Next Article in Journal
Role of Pascalization in Milk Processing and Preservation: A Potential Alternative towards Sustainable Food Processing
Next Article in Special Issue
Review on All-Fiber Online Raman Sensor with Hollow Core Microstructured Optical Fiber
Previous Article in Journal
Thermo-Optic Phase Tuners Analysis and Design for Process Modules on a Silicon Nitride Platform
Previous Article in Special Issue
A Broadband Gold-Coated Photonic Crystal Fiber Polarization Filter with a High Loss Ratio of Both Polarizations at 1550 and 1310 nm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chalcogenide Glass Microfibers for Mid-Infrared Optics

1
State Key Laboratory of Modern Optical Instrumentation, College of Optical Science and Engineering, Zhejiang University, Hangzhou 310027, China
2
Intelligent Optics & Photonics Research Center, Jiaxing Institute of Zhejiang University, Jiaxing 314031, China
3
Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan 030006, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this paper.
Photonics 2021, 8(11), 497; https://doi.org/10.3390/photonics8110497
Submission received: 5 October 2021 / Revised: 3 November 2021 / Accepted: 3 November 2021 / Published: 5 November 2021
(This article belongs to the Special Issue Novel Specialty Optical Fibers and Applications)

Abstract

:
With diameters close to the wavelength of the guided light, optical microfibers (MFs) can guide light with tight optical confinement, strong evanescent fields and manageable waveguide dispersion and have been widely investigated in the past decades for a variety of applications. Compared to silica MFs, which are ideal for working in visible and near-infrared regions, chalcogenide glass (ChG) MFs are promising for mid-infrared (mid-IR) optics, owing to their easy fabrication, broad-band transparency and high nonlinearity, and have been attracting increasing attention in applications ranging from near-field coupling and molecular sensing to nonlinear optics. Here, we review this emerging field, mainly based on its progress in the last decade. Starting from the high-temperature taper drawing technique for MF fabrication, we introduce basic mid-IR waveguiding properties of typical ChG MFs made of As2S3 and As2Se3. Then, we focus on ChG-MF-based passive optical devices, including optical couplers, resonators and gratings and active and nonlinear applications of ChG MFs for mid-IR Raman lasers, frequency combs and supercontinuum (SC) generation. MF-based spectroscopy and chemical/biological sensors are also introduced. Finally, we conclude the review with a brief summary and an outlook on future challenges and opportunities of ChG MFs.

1. Introduction

In the past decades, optical microfibers (MFs) have brought wide opportunities in renewing and expanding fiber optics and technology at a wavelength scale [1,2,3,4,5]. Fabricated by a high-temperature taper-drawing technique [6], MFs exhibit excellent surface smoothness and diameter uniformity, which bestow them favorable features of low waveguiding loss [1,7], tight optical confinement, high fractional evanescent fields and large manageable waveguide dispersion [8], making them a versatile platform for both scientific research (e.g., optical nonlinearity [9,10,11] and atom optics [12,13,14]) and technological applications (e.g., optical sensors [15,16,17] and microlasers [18,19,20]). As one of the most widely studied optical materials, MFs based on silica have been employed for a variety of applications [3,4]; however, their uses in the mid-infrared (mid-IR, 2.5–20 μm) spectral range are restricted by strong absorption of silica glass [21]. In recent years, along with the rapid progress in mid-IR photonics, MFs for mid-IR photonics have attracted increasing interest [22,23].
Typically, mid-IR MFs are fabricated from a category of mid-IR-transparent materials, including oxide glasses (e.g., germanates [24], fluorotellurites [25] and tellurites [26]), fluoride glasses (ZBLAN) [27,28], chalcogenide glasses (ChGs, glasses containing one or more chalcogens: sulfur (S), selenium (Se) and tellurium (Te)) [29,30,31], as well as semiconductors (e.g., cadmium telluride (CdTe) and [32] silicon (Si) [33]). Among these materials, ChGs are mostly investigated for their special merits, including broadband intrinsic transparency (0.5–25 μm), high optical nonlinearity (about 100–1000 times larger than that of silica glass) and hospitality to rare-earth dopants [29,34,35]. The typical optical loss of ChG fibers in the mid-IR is around 0.1–10 dB/m [35], while the lowest loss, down to 0.012 dB/m at 3 μm, has been reported in multimode As2S3 fibers [36]. Additionally, photonic devices based on ChG MFs, including resonators [37,38], gratings [39,40,41] and sensors [42,43], have been investigated in the visible (VIS) or near-infrared (near-IR) region and have recently extended to the mid-IR region [44,45,46]. In this review, combined with the latest progress in ChG MFs, we first introduce the fabrication and optical waveguiding properties of ChG MFs. Secondly, we introduce ChG MF-based passive optical devices (e.g., couplers, resonators and gratings), as well as active and nonlinear applications of ChG MFs in the mid-IR region including Raman lasers, frequency combs and supercontinuum (SC) generation, followed by MF-based spectroscopy and chemical/biological sensors. Finally, we summarize this review and present a brief outlook into future challenges and opportunities in this field.

2. Fabrication

Excellent surface smoothness and geometric uniformity are regarded as essential guarantees for achieving low-loss optical waveguiding in MFs. Typically, for MFs made of glass, the high-temperature taper-drawing process is the best technique for fabricating high-quality MFs with extremely low surface roughness (e.g., <0.5 nm [47]), excellent diameter uniformity and circular cross-section [48]. Tapered fibers based on mid-IR materials such as tellurite glass and fluorotellurite glass have been successfully fabricated [25,26]. Besides, owing to the relatively low softening temperature (Ts, e.g., 180 °C for As2S3 glass) of ChGs, ChG MFs are typically fabricated from ChG fibers heated by an electric heater (several millimeters in heating zone length) at a temperature slightly higher than the Ts [31,49], as schematically illustrated in Figure 1a. When stretching the fiber bi-directionally at both sides, the fiber is elongated and forms a waist (i.e., the MF) with two transition tapers connected to the untapered fibers. Close-up scanning electron microscope (SEM) images of an as-fabricated ChG (As2S3 glass) MF are shown in the inset of Figure 1a, showing the excellent surface smoothness and diameter uniformity.
Usually, made from soft glass, as-fabricated ChG MF is relatively fragile and easily damaged by mechanical disturbance. To enhance the mechanical strength of ChG MF and reduce its sensitivity to the surrounding environment, in 2010, Baker et al. reported a kind of hybrid ChG-PMMA fiber taper consisting of an As2Se3 core and a PMMA cladding [31]. The PMMA cladding possesses a Ts compatible with As2Se3 glass and enables taper-drawing fabrication of an MF from the two materials at the same time. A 9.7-cm-length hybrid biconical As2Se3 fiber taper with a core diameter down to 0.8 μm and a PMMA cladding diameter of 2.4 μm was fabricated successfully. In 2011, they demonstrated a generalized heat-brush tapering approach, which allowed the ratio of the feed and drawing velocities changes within each tapering sweep [50]. The waist diameter of the as-fabricated As2Se3 fiber taper decreased linearly from 15 to 10 µm over a 2.0-centimeter length. In 2016, Li et al. fabricated polymer-cladded (PC, COP or PMMA) As2Se3 fiber tapers with a waist diameter down to 1.5 µm and a waist length of 10 cm [49], as illustrated in Figure 1b. Both ends of the ChG fiber tapers were polished and adhered to single-mode silica fibers using UV-cured epoxy for efficient light input and output. Owing to the large waveguide nonlinear parameter and engineerable chromatic dispersion of the MFs, nonlinear optical effects including self-phase modulation, four-wave mixing and Raman scattering with wavelength up to 2.2 µm could be achieved.
To ease the tapering system and obtain long MFs with uniform diameters with a large reduction ratio of fiber diameter between the preform fiber and the final MF, in 2020, Xie et al. demonstrated a two-step taper-drawing process [45]. As schematically shown in Figure 1c1, commercial As2S3 fiber was first heated and stretched bi-directionally with a low speed (0.4 mm·s−1). Secondly, the pre-drawn MF with a diameter in several micrometers was stretched unidirectionally at a fast rate (6 mm·s−1) at one end and a slow rate (0.4 mm·s−1) at the other end to draw a long-waist MF with a uniform diameter. Biconical fiber tapers with a waist length longer than 5 cm and diameters of around 3 μm were easily fabricated for assembling mid-IR photonic devices. In addition, as-drawn As2S3 MF (3.5-μm diameter) could be assembled into a 62-μm diameter knot structure with an elastic strain of up to 5.6% (Figure 1c2), indicating the excellent structural uniformity and surface-defect-free condition of the MF.

3. Optical Waveguiding Properties

MFs offer a number of attractive optical properties including tight optical confinement, strong evanescent field and tailorable waveguide dispersion [4,8]. The special merits of ChGs bestow the ChG MFs more possibilities. In this section, we discuss the optical properties of As2S3 (transmission range: 0.7–6 μm [35]) and As2Se3 (transmission range: 1–10 μm [35]) MFs contributed to applications such as optical waveguiding and nonlinearity. We assumed that the MF has a circular cross-section and an infinite air cladding with a step-index profile, and only the fundamental mode (HE11) was considered. The refractive indices of the As2S3 (As2Se3) glass around 1.55-μm and 4-μm wavelengths are 2.44 (2.73) and 2.41 (2.68), respectively [51,52]. The refractive index of As2Se3 glass around 8-μm wavelength is 2.67.
The effective refractive index (neff = β/k0, β is the propagation constant of the guided mode, k0 = 2π/λ is the wavenumber of the transmission wavelength λ), effective mode area (Aeff) and fractional evanescent fields (η) are essential waveguiding parameters for designing fiber-optic devices, including near-field couplers, gratings, resonators and sensors [31,53]. Figure 2a–c shows the calculated As2S3 and As2Se3 MF diameter-dependent neff, Aeff, and η in the mid-IR region, respectively. The results in the 1.55-μm wavelength (dash lines) are also shown for comparison. ChG MFs with diameters larger than the wavelengths of the guided light exhibit neff of HE11 mode close to the refractive indices of the bulk materials (Figure 2a), which is independent on the guided wavelength. When the diameter reduces close to or smaller than the wavelength, the neff decreases dramatically, indicating more fractional evanescent fields exist around the MF (Figure 2b), beneficial for enhancing interactions between the mid-IR light and samples surrounding the MF in applications such as optical trapping, mid-IR spectroscopy and chemical/biological sensing related to molecular fingerprints [3,17,54]. Meanwhile, owing to the high nonlinear refractive index n2 of As2S3 and As2Se3 materials and tight optical confinement (i.e., small Aeff) of the guided modes in MFs, these MFs usually have large effective nonlinearity γ (γ = k0 n2/Aeff) and relatively low-threshold nonlinear optical effects when their diameters are close to the wavelength of the guided light in the material (i.e., λ/n, λ and n are the wavelength in air and refractive index of material, respectively. See Figure 2c) [31,55].

4. Photonic Applications

4.1. Near-Field Optical Couplers

MF near-field couplers are simple fiber-optic components that can be used for efficient in/out-coupling with other optical structures such as micro/nano waveguides [48,56,57,58], whispering gallery mode (WGM) microcavities [59,60] and photonic crystal structures [61,62,63]. By precisely adjusting the coupling length or gap distance between the MF and the photonic structure, the phase-matching condition (and thus the coupling efficiency) can be precisely adjusted. MF near-field couplers possess the advantages of easy operation and small footprints.
With increasing demands of in/out coupler at mid-IR range in recent years, mid-IR MFs have been applied in the near-field optical coupling of a variety of photonic structures, including mid-IR waveguides and microcavities. For example, in 2018, to test the waveguiding properties of CdTe microwires (a quasi-one-dimensional single-crystal semiconductor) in the mid-IR region, Xin et al. used a 1.5-μm-diameter As2S3 MF as input coupler and a 5-μm-diameter As2S3 MF as output coupler for a 1.5-μm-diameter CdTe microwire (Figure 3a) [32]. The input MF diameter was carefully chosen to ensure single-mode operation within the 4.45–4.7 μm range. The relatively large effective refractive index contrast between the As2S3 MF and CdTe microwire can decrease the end-face reflection and achieve flat broadband coupling over the 4.45–4.7 μm and 8.36–8.6 μm range.
High-quality factor (Q, defined as the ratio of the resonant wavelength to the full width at half-maximum (FWHM)) resonators in the mid-IR are critical to promote the performances of mid-IR laser sources, cavity-based sensors and mid-IR frequency comb [66]. However, achieving ultra-high Q cavities and measuring their quality factors in the mid-IR remain challenging. In 2016, Lecaplain et al. developed an efficient coupling technique for WGM-based fluoride crystalline microresonators assisted by tapered ChG fibers (Figure 3b1) [64]. The ChG tapered fibers were pulled down to the 1-μm diameter (i.e., a ChG MF) to achieve phase matching with the fundamental WGMs of the resonator. By translating the microresonator along the taper relative to the waist, the phase-matching conditions were tuned precisely. Figure 3b2 shows the normalized transmission as a function of the coupling parameter. The data were consistent with the theoretical model ( K = κ e x / κ 0 = ( 1 ± T ) / ( 1 T ) , in which κ 0 is the intrinsic loss rate, κ e x is the coupling loss rate and T is the transmission). More recently, Xie et al. used As2S3 biconical tapered fibers for the first characterization of ChG microsphere resonators in the mid-IR region (Figure 3c) [65]. A 1.6-μm-diameter As2S3 fiber taper was used for near-field coupling with ChG microspheres with the coupling loss less than 3 dB over 4.465–4.705 μm.

4.2. MF-Based Resonators

The combination of low-loss waveguiding and high-efficiency evanescent coupling between adjacent MFs bestows a kind of competitive high-Q resonators in various structures (e.g., loop [67,68], knot [1,69], ring [70,71,72] and coil [73,74]), showing the desirable features of easy fabrication, fiber compatibility and resonance tunability. During the past two decades, silica MF-based resonators have attracted wide attention in the VIS and near-IR applications, including optical filters, sensors and lasers [57,74,75,76]. Additionally, there are several kinds of high-Q (104–105) resonators based on ChG MFs in the near-IR region, including SU8 polymer-embedded As2S3 MF knot resonators (MKRs) (Figure 4a1,a2) [38], self-touching As2Se3 MF loop resonators (MLRs) by thermally shaping As2Se3 glass into an MF and splicing to cleaved silica fiber tapers (Figure 4b1,b2) [37] and resonators based on photoinduced WGMs localized in the cross-section of the ChG MFs [77]. These studies are beneficial to the further exploitations of ChGs, but they were not extended to the mid-IR region.
In 2020, Xie et al. demonstrated the first ChG MF-based resonator in the mid-IR region [45]. As2S3 MFs with uniform diameters and long waists were fabricated via a two-step taper drawing process (Figure 1c1). The MF structure was immersed in a drop of ethyl alcohol to avoid possible surface damage and assembled to a knot with double ends naturally connected to fiber tapers. Figure 4c1 shows an 824-μm diameter MKR assembled from a 3.2-μm-diameter As2S3 MF. A Q factor of about 2.84 × 104 was obtained at the wavelength of 4469.14 nm (Figure 4c2,c3). The free spectral range (FSR) of the MKR can not only be tuned by tightening the knot structure in liquid (Figure 4c4), but the resonance peaks can also be thermally tuned at a thermal tuning rate of 110 pm·°C−1 (Figure 4c5). Furthermore, to increase the long-term stability of the MKR device and isolate it from contaminations, a 551-μm-diameter MKR was embedded in low-index polymer (PMMA) film while a Q factor of 1.1 × 104 around 4.5-μm wavelength was retained.
As MF-based mid-IR resonators are essential for MF-based optical technology, they are expected to attract increasing attention in mid-IR optical applications, including mid-IR microlasers, narrow-wide filters, cavity-enhanced spectroscopy and the label-free detection of molecules [54,78].

4.3. MF Gratings

MF gratings have attracted intense attention owing to their miniaturized footprints, large fraction of evanescent field and high flexibility. Typical grating fabrication techniques include femtosecond (fs) laser inscription [79,80], ultraviolet laser irradiation [81] and focused ion beam milling [82,83]. So far, MF gratings have been employed for photonic sensors for measuring the refractive index, force and temperature, as well as for functionalized devices, including tunable filters and Fabry–Perot cavities [84,85,86,87].
Previously, ChG fiber gratings working in the mid-IR region [88] and ChG MF gratings working in the near-IR have been reported [40,89] (Figure 5a), while mid-IR MF gratings have been seldomly studied. In 2020, Cai et al. demonstrated the first mid-IR MF Bragg grating (mFBG) based on an As2S3 MF (Figure 5b1) [46]. The transmission spectrum shows a 15 dB depth around 4.5-micrometer wavelength (Figure 5b3), indicating a photo-induced refractive index change of 0.02. The dependence of the grating formation on the accumulated influence of exposure power density and time was also investigated. The interference pattern exposure technique can be applied to other photosensitive ChG MFs.
Compared with conventional long-period grating (LPG), MFs could increase the depth of the evanescent field into the ambient environment, which may be helpful to enhance the performance of MF-based optical sensors. In 2019, Wang et al. reported a temperature sensor based on long-period fiber gratings (LPFGs) inscribed on tapered multimode ChG fibers [90]. Simulation results show that the temperature sensitivity could be increased by reducing the waist diameter of tapered fiber. A resonant wavelength at 3 μm was obtained when the period of LPFGs was 176 μm and the maximum sensitivity of temperature could achieve 12.6 nm/°C at a 3-micrometer wavelength.

4.4. Raman Lasers

The Raman gain provided by silica fibers has led to Raman fiber lasers in the near-IR range for optical imaging and telecommunications-related applications [91]. The mid-IR Raman fiber lasers have been applied in LIDAR, pharmaceutical science and military applications [92,93]. The Raman gain coefficients of typical ChGs (e.g., 4.3~5.7 × 10−12 m/W for As2S3 and 2~5 × 10−11 m/W for As2Se3 at 1.5 μm [94]) are more than 900 times larger than that of silica, making them excellent candidates for the mid-IR Raman lasers. In 2017, Abdukerim et al. demonstrated an all-fiber Raman laser based on a PMMA-coated 10-centimeter As38Se62/As38S62 MF [95]. The Raman laser emitted at 2.025 μm with pump pulses at 1.938 μm and a threshold pump power of 4.6 W (Figure 6a). To generate a shower of Raman solitons at the same wavelength, in 2021, Guo et al. pumped a 1960-nanometer picosecond laser into a tapered fluorotellurite MF and demonstrated Raman solitons around 3 μm [96]. The group velocity dispersion was controlled by the diameter of the fluorotellurite MF, and all the generated solitons could be put at the same wavelength.

4.5. Frequency Conversion

Besides the large nonlinear optical coefficient and broad mid-IR transparency, ChGs are excellent candidates for mid-IR frequency conversion (e.g., optical parametric oscillation (OPO) and optical parametric amplification (OPA)) for their high laser damage threshold and moderate birefringence [100]. The efficiency of frequency conversion depends on phase matching, nonlinear gain and losses. The tapering of ChG fibers to increase nonlinearity while compensating the normal material dispersion has been demonstrated. In 2016, Abdukerim et al. reported the first fiber optical parametric oscillator (FOPO) using an As2Se3-COP MF at 2 μm based on four-wave mixing (FWM) [101]. The MF with a core diameter of 1.47 μm led to a waveguide nonlinearity of 24 W−1m−1 and a zero-dispersion wavelength of 1.875 μm. The FOPO had a pump threshold power of 5 W and covered a tunable wavelength range of 290 nm. In the FWM process, where pump and idler are far-detuned, the exact phase-matching wavelength is highly sensitive to chromatic dispersion and hard to control. In 2019, Alamgir et al. demonstrated the far-detuned wavelength converter by in-situ monitoring the output wavelength while tapering down the As2Se3 fiber [97]. As the tapering proceeded, the phase-matching condition was satisfied closer to the pump wavelength and the gain spectrum gradually shifted towards shorter wavelengths. The idlers were generated with a spectral range of 2.347–2.481 μm from a pump wavelength of 1.938 μm (Figure 6b).
The modulation instability (MI) is another frequency conversion process where the low-amplitude noise is parametrically amplified on the pump signal together with the growth of symmetric sidebands on both sides of the pump. Although most of the spontaneous MI requires pumping in the anomalous dispersion regime, MI can also be pumped in a normal dispersion regime when the fiber possesses a higher-order group velocity dispersion profile. To prove the broadband frequency conversion due to normal dispersion MI, in 2014, Godin et al. demonstrated a parametric frequency conversion based on MI at 2 and 3.5 μm by pumping an As2Se3-polymer MF with an fs OPO at 2.63 μm [98]. The measured frequency shift of 30 THz was the largest reported, using normal dispersion pumped MI in a single-pass configuration (Figure 6c). To prove the potential for improving the technical specifications of mid-IR fiber wavelength converter in terms of tunability and conversion efficiency, in 2017, Li et al. generated tunable parametric sidebands via MI in a 10-centimeter long As2Se3-CYTOP tapered MF with a diameter of 1.625 μm (Figure 6d) [99]. The widely spaced Stokes and anti-Stokes bands were generated with a frequency shift as large as 49.3 THz, the largest reported in soft glass materials.

4.6. Supercontinuum Generation

Supercontinuum (SC) generation in optical fibers originates from the interaction between ultrashort laser pulses and fibers with high nonlinearity [107]. ChGs are regarded as good candidates for mid-IR SC generation with a large transparency window and comparatively high optical nonlinearity [22].
The uses of MFs reduce the power consumption and allow a shorter interaction length for SC generation. In 2013, Al-kadry et al. demonstrated SC generation from 1260 to 2200 nm using a 10-centimeter long As2Se3 MF [108]. The pump wavelength was self-frequency shifted to 1775 nm in the anomalous dispersion regime of the As2Se3 MF to avoid two photons’ absorption. In the same year, Rudy et al. reported a generated SC from 1 to 3.7 μm in a 2.1-millimeter length As2S3 MF with 300-pJ 2-nanometer ultrafast pump pulses [109]. The spectral bandwidth was limited by the absorption caused by third harmonic generation. In 2014, Al-kadry et al. further used pump pulses of 800 fs at a 1.94-μm wavelength to generate an SC covered bandwidth of two octaves from 1.1 to 4.4 μm [102]. The waveguide nonlinearity of a 1.6-micrometer diameter As2Se3 MF was as high as 32.2 W−1m−1, resulting in a pumping power of only 500 pJ (Figure 7a). In 2017, Wang et al. investigated the dependence of the SC spectral behavior on the waist diameter and the transition region length of the tapered MF (Figure 7b) [103]. They reported a 1.4–7.2-micrometer SC generation with an average power of 1.06 mW from a 12-cm-length As-S tapered MF pumped at 3.25 μm.
Tapering the fiber is also an important method to obtain a highly coherent SC by designing an all-normal dispersion of fibers to suppress soliton effects including soliton fission and Raman soliton self-frequency shift. In 2019, Li et al. reported a 1.5–8.3 μm SC generation with a high coherence property from As-S tapered MF pumped at 3.75 μm (Figure 7c) [104]. They found that the coherence properties gradually become worse when the waist core diameters of tapered MFs decrease and a flat and near-zero dispersion is beneficial for high-coherence SC generation.
To meet the demanding specifications including robust and alignment-free sources of broadband SC, the multimaterial all-fiber-based SC systems have drawn much attention. In 2015, Sun et al. fabricated multimaterial As2Se3-As2S3 ChG MFs with a zero-dispersion wavelength less than 2 μm [110]. The SC generation from 1.4 μm to longer than 4.8 μm was demonstrated by pumping the MF with 100 fs pulses at 3.4 μm. In 2017, Hudson et al. reported a 2.4 octave-SC generation by pumping a robust polymer-protected As2Se3-As2S3 MF with 230 fs, 4.2 kW peak power pulses at 3 μm (Figure 7d) [105]. An average power spectral density of 0.003 mW/nm was achieved. Furthermore, significant power existed at various octave points throughout the SC, making the system a good candidate for an f-2f interferometer.
It is worth mentioning that another type of ChG MFs, tapered fibers with suspended-core geometry that operate in the anomalous dispersion regime, have been used to generate SC within a broad spectrum. In 2018, Anashkina et al. designed and developed As39Se61 suspended-core MFs for mid-IR SC generation. SC generation ranging from 1 to 10 μm was proposed with 150-fs 100-pJ pump pulses at 2 μm [111]. In 2020, Leonov et al. reported SC generation covering 1.4 to 4.2 μm in As39S61 suspended-core MFs pumped by a Cr:ZnSe laser. A maximum SC average power of ~35 mW was obtained with an optical efficiency of ~35% [112].
In addition, materials other than ChGs have also been investigated for SC generation. For example, compared to that of ChG, telluride glasses have shorter zero-dispersion wavelengths that are desirable for near-to-mid IR SC generation. In 2019, Saini et al. obtained an SC spectrum spanning 1.28–3.31 μm using a 3.2-centimeter-long tellurite tapered MF pumped with 200 fs laser pulse at 2.0 μm [26]. Compared to fluoride and chalcogenide fibers, fluorotellurite fibers based on TeO2-BaF2-Y2O3 have better chemical and thermal properties. In 2019, Li et al. demonstrated broadband SC generation from 0.6 to 5.4 μm in a fluorotellurite tapered MF pumped by a 2010 nm fs pulses [25]. Recently, polysilicon waveguides have emerged as an alternative platform as they are cheap and flexible to be incorporated into a wide range of geometries. In 2019, Ren et al. fabricated the tapered polycrystalline silicon core MFs (SCFs), and an SC spanning from 1.8 to 3.4 μm was obtained [33].

4.7. Frequency Combs

In the past years, frequency combs have been studied from the VIS and near-IR to the mid-IR regions. Mid-IR combs are of great use for molecular fingerprint spectroscopy and are typically produced by OPOs [113,114], difference frequency generation [115,116], microresonators [117,118] and fs laser generation [119]. Wide frequency combs spectra in mid-IR have also been demonstrated using PPLN-based OPOs pumped by fiber-laser-based near-IR combs [120]. The mid-IR MFs could further broaden the frequency combs by SC generation when the generated SC maintains coherence with the pump laser. In 2012, Marandi et al. used a subharmonic OPO and subsequent SC generation in the tapered MF to convert the pulses of a conventional 1.5-micrometer frequency comb source to mid-IR [106]. The spectrum of the MF output extended from 2.2 to 5 μm at 40 dB below the peak (Figure 7e1). The beat frequencies in Figure 7e2 verified that the coherence properties of the initial frequency combs presevered. It is worth mentioning that, to simplify the mid-IR frequency comb generation, in 2014, Lee et al. demonstrated mid-IR frequency comb generation using a hybrid ChG-silica nanospike waveguide [121]. The waveguide was pumped directly by a two-micrometer frequency comb and could minimize the unwanted dispersion of pump pulses. The waveguide was small and simple using free-space coupling into and out of the fiber without an intermediate OPO.

4.8. MF-Based Sensors

The optical fiber sensors operating in the mid-IR region where the vibrational and rotational states of molecules are located play an important role in biomedical [122], chemical [123] and environmental applications [124].
The principle of fiber evanescent wave spectroscopy (FEWS) is based on the interaction between evanescent fields and chemical or biological specimens to be investigated, showing desirable advantages of analyzing samples in situ in real time. In past decades, many FEWS experiments have been applied in different areas including the detection of contaminants in water [128], metabolic profiling of chronic diseases [129] and monitoring of chemical processes [130]. The characterization of human tissue based on mid-IR FEWS is an important method for early and rapid diagnosis of diseases including tumors or cancers [131]. Typically, the fiber diameter used for FEWS is much larger than the working wavelength, leaving much space for sensitivity enhancement.
Owing to the characteristic fingerprints of a wide range of gases (e.g., CO, CO2, NO and CH4) falling in the mid-IR region, MF-based gas sensors have been proposed based on mid-IR absorption spectroscopy in the last decade [132]. In 2019, Huang et al. demonstrated a nonlinear gas sensor based on third harmonic generation (THG) in cascaded ChG MFs for detecting CH4 (Figure 8a1) [125]. Mid-IR fingerprint light was absorbed in the first section of MF and the rest transmission signal was converted into a near-IR signal after THG at the second section. The proposed mid-IR ChG MF CH4 sensor had a theoretical detection limit as low as 7.4 × 10−8 (Figure 8a2). Another configuration combines graphene with MFs, which can provide high-sensitivity optical gas sensing from permittivity change of the graphene. In 2019, Wang et al. reported a mid-IR gas sensor based on the graphene Bragg grating integrated with the Si slot MF (Figure 8b1) [126]. The NO2 concentration could be determined from the 3 dB bandwidth of the Bragg grating’s reflection band (Figure 8b2). A detection limit below 1 ppm and a sensitivity of 1.02 nm/ppm were achieved.
The optimization of the tapered transition structure of the fiber sensor is essential to obtain high-efficiency excitation of the evanescent fields to further enhance the sensor sensitivity. In 2021, Wang et al. reported a geometrical optimization of Ge-Sb-Se tapered MF for concentration sensing of methanol (Figure 8c1,c2) [127]. The highest sensitivity of 0.0120 a.u./% was obtained from the MF with a down-taper transition length of 5.4 mm and an up-taper transition length of 7.9 mm. Meanwhile, Wang et al. from the same group investigated the dependence of the sensitivity (of aqueous ethanol solution) on the MF parameters (e.g., diameter and length) [133] and predicted that the sensitivity could be improved by decreasing the MF diameter and increasing the length of the taper waist.

5. Discussion

In this article, a comprehensive survey of recent works on ChG MFs for the mid-IR range is presented, where the fabrication methods and the optical waveguiding properties (i.e., neff, Aeff and η) of ChG MFs are discussed. So far, high-quality ChG MFs (e.g., As2S3 and As2Se3 MFs) have been fabricated by a high-temperature taper drawing process and have shown great promise for broadband low-loss optical waveguiding in the mid-IR region. Based on these MFs, a new category of compact and fiber-compatible photonic devices has been developed for passive, active and nonlinear applications in the mid-IR region. These initial results have unambiguously suggested the great opportunities of using the ChG MF as a versatile platform for a miniaturized fiber-based platform for mid-IR optics.

6. Conclusions

Looking into the future, there are obvious challenges and opportunities for ChG MFs. First, compared with silica glass, ChGs possess a much higher refractive index and optical nonlinearity that are advantageous for tighter optical confinement and lower threshold for nonlinear optical effects but exhibit much lower stability against optical, thermal, chemical and mechanical damages. Therefore, better package or protection techniques/protocols should be developed for real applications of ChG MFs, and more research efforts could be made on the material aspects. Secondly, compared to those for the VIS and near-IR spectral ranges, fiber-compatible devices for mid-IR, ranging from light sources, couplers, polarizers and filters to amplifiers, are typically not mature or optimized, leaving both challenges and opportunities in the future study of mid-IR MFs. Finally, other potentials, such as MF-incorporated plasmonic and atom optics that have been quite successful in the VIS/near-IR region, have not yet been equivalently investigated in mid-IR, which could be the new driving force in this field.

Author Contributions

Conceptualization, X.G., P.W. and L.T.; writing—original draft preparation, D.C. and Y.X.; writing—review and editing, D.C., Y.X., X.G., P.W. and L.T.; supervision, L.T.; project administration and funding acquisition, L.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (11527901), Natural Science Foundation of Zhejiang Province (LR21F050002), and Fundamental Research Funds for the Central Universities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tong, L.M.; Gattass, R.R.; Ashcom, J.B.; He, S.L.; Lou, J.Y.; Shen, M.Y.; Maxwell, I.; Mazur, E. Subwavelength-diameter silica wires for low-loss optical wave guiding. Nature 2003, 426, 816–819. [Google Scholar] [CrossRef] [PubMed]
  2. Brambilla, G.; Xu, F.; Horak, P.; Jung, Y.M.; Koizumi, F.; Sessions, N.P.; Koukharenko, E.; Feng, X.; Murugan, G.S.; Wilkinson, J.S.; et al. Optical fiber nanowires and microwires: Fabrication and applications. Adv. Opt. Photonics 2009, 1, 107–161. [Google Scholar] [CrossRef]
  3. Tong, L.M.; Zi, F.; Guo, X.; Lou, J.Y. Optical microfibers and nanofibers: A tutorial. Opt. Commun. 2012, 285, 4641–4647. [Google Scholar] [CrossRef]
  4. Wu, X.Q.; Tong, L.M. Optical microfibers and nanofibers. Nanophotonics 2013, 2, 407–428. [Google Scholar] [CrossRef]
  5. Ismaeel, R.; Lee, T.; Ding, M.; Belal, M.; Brambilla, G. Optical microfiber passive components. Laser Photonics Rev. 2013, 7, 350–384. [Google Scholar] [CrossRef] [Green Version]
  6. Ward, J.M.; O’Shea, D.G.; Shortt, B.J.; Morrissey, M.J.; Deasy, K.; Nic Chormaic, S.G. Heat-and-pull rig for fiber taper fabrication. Rev. Sci. Instrum. 2006, 77, 083105. [Google Scholar] [CrossRef] [Green Version]
  7. Xu, Y.X.; Fang, W.; Tong, L.M. Real-time control of micro/nanofiber waist diameter with ultrahigh accuracy and precision. Opt. Express 2017, 25, 10434–10440. [Google Scholar] [CrossRef] [PubMed]
  8. Tong, L.M.; Lou, J.Y.; Mazur, E. Single-mode guiding properties of subwavelength-diameter silica and silicon wire waveguides. Opt. Express 2004, 12, 1025–1035. [Google Scholar] [CrossRef] [PubMed]
  9. Foster, M.A.; Moll, K.D.; Gaeta, A.L. Optimal waveguide dimensions for nonlinear interactions. Opt. Express 2004, 12, 2880–2887. [Google Scholar] [CrossRef] [Green Version]
  10. Foster, M.A.; Turner, A.C.; Lipson, M.; Gaeta, A.L. Nonlinear optics in photonic nanowires. Opt. Express 2008, 16, 1300–1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Leon-Saval, S.G.; Birks, T.A.; Wadsworth, W.J.; Russell, P.S.J.; Mason, M.W. Supercontinuum generation in submicron fibre waveguides. Opt. Express 2004, 12, 2864–2869. [Google Scholar] [CrossRef]
  12. Balykin, V.I.; Hakuta, K.; Le Kien, F.; Liang, J.Q.; Morinaga, M. Atom trapping and guiding with a subwavelength-diameter optical fiber. Phys. Rev. A 2004, 70, 011401. [Google Scholar] [CrossRef] [Green Version]
  13. Le Kien, F.; Balykin, V.I.; Hakuta, K. Atom trap and waveguide using a two-color evanescent light field around a subwavelength-diameter optical fiber. Phys. Rev. A 2004, 70, 063403. [Google Scholar] [CrossRef] [Green Version]
  14. Sague, G.; Vetsch, E.; Alt, W.; Meschede, D.; Rauschenbeutel, A. Cold-atom physics using ultrathin optical fibers: Light-induced dipole forces and surface interactions. Phys. Rev. Lett. 2007, 99, 163602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Guo, X.; Ying, Y.B.; Tong, L.M. Photonic nanowires: From subwavelength waveguides to optical sensors. Acc. Chem. Res. 2014, 47, 656–666. [Google Scholar] [CrossRef] [PubMed]
  16. Lou, J.Y.; Wang, Y.P.; Tong, L.M. Microfiber optical sensors: A review. Sensors 2014, 14, 5823–5844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhang, L.; Tang, Y.; Tong, L.M. Micro-/nanofiber optics: Merging photonics and material science on nanoscale for advanced sensing technology. IScience 2020, 23, 100810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Yang, Q.; Jiang, X.S.; Guo, X.; Chen, Y.; Tong, L.M. Hybrid structure laser based on semiconductor nanowires and a silica microfiber knot cavity. Appl. Phys. Lett. 2009, 94, 101108. [Google Scholar]
  19. Jiang, X.S.; Song, Q.H.; Xu, L.; Fu, J.; Tong, L.M. Microfiber knot dye laser based on the evanescent-wave-coupled gain. Appl. Phys. Lett. 2007, 90, 233501. [Google Scholar] [CrossRef]
  20. Jiang, X.S.; Yang, Q.; Vienne, G.; Li, Y.H.; Tong, L.M.; Zhang, J.J.; Hu, L.L. Demonstration of microfiber knot laser. Appl. Phys. Lett. 2006, 89, 143513. [Google Scholar] [CrossRef]
  21. Nagel, S.R.; MacChesney, J.B.; Walker, K.L. Optical Fiber Communications, 1st ed.; Elsevier: New York, NY, USA, 1985; pp. 1–64. [Google Scholar]
  22. Dai, S.X.; Wang, Y.Y.; Peng, X.F.; Zhang, P.Q.; Wang, X.S.; Xu, Y.S. A review of mid-Infrared supercontinuum generation in chalcogenide glass fibers. Appl. Sci. 2018, 8, 707. [Google Scholar] [CrossRef] [Green Version]
  23. Gao, S.; Bao, X.Y. Chalcogenide taper and its nonlinear effects and sensing applications. IScience 2020, 23, 100802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Anashkina, E.A.; Andrianov, A.V.; Kim, A.V. Conversion of ultrashort laser pulses to wavelengths above 3 mm in tapered germanate fibres. Quantum Electron. 2015, 45, 437–442. [Google Scholar] [CrossRef]
  25. Li, Z.R.; Yao, C.F.; Jia, Z.X.; Wang, F.; Qin, G.S.; Ohishi, Y.; Qin, W.P. Broadband supercontinuum generation from 600 to 5400 nm in a tapered fluorotellurite fiber pumped by a 2010 nm femtosecond fiber laser. Appl. Phys. Lett. 2019, 115, 091103. [Google Scholar] [CrossRef]
  26. Saini, T.S.; Hoa, N.P.T.; Tuan, T.H.; Luo, X.; Suzuki, T.; Nish, Y. Tapered tellurite step-index optical fiber for coherent near-to-mid-IR supercontinuum generation: Experiment and modeling. Appl. Opt. 2019, 58, 415–421. [Google Scholar] [CrossRef] [PubMed]
  27. Kubat, I.; Agger, C.S.; Moselund, P.M.; Bang, O. Mid-infrared supercontinuum generation to 45 μm in uniform and tapered ZBLAN step-index fibers by direct pumping at 1064 or 1550 nm. J. Opt. Soc. Am. B 2013, 30, 2743–2757. [Google Scholar] [CrossRef]
  28. Chen, Z.G.; Taylor, A.J.; Efimov, A. Coherent mid-infrared broadband continuum generation in non-uniform ZBLAN fiber taper. Opt. Express 2009, 17, 5852–5860. [Google Scholar] [CrossRef] [PubMed]
  29. Zakery, A.; Elliott, S.R. Optical properties and applications of chalcogenide glasses: A review. J. Non-Cryst. Solids 2003, 330, 1–12. [Google Scholar] [CrossRef]
  30. Eggleton, B.J.; Luther-Davies, B.; Richardson, K. Chalcogenide photonics. Nat. Photonics 2011, 5, 141–148. [Google Scholar] [CrossRef]
  31. Baker, C.; Rochette, M. Highly nonlinear hybrid AsSe-PMMA microtapers. Opt. Express 2010, 18, 12391–12398. [Google Scholar] [CrossRef]
  32. Xin, C.G.; Wu, H.; Xie, Y.; Yu, S.L.; Zhou, N.; Shi, Z.X.; Guo, X.; Tong, L.M. CdTe microwires as mid-infrared optical waveguides. Opt. Express 2018, 26, 10944–10952. [Google Scholar] [CrossRef] [PubMed]
  33. Ren, H.; Shen, L.; Wu, D.; Aktas, O.; Hawkins, T.; Ballato, J.; Gibson, U.J.; Peacock, A.C. Nonlinear optical properties of polycrystalline silicon core fibers from telecom wavelengths into the mid-infrared spectral region. Opt. Mater. Express 2019, 9, 1271–1279. [Google Scholar] [CrossRef]
  34. Shiryaev, V.S.; Churbanov, M.F. Recent advances in preparation of high-purity chalcogenide glasses for mid-IR photonics. J. Non-Cryst. Solids 2017, 475, 1–9. [Google Scholar] [CrossRef]
  35. Tao, G.M.; Ebendorff-Heidepriem, H.; Stolyarov, A.M.; Danto, S.; Badding, J.V.; Fink, Y.; Ballato, J.; Abouraddy, A.F. Infrared fibers. Adv. Opt. Photonics 2015, 7, 379–458. [Google Scholar] [CrossRef]
  36. Churbanov, M.F.; Snopatin, G.E.; Shiryaev, V.S.; Plotnichenko, V.G.; Dianov, E.M. Recent advances in preparation of high-purity glasses based on arsenic chalcogenides for fiber optics. J. Non-Cryst. Solids 2011, 357, 2352–2357. [Google Scholar] [CrossRef]
  37. Aktas, O. Chalcogenide microresonators tailored to distinct morphologies by the shaping of glasses on silica tapers. Opt. Lett. 2017, 42, 907–910. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Q.M.; Li, M.; Hao, Q.; Deng, D.H.; Zhou, H.; Zeng, H.P.; Zhan, L.; Wu, X.; Liu, L.Y.; Xu, L. Fabrication and characterization of on-chip optical nonlinear chalcogenide nanofiber devices. Opt. Lett. 2010, 35, 3829–3831. [Google Scholar] [CrossRef] [PubMed]
  39. Ahmad, R.; Rochette, M. Photosensitivity at 1550 nm and Bragg grating inscription in As2Se3 chalcogenide microwires. Appl. Phys. Lett. 2011, 99, 061109. [Google Scholar] [CrossRef]
  40. Ahmad, R.; Rochette, M.; Baker, C. Fabrication of Bragg gratings in subwavelength diameter As2Se3 chalcogenide wires. Opt. Lett. 2011, 36, 2886–2888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Florea, C.; Sanghera, J.S.; Shaw, B.; Aggarwal, I.D. Fiber Bragg gratings in As2S3 fibers obtained using a 0/−1 phase mask. Opt. Mater. 2009, 31, 942–944. [Google Scholar] [CrossRef]
  42. Gao, S.; Baker, C.; Cai, W.; Chen, L.; Bao, X.Y. 10 kHz-34 MHz ultrasound detection based on a dual-core hybrid taper. APL Photonics 2019, 4, 110805. [Google Scholar] [CrossRef] [Green Version]
  43. Fan, H.B.; Chen, L.; Bao, X.Y. Chalcogenide microfiber-assisted silica microfiber for ultrasound detection. Opt. Lett. 2020, 45, 1128–1131. [Google Scholar] [CrossRef] [PubMed]
  44. Yeom, D.I.; Magi, E.C.; Lamont, M.R.; Roelens, M.A.; Fu, L.B.; Eggleton, B.J. Low-threshold supercontinuum generation in highly nonlinear chalcogenide nanowires. Opt. Lett. 2008, 33, 660–662. [Google Scholar] [CrossRef]
  45. Xie, Y.; Cai, D.W.; Wu, H.; Pan, J.; Zhou, N.; Xin, C.G.; Yu, S.L.; Wang, P.; Jiang, X.S.; Qiu, J.R.; et al. Mid-infrared chalcogenide microfiber knot resonators. Photonics Res. 2020, 8, 616–621. [Google Scholar] [CrossRef]
  46. Cai, D.W.; Xie, Y.; Wang, P.; Zhang, L.; Guo, X.; Tong, L.M. Mid-infrared microfiber Bragg gratings. Opt. Lett. 2020, 45, 6114–6117. [Google Scholar] [CrossRef] [PubMed]
  47. Tong, L.M.; Lou, J.Y.; Ye, Z.Z.; Svacha, G.T.; Mazur, E. Self-modulated taper drawing of silica nanowires. Nanotechnology 2005, 16, 1445–1448. [Google Scholar] [CrossRef]
  48. Tong, L.M.; Hu, L.L.; Zhang, J.J.; Qiu, J.R.; Yang, Q.; Lou, J.Y.; Shen, Y.H.; He, J.L.; Ye, Z.Z. Photonic nanowires directly drawn from bulk glasses. Opt. Express 2006, 14, 82–87. [Google Scholar] [CrossRef]
  49. Li, L.Z.; Al-Kadry, A.; Abdukerim, N.; Rochette, M. Design, fabrication and characterization of PC, COP and PMMA-cladded As2Se3 microwires. Opt. Mater. Express 2016, 6, 912–921. [Google Scholar] [CrossRef]
  50. Baker, C.; Rochette, M. A generalized heat-brush approach for precise control of the waist profile in fiber tapers. Opt. Mater. Express 2011, 1, 1065–1076. [Google Scholar] [CrossRef]
  51. Rodney, W.S.; Malitson, I.H.; King, T.A. Refractive index of arsenic trisulfide. J. Opt. Soc. Am. 1958, 48, 633–636. [Google Scholar] [CrossRef]
  52. Van Popta, A.C.; DeCorby, R.G.; Haugen, C.J.; Robinson, T.; McMullin, J.N.; Tonchev, D.; Kasap, S.O. Photoinduced refractive index change in As2Se3 by 633nm illumination. Opt. Express 2002, 10, 639–644. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, K.J.; Yang, S.Y.; Tong, L.M. Modeling of evanescent coupling between two parallel optical nanowires. Appl. Opt. 2007, 46, 1429–1434. [Google Scholar] [CrossRef]
  54. Singh, V.; Lin, P.T.; Patel, N.; Lin, H.; Li, L.; Zou, Y.; Deng, F.; Ni, C.; Hu, J.; Giammarco, J.; et al. Mid-infrared materials and devices on a Si platform for optical sensing. Sci. Technol. Adv. Mater. 2014, 15, 014603. [Google Scholar] [CrossRef]
  55. Hudson, D.D.; Dekker, S.A.; Maegi, E.C.; Judge, A.C.; Jackson, S.D.; Li, E.B.; Sanghera, J.S.; Shaw, L.B.; Aggarwal, I.D.; Eggleton, B.J. Octave spanning supercontinuum in an As2S3 taper using ultralow pump pulse energy. Opt. Lett. 2011, 36, 1122–1124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Guo, X.; Qiu, M.; Bao, J.M.; Wiley, B.J.; Yang, Q.; Zhang, X.N.; Ma, Y.G.; Yu, H.K.; Tong, L.M. Direct coupling of plasmonic and photonic nanowires for hybrid nanophotonic components and circuits. Nano Lett. 2009, 9, 4515–4519. [Google Scholar] [CrossRef] [PubMed]
  57. Ding, Z.X.; Huang, Z.N.; Chen, Y.; Mou, C.B.; Lu, Y.Q.; Xu, F. All-fiber ultrafast laser generating gigahertz-rate pulses based on a hybrid plasmonic microfiber resonator. Adv. Photon. 2020, 2, 026002. [Google Scholar] [CrossRef] [Green Version]
  58. Ma, Y.Q.; Farrell, G.; Semenova, Y.; Li, B.H.; Yuan, J.H.; Sang, X.Z.; Yan, B.B.; Yu, C.X.; Guo, T.; Wu, Q. Optical microfiber-loaded surface plasmonic TE-pass polarizer. Opt. Laser Technol. 2016, 78, 101–105. [Google Scholar] [CrossRef]
  59. Vahala, K.J. Optical microcavities. Nature 2003, 424, 839–846. [Google Scholar] [CrossRef] [PubMed]
  60. He, L.N.; Ozdemir, K.; Zhu, J.G.; Kim, W.; Yang, L. Detecting single viruses and nanoparticles using whispering gallery microlasers. Nat. Nanotechnol. 2011, 6, 428–432. [Google Scholar] [CrossRef] [Green Version]
  61. Choi, J.; Noh, H. Enhanced absorption by coherent control in a photonic crystal resonator coupled with a microfiber. Opt. Lett. 2018, 43, 5532–5534. [Google Scholar] [CrossRef] [PubMed]
  62. Kim, M.K.; Hwang, I.K.; Kim, S.H.; Chang, H.J.; Lee, Y.H. All-optical bistable switching in curved microfiber-coupled photonic crystal resonators. Appl. Phys. Lett. 2007, 90, 161118. [Google Scholar] [CrossRef] [Green Version]
  63. Hwang, I.K.; Kim, S.K.; Yang, J.K.; Kim, S.H.; Lee, S.H.; Lee, Y.H. Curved-microfiber photon coupling for photonic crystal light emitter. Appl. Phys. Lett. 2005, 87, 131107. [Google Scholar] [CrossRef] [Green Version]
  64. Lecaplain, C.; Javerzac-Galy, C.; Gorodetsky, M.L.; Kippenberg, T.J. Mid-infrared ultra-high-Q resonators based on fluoride crystalline materials. Nat. Commun. 2016, 7, 13383. [Google Scholar] [CrossRef] [PubMed]
  65. Xie, Y.; Cai, D.W.; Pan, J.; Zhou, N.; Gao, Y.X.; Jin, Y.Y.; Jiang, X.S.; Qiu, J.R.; Wang, P.; Guo, X.; et al. Batch fabrication of high-quality infrared chalcogenide microsphere resonators. Small 2021, 17, 2100140. [Google Scholar] [CrossRef] [PubMed]
  66. Heylman, K.D.; Knapper, K.A.; Horak, E.H.; Rea, M.T.; Vanga, S.K.; Goldsmith, R.H. Optical microresonators for sensing and transduction: A materials perspective. Adv. Mater. 2017, 29, 1700037. [Google Scholar] [CrossRef]
  67. Sumetsky, M.; Dulashko, Y.; Fini, J.M.; Hale, A.; DiGiovanni, D.J. The microfiber loop resonator: Theory, experiment, and application. J. Lightwave Technol. 2006, 24, 242–250. [Google Scholar] [CrossRef]
  68. Guo, X.; Li, Y.H.; Jiang, X.S.; Tong, L.M. Demonstration of critical coupling in microfiber loops wrapped around a copper rod. Appl. Phys. Lett. 2007, 91, 073512. [Google Scholar] [CrossRef]
  69. Jiang, X.S.; Tong, L.M.; Vienne, G.; Guo, X.; Tsao, A.; Yang, Q.; Yang, D.R. Demonstration of optical microfiber knot resonators. Appl. Phys. Lett. 2006, 88, 223501. [Google Scholar] [CrossRef]
  70. Wang, P.; Zhang, L.; Yang, Z.Y.; Gu, F.X.; Wang, S.S.; Yang, Q.; Tong, L.M. Fusion spliced microfiber closed-loop resonators. IEEE Photonics Technol. Lett. 2010, 22, 1075–1077. [Google Scholar] [CrossRef]
  71. Li, W.; Wang, P.; Hu, Z.F.; Tong, L.M. Fusion splicing soft glass microfibers for photonic devices. IEEE Photonics Technol. Lett. 2011, 23, 831–833. [Google Scholar] [CrossRef]
  72. Zhang, B.; Zeng, P.Y.; Yang, Z.L.; Xia, D.; Zhao, J.X.; Sun, Y.D.; Huang, Y.F.; Song, J.C.; Pan, J.S.; Cheng, H.J.; et al. On-chip chalcogenide microresonators with low-threshold parametric oscillation. Photonics Res. 2021, 9, 1272–1279. [Google Scholar] [CrossRef]
  73. Sumetsky, M. Optical fiber microcoil resonators. Opt. Express 2004, 12, 2303–2316. [Google Scholar] [CrossRef]
  74. Scheuer, J.; Sumetsky, M. Optical-fiber microcoil waveguides and resonators and their applications for interferometry and sensing. Laser Photon. Rev. 2011, 5, 465–478. [Google Scholar] [CrossRef]
  75. Xu, Z.L.; Sun, Q.Z.; Li, B.R.; Luo, Y.Y.; Lu, W.G.; Liu, D.M.; Shum, P.P.; Zhang, L. Highly sensitive refractive index sensor based on cascaded microfiber knots with Vernier effect. Opt. Express 2015, 23, 6662–6672. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, M.; Tang, R.; Luo, A.P.; Xu, W.C.; Luo, Z.C. Graphene-decorated microfiber knot as a broadband resonator for ultrahigh-repetition-rate pulse fiber lasers. Photonics Res. 2018, 6, C1–C7. [Google Scholar] [CrossRef]
  77. Luan, F.; Magi, E.; Gong, T.X.; Kabakova, I.; Eggleton, B.J. Photoinduced whispering gallery mode microcavity resonator in a chalcogenide microfiber. Opt. Lett. 2011, 36, 4761–4763. [Google Scholar] [CrossRef] [PubMed]
  78. Ma, P.; Choi, D.Y.; Yu, Y.; Yang, Z.Y.; Vu, K.; Nguyen, T.; Mitchell, A.; Luther-Davies, B.; Madden, S. High Q factor chalcogenide ring resonators for cavity-enhanced MIR spectroscopic sensing. Opt. Express 2015, 23, 19969–19979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Grobnic, D.; Mihailov, S.J.; Ding, H.M.; Smelser, C.W. Bragg grating evanescent field sensor made in biconical tapered fiber with femtosecond IR radiation. IEEE Photonics Technol. Lett. 2006, 18, 160–162. [Google Scholar] [CrossRef]
  80. Fang, X.; Liao, C.R.; Wang, D.N. Femtosecond laser fabricated fiber Bragg grating in microfiber for refractive index sensing. Opt. Lett. 2010, 35, 1007–1009. [Google Scholar] [CrossRef] [Green Version]
  81. Ran, Y.; Jin, L.; Tan, Y.N.; Sun, L.P.; Li, J.; Guan, B.O. High-efficiency ultraviolet inscription of Bragg gratings in microfibers. IEEE Photonics J. 2012, 4, 181–186. [Google Scholar] [CrossRef]
  82. Kou, J.L.; Qiu, S.J.; Xu, F.; Lu, Y.Q. Demonstration of a compact temperature sensor based on first-order Bragg grating in a tapered fiber probe. Opt. Express 2011, 19, 18452–18457. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, Y.X.; Meng, C.; Zhang, A.P.; Xiao, Y.; Yu, H.K.; Tong, L.M. Compact microfiber Bragg gratings with high-index contrast. Opt. Lett. 2011, 36, 3115–3117. [Google Scholar] [CrossRef] [PubMed]
  84. Nayak, K.P.; Le Kien, F.; Kawai, Y.; Hakuta, K.; Nakajima, K.; Miyazaki, H.T.; Sugimoto, Y. Cavity formation on an optical nanofiber using focused ion beam milling technique. Opt. Express 2011, 19, 14040–14050. [Google Scholar] [CrossRef] [PubMed]
  85. Guan, B.O.; Li, J.; Jin, L.; Ran, Y. Fiber Bragg gratings in optical microfibers. Opt. Fiber Technol. 2013, 19, 793–801. [Google Scholar] [CrossRef]
  86. Gao, S.; Jin, L.; Ran, Y.; Sun, L.P.; Li, J.; Guan, B.O. Temperature compensated microfiber Bragg gratings. Opt. Express 2012, 20, 18281–18286. [Google Scholar] [CrossRef] [PubMed]
  87. Luo, W.; Kou, J.L.; Chen, Y.; Xu, F.; Lu, Y.Q. Ultra-highly sensitive surface-corrugated microfiber Bragg grating force sensor. Appl. Phys. Lett. 2012, 101, 133502. [Google Scholar] [CrossRef]
  88. Bernier, M.; El-Amraoui, M.; Couillard, J.F.; Messaddeq, Y.; Vallee, R. Writing of Bragg gratings through the polymer jacket of low-loss As2S3 fibers using femtosecond pulses at 800 nm. Opt. Lett. 2012, 37, 3900–3902. [Google Scholar] [CrossRef] [PubMed]
  89. Baker, C.; Gao, S.; Chen, L.; Bao, X.Y. Self-inscribed antisymmetric long-period grating in a dual-core As2Se3-PMMA fiber. Opt. Express 2017, 25, 12409–12414. [Google Scholar] [CrossRef]
  90. Wang, L.L.; Ma, W.Q.; Zhang, P.Q.; Yang, D.D.; Zhu, L.; Wang, X.S.; Dai, S.X. Design and analysis of long-period fiber gratings in tapered multimode chalcogenide glass fiber for temperature measurement. J. Opt. Soc. Am. B 2019, 36, 1792–1798. [Google Scholar] [CrossRef]
  91. Supradeepa, V.R.; Feng, Y.; Nicholson, J.W. Raman fiber lasers. J. Opt. 2017, 19, 023001. [Google Scholar] [CrossRef]
  92. Pierce, M.C.; Jackson, S.D.; Dickinson, M.R.; King, T.A.; Sloan, P. Laser-tissue interaction with a continuous wave 3-μm fibre laser: Preliminary studies with soft tissue. Lasers Surg. Med. 2000, 26, 491–495. [Google Scholar] [CrossRef]
  93. Jackson, S.D. Towards high-power mid-infrared emission from a fibre laser. Nat. Photonics 2012, 6, 423–431. [Google Scholar] [CrossRef]
  94. White, R.T.; Monro, T.M. Cascaded Raman shifting of high-peak-power nanosecond pulses in As2S3 and As2Se3 optical fibers. Opt. Lett. 2011, 36, 2351–2353. [Google Scholar] [CrossRef] [PubMed]
  95. Abdukerim, N.; Li, L.Z.; El Amraoui, M.; Messaddeq, Y.; Rochette, M. 2 μm Raman fiber laser based on a multimaterial chalcogenide microwire. Appl. Phys. Lett. 2017, 110, 161103. [Google Scholar] [CrossRef]
  96. Guo, X.H.; Li, Z.R.; Jiao, Y.D.; Zhao, Z.P.; Yao, C.F.; Jia, Z.X.; Dong, K.G.; Zhang, H.Y.; Lin, H.H.; Qin, W.P.; et al. A shower of mid-infrared Raman solitons at designed wavelength of similar to 3 μm from a tapered fluorotellurite fiber. Laser Phys. 2021, 31, 095103. [Google Scholar] [CrossRef]
  97. Alamgir, I.; Abdukerim, N.; Rochette, M. In situ fabrication of far-detuned optical fiber wavelength converters. Opt. Lett. 2019, 44, 4467–4470. [Google Scholar] [CrossRef] [PubMed]
  98. Godin, T.; Combes, Y.; Ahmad, R.; Rochette, M.; Sylvestre, T.; Dudley, J.M. Far-detuned mid-infrared frequency conversion via normal dispersion modulation instability in chalcogenide microwires. Opt. Lett. 2014, 39, 1885–1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Li, L.Z.; Abdukerim, N.; Rochette, M. Mid-infrared wavelength conversion from As2Se3 microwires. Opt. Lett. 2017, 42, 639–642. [Google Scholar] [CrossRef]
  100. Liang, F.; Kang, L.; Lin, Z.S.; Wu, Y.C. Mid-infrared nonlinear optical materials based on metal chalcogenides: Structure-property relationship. Cryst. Growth Des. 2017, 17, 2254–2289. [Google Scholar] [CrossRef]
  101. Abdukerim, N.; Li, L.Z.; Rochette, M. Chalcogenide-based optical parametric oscillator at 2 μm. Opt. Lett. 2016, 41, 4364–4367. [Google Scholar] [CrossRef]
  102. Al-Kadry, A.; El Amraoui, M.; Messaddeq, Y.; Rochette, M. Two octaves mid-infrared supercontinuum generation in As2Se3 microwires. Opt. Express 2014, 22, 31131–31137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Wang, Y.Y.; Dai, S.X.; Li, G.T.; Xu, D.; You, C.Y.; Han, X.; Zhang, P.Q.; Wang, X.S.; Xu, P.P. 1.4–7.2 μm broadband supercontinuum generation in an As-S chalcogenide tapered fiber pumped in the normal dispersion regime. Opt. Lett. 2017, 42, 3458–3461. [Google Scholar] [CrossRef] [PubMed]
  104. Li, G.T.; Peng, X.F.; Dai, S.X.; Wang, Y.Y.; Xie, M.; Yang, L.; Yang, C.F.; Wei, W.Y.; Zhang, P.Q. Highly coherent 1.5–8.3 μm broadband supercontinuum generation in tapered As-S chalcogenide fibers. J. Lightwave Technol. 2019, 37, 1847–1852. [Google Scholar] [CrossRef]
  105. Hudson, D.D.; Antipov, S.; Li, L.Z.; Alamgir, I.; Hu, T.; Amraoui, M.; Messaddeq, Y.; Rochette, M.; Jackson, S.D.; Fuerbach, A. Toward all-fiber supercontinuum spanning the mid-infrared. Optica 2017, 4, 1163–1166. [Google Scholar] [CrossRef]
  106. Marandi, A.; Rudy, C.W.; Plotnichenko, V.G.; Dianov, E.M.; Vodopyanov, K.L.; Byer, R.L. Mid-infrared supercontinuum generation in tapered chalcogenide fiber for producing octave-spanning frequency comb around 3 μm. Opt. Express 2012, 20, 24218–24225. [Google Scholar] [CrossRef] [PubMed]
  107. Dudley, J.M.; Genty, G.; Coen, S. Supercontinuum generation in photonic crystal fiber. Rev. Mod. Phys. 2006, 78, 1135–1184. [Google Scholar] [CrossRef]
  108. Al-kadry, A.; Baker, C.; El Amraoui, M.; Messaddeq, Y.; Rochette, M. Broadband supercontinuum generation in As2Se3 chalcogenide wires by avoiding the two-photon absorption effects. Opt. Lett. 2013, 38, 1185–1187. [Google Scholar] [CrossRef]
  109. Rudy, C.W.; Marandi, A.; Vodopyanov, K.L.; Byer, R.L. Octave-spanning supercontinuum generation in in situ tapered As2S3 fiber pumped by a thulium-doped fiber laser. Opt. Lett. 2013, 38, 2865–2868. [Google Scholar] [CrossRef] [Green Version]
  110. Sun, Y.A.; Dai, S.X.; Zhang, P.Q.; Wang, X.S.; Xu, Y.S.; Liu, Z.J.; Chen, F.F.; Wu, Y.A.; Zhang, Y.J.; Wang, R.P.; et al. Fabrication and characterization of multimaterial chalcogenide glass fiber tapers with high numerical apertures. Opt. Express 2015, 23, 23472–23483. [Google Scholar] [CrossRef]
  111. Anashkina, E.A.; Shiryaev, V.S.; Koptev, M.Y.; Stepanov, B.S.; Muravyev, S.V. Development of of As-Se tapered suspended-core fibers for ultra-broadband mid-IR wavelength conversion. J. Non-Cryst. Solids 2018, 480, 43–50. [Google Scholar] [CrossRef] [Green Version]
  112. Leonov, S.O.; Wang, Y.C.; Shiryaev, V.S.; Snopatin, G.E.; Stepanov, B.S.; Plotnichenko, V.G.; Vicentini, E.; Gambetta, A.; Coluccelli, N.; Svelto, C.; et al. Coherent mid-infrared supercontinuum generation in tapered suspended-core As39Se61 fibers pumped by a few-optical-cycle Cr:ZnSe laser. Opt. Lett. 2020, 45, 1346–1349. [Google Scholar] [CrossRef]
  113. Leindecker, N.; Marandi, A.; Byer, R.L.; Vodopyanov, K.L. Broadband degenerate OPO for mid-infrared frequency comb generation. Opt. Express 2011, 19, 6304–6310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Reid, D.T.; Gale, B.J.S.; Sun, J. Frequency comb generation and carrier-envelope phase control in femtosecond optical parametric oscillators. Laser Phys. 2008, 18, 87–103. [Google Scholar] [CrossRef]
  115. Keilmann, F.; Gohle, C.; Holzwarth, R. Time-domain mid-infrared frequency-comb spectrometer. Opt. Lett. 2004, 29, 1542–1544. [Google Scholar] [CrossRef] [Green Version]
  116. Erny, C.; Moutzouris, K.; Biegert, J.; Kuhlke, D.; Adler, F.; Leitenstorfer, A.; Keller, U. Mid-infrared difference-frequency generation of ultrashort pulses tunable between 3.2 and 4.8 μm from a compact fiber source. Opt. Lett. 2007, 32, 1138–1140. [Google Scholar] [CrossRef] [Green Version]
  117. Del’Haye, P.; Herr, T.; Gavartin, E.; Gorodetsky, M.L.; Holzwarth, R.; Kippenberg, T.J. Octave spanning tunable frequency comb from a microresonator. Phys. Rev. Lett. 2011, 107, 063901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Del’Haye, P.; Schliesser, A.; Arcizet, O.; Wilken, T.; Holzwarth, R.; Kippenberg, T.J. Optical frequency comb generation from a monolithic microresonator. Nature 2007, 450, 1214–1217. [Google Scholar] [CrossRef] [Green Version]
  119. Bernhardt, B.; Sorokin, E.; Jacquet, P.; Thon, R.; Becker, T.; Sorokina, I.T.; Picque, N.; Hansch, T.W. Mid-infrared dual-comb spectroscopy with 2.4 μm Cr2+:ZnSe femtosecond lasers. Appl. Phys. B 2010, 100, 3–8. [Google Scholar] [CrossRef] [Green Version]
  120. Adler, F.; Cossel, K.C.; Thorpe, M.J.; Hartl, I.; Fermann, M.E.; Ye, J. Phase-stabilized, 1.5 W frequency comb at 2.8–4.8 μm. Opt. Lett. 2009, 34, 1330–1332. [Google Scholar] [CrossRef] [PubMed]
  121. Lee, K.F.; Granzow, N.; Schmidt, M.A.; Chang, W.; Wang, L.; Coulombier, Q.; Troles, J.; Leindecker, N.; Vodopyanov, K.L.; Schunemann, P.G.; et al. Midinfrared frequency combs from coherent supercontinuum in chalcogenide and optical parametric oscillation. Opt. Lett. 2014, 39, 2056–2059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Sanghera, J.S.; Shaw, L.B.; Busse, L.E.; Talley, D.; Aggarwal, I.D. Infrared transmitting fiber optics for biomedical applications. In Proceedings of the BiOS ’99 International Biomedical Optics Symposium, San Jose, CA, USA, 23–29 January 1999; Volume 3596, pp. 178–187. [Google Scholar]
  123. Sanghera, J.S.; Kung, F.H.; Busse, L.E.; Pureza, P.C.; Aggarwal, I.D. Infrared evanescent absorption-spectroscopy of toxic-chemicals using chalcogenide class fibers. J. Am. Ceram. Soc. 1995, 78, 2198–2202. [Google Scholar] [CrossRef]
  124. Michel, K.; Bureau, B.; Pouvreau, C.; Sangleboeuf, J.C.; Boussard-Pledel, C.; Jouan, T.; Rouxel, T.; Adam, J.L.; Staubmann, K.; Steinner, H.; et al. Development of a chalcogenide glass fiber device for in situ pollutant detection. J. Non-Cryst. Solids 2003, 326, 434–438. [Google Scholar] [CrossRef]
  125. Huang, P.; Huang, T.Y.; Zeng, S.W.; Pan, J.X.; Wu, X.; Zhao, X.; Wu, Y.H.; Ping, P.S.; Brambilla, G. Nonlinear gas sensing based on third-harmonic generation in cascaded chalcogenide microfibers. J. Opt. Soc. Am. B 2019, 36, 300–305. [Google Scholar] [CrossRef]
  126. Wang, J.Q.; Chen, Y.Z.; Geng, Y.F.; Hong, X.M.; Li, X.J. Theoretical design of mid-infrared graphene optical gas sensor based on slot Si core fiber. IEEE Photonics Technol. Lett. 2019, 31, 1096–1099. [Google Scholar] [CrossRef]
  127. Wang, M.; Yang, F.; Dai, S.X.; Cao, Z.F.; Su, J.X.; Ding, S.J.; Zhang, P.Q. Effect of the geometries of Ge-Sb-Se chalcogenide glass tapered fiber on the sensitivity of evanescent wave sensors. J. Lightwave Technol. 2021, 39, 4828–4836. [Google Scholar] [CrossRef]
  128. Michel, K.; Bureau, B.; Boussard-Pledel, C.; Jouan, T.; Adam, J.L.; Staubmann, K.; Baumann, T. Monitoring of pollutant in waste water by infrared spectroscopy using chalcogenide glass optical fibers. Sens. Actuators B Chem. 2004, 101, 252–259. [Google Scholar] [CrossRef]
  129. Keirsse, J.; Boussard-Pledel, C.; Loreal, O.; Sire, O.; Bureau, B.; Turlin, B.; Leroyer, P.; Lucas, J. Chalcogenide glass fibers used as biosensors. J. Non-Cryst. Solids 2003, 326, 430–433. [Google Scholar] [CrossRef]
  130. Le Coq, D.; Michel, K.; Keirsse, J.; Boussard-Pledel, C.; Fonteneaua, G.; Bureau, B.; Le Quere, J.M.; Sire, O.; Lucas, J. Infrared glass fibers for in-situ sensing, chemical and biochemical reactions. Comptes Rendus Chim. 2002, 5, 907–913. [Google Scholar] [CrossRef]
  131. Hocde, S.; Loreal, O.; Sire, O.; Boussard-Pledel, C.; Bureau, B.; Turlin, B.; Keirsse, J.; Leroyer, P.; Lucas, J. Metabolic imaging of tissues by infrared fiber-optic spectroscopy: An efficient tool for medical diagnosis. J. Biomed. Opt. 2004, 9, 404–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Gutierrez-Arroyo, A.; Baudet, E.; Bodiou, L.; Nazabal, V.; Rinnert, E.; Michel, K.; Bureau, B.; Colas, F.; Charrier, J. Theoretical study of an evanescent optical integrated sensor for multipurpose detection of gases and liquids in the mid-infrared. Sens. Actuators B Chem. 2017, 242, 842–848. [Google Scholar] [CrossRef]
  133. Wang, X.M.; Su, J.X.; Wang, Y.Y.; Yang, C.F.; Dai, S.X.; Zhang, P.Q. High-sensitivity sensing in bare Ge-Sb-Se chalcogenide tapered fiber with optimal structure parameters. J. Non-Cryst. Solids 2021, 559, 120686. [Google Scholar] [CrossRef]
Figure 1. Fabrication of ChG MFs. (a) Schematic diagram of taper drawing a ChG MF from a conventional ChG fiber. Insets: SEM images of an as-fabricated As2S3 MF. (b) Schematic diagram of taper drawing a hybrid ChG-PMMA MF. Inset: optical image of as-fabricated hybrid ChG biconical fiber taper. Reprinted with the permission from [49] © The Optical Society. (c1) Schematic diagram of taper drawing a long-waist ChG MF by a two-step taper-drawing process. (c2) 62-micrometer diameter As2S3 MF-based knot assembled from a 3.5-micrometer diameter MF fabricated by the method in (c1). Reprinted with the permission from [45] © CLP Publishing.
Figure 1. Fabrication of ChG MFs. (a) Schematic diagram of taper drawing a ChG MF from a conventional ChG fiber. Insets: SEM images of an as-fabricated As2S3 MF. (b) Schematic diagram of taper drawing a hybrid ChG-PMMA MF. Inset: optical image of as-fabricated hybrid ChG biconical fiber taper. Reprinted with the permission from [49] © The Optical Society. (c1) Schematic diagram of taper drawing a long-waist ChG MF by a two-step taper-drawing process. (c2) 62-micrometer diameter As2S3 MF-based knot assembled from a 3.5-micrometer diameter MF fabricated by the method in (c1). Reprinted with the permission from [45] © CLP Publishing.
Photonics 08 00497 g001
Figure 2. Numerical solutions of diameter-dependent (a) effective refractive index (neff), (b) fractional evanescent fields (η), and (c) effective mode area (Aeff) of As2S3 and As2Se3 MFs at wavelengths of 1.55 μm, 4 μm, and 8 μm, respectively.
Figure 2. Numerical solutions of diameter-dependent (a) effective refractive index (neff), (b) fractional evanescent fields (η), and (c) effective mode area (Aeff) of As2S3 and As2Se3 MFs at wavelengths of 1.55 μm, 4 μm, and 8 μm, respectively.
Photonics 08 00497 g002
Figure 3. Mid-IR MF near-field couplers. (a) Schematic diagram of As2S3 tapered fiber coupler used for CdTe microwire transmission spectra measurement. Reprinted with the permission from [32] © The Optical Society. (b1) Optical micrograph and SEM image of the MgF2 microresonator coupled with an As2S3 biconical fiber taper. Inset: Finite element model simulation of the optical intensity profile of the fundamental WGM at λ = 4.5 μm. (b2) Transmission as a function of the coupling parameter K = κ e x / κ 0 for varying taper waist radius when taper-resonator coupling was achieved. Reprinted with the permission from [64] © Springer Nature. (c) Schematic illustration of As2S3 biconical fiber taper used for characterization of an As2S3 microsphere. Reprinted with the permission from [65] © John Wiley and Sons.
Figure 3. Mid-IR MF near-field couplers. (a) Schematic diagram of As2S3 tapered fiber coupler used for CdTe microwire transmission spectra measurement. Reprinted with the permission from [32] © The Optical Society. (b1) Optical micrograph and SEM image of the MgF2 microresonator coupled with an As2S3 biconical fiber taper. Inset: Finite element model simulation of the optical intensity profile of the fundamental WGM at λ = 4.5 μm. (b2) Transmission as a function of the coupling parameter K = κ e x / κ 0 for varying taper waist radius when taper-resonator coupling was achieved. Reprinted with the permission from [64] © Springer Nature. (c) Schematic illustration of As2S3 biconical fiber taper used for characterization of an As2S3 microsphere. Reprinted with the permission from [65] © John Wiley and Sons.
Photonics 08 00497 g003
Figure 4. ChG MF-based resonators. (a1) Optical micrograph of the intertwisted overlap region of a SU8 polymer-embedded As2S3 MKR. (a2) Transmission spectrum of the MKR. Reprinted with the permission from [38] © The Optical Society. (b1) SEM image of a 178.2-μm diameter As2Se3 MLR. The inset shows the self-coupling region. (b2) Transmission spectrum of the MLR. Inset: Lorentzian fitting to a resonance mode at 1552.9 nm. Reprinted with the permission from [37] © The Optical Society. (c1) Optical micrograph of an 824-μm diameter As2S3 MKR. Inset: optical micrograph of the intertwisted overlap area. (c2) Transmission spectrum of the MKR. (c3) Lorentzian fitting to a resonance mode centered at 4469.14-nm wavelength. (c4) Transmission spectra of an As2S3 MKR by tightening the knot structure in liquid (from top to bottom). (c5) Resonance shift of an MKR with the temperature rising from 31.4 °C to 59.8 °C. Inset: transmission spectra of resonance modes corresponding to 31.4 °C (blue line) and 40.6 °C (red line). Reprinted with the permission from [45] © CLP Publishing.
Figure 4. ChG MF-based resonators. (a1) Optical micrograph of the intertwisted overlap region of a SU8 polymer-embedded As2S3 MKR. (a2) Transmission spectrum of the MKR. Reprinted with the permission from [38] © The Optical Society. (b1) SEM image of a 178.2-μm diameter As2Se3 MLR. The inset shows the self-coupling region. (b2) Transmission spectrum of the MLR. Inset: Lorentzian fitting to a resonance mode at 1552.9 nm. Reprinted with the permission from [37] © The Optical Society. (c1) Optical micrograph of an 824-μm diameter As2S3 MKR. Inset: optical micrograph of the intertwisted overlap area. (c2) Transmission spectrum of the MKR. (c3) Lorentzian fitting to a resonance mode centered at 4469.14-nm wavelength. (c4) Transmission spectra of an As2S3 MKR by tightening the knot structure in liquid (from top to bottom). (c5) Resonance shift of an MKR with the temperature rising from 31.4 °C to 59.8 °C. Inset: transmission spectra of resonance modes corresponding to 31.4 °C (blue line) and 40.6 °C (red line). Reprinted with the permission from [45] © CLP Publishing.
Photonics 08 00497 g004
Figure 5. Mid-IR MF gratings. (a) Experimental setup for the As2Se3 MF Bragg grating photo inscription assisted by prism and in-situ monitoring of the process. Reprinted with the permission from [40] © The Optical Society. (b1) Schematic illustration of the As2S3 mFBG fabrication setup using interference pattern formed by two 532 nm beams. Inset: optical micrograph of a 2.4-μm diameter ChG mFBG. (b2) Schematic illustration of the optical characterization of ChG mid-IR mFBGs. (b3) Transmission spectra of mFBGs with different MF diameters. Reprinted with the permission from [46] © The Optical Society.
Figure 5. Mid-IR MF gratings. (a) Experimental setup for the As2Se3 MF Bragg grating photo inscription assisted by prism and in-situ monitoring of the process. Reprinted with the permission from [40] © The Optical Society. (b1) Schematic illustration of the As2S3 mFBG fabrication setup using interference pattern formed by two 532 nm beams. Inset: optical micrograph of a 2.4-μm diameter ChG mFBG. (b2) Schematic illustration of the optical characterization of ChG mid-IR mFBGs. (b3) Transmission spectra of mFBGs with different MF diameters. Reprinted with the permission from [46] © The Optical Society.
Photonics 08 00497 g005
Figure 6. Raman scattering, four-wave mixing and modulation instability in mid-IR MFs. (a) Output spectra of the Raman laser based on a multi material ChG MF with increasing pump power. Inset: Average output power versus average pump power. Reprinted with the permission from [95] © AIP Publishing. (b) Conversion efficiency vs. idler wavelength for different tapering periods of all-fiber wavelength converters based on tapering As2Se3 MFs. Reprinted with the permission from [97] © The Optical Society. (c) Experimental observation of MI when pumping an As2Se3-polymer MF at λ = 2620 nm in the normal dispersion regime. Input peak powers estimated at (A) 200 mW, (B) 1 W and (C) 10 W. Reprinted with the permission from [98] © The Optical Society. (d) Measured output idler spectra resulting from wavelength tuning of the probe from 1.587 to 1.650 μm with a power of 0.54 mW. Reprinted with the permission from [99] © The Optical Society.
Figure 6. Raman scattering, four-wave mixing and modulation instability in mid-IR MFs. (a) Output spectra of the Raman laser based on a multi material ChG MF with increasing pump power. Inset: Average output power versus average pump power. Reprinted with the permission from [95] © AIP Publishing. (b) Conversion efficiency vs. idler wavelength for different tapering periods of all-fiber wavelength converters based on tapering As2Se3 MFs. Reprinted with the permission from [97] © The Optical Society. (c) Experimental observation of MI when pumping an As2Se3-polymer MF at λ = 2620 nm in the normal dispersion regime. Input peak powers estimated at (A) 200 mW, (B) 1 W and (C) 10 W. Reprinted with the permission from [98] © The Optical Society. (d) Measured output idler spectra resulting from wavelength tuning of the probe from 1.587 to 1.650 μm with a power of 0.54 mW. Reprinted with the permission from [99] © The Optical Society.
Photonics 08 00497 g006
Figure 7. Supercontinuum generation and frequency comb generation in mid-IR MF. (a) SC spectra from amplified pump pulses using a 1.6-micrometer diameter As2Se3 MF for various amplified pump power levels. Inset: the autocorrelation trace of the amplified pump pulse. Reprinted with the permission from [102] © The Optical Society. (b) Measured SC spectra generated from 12-centimeter long As-S tapered MFs with diameters varying from 4.3 and 7.9 μm pumped at 3.25 μm. Reprinted with the permission from [103] © The Optical Society. (c) Simulated SC spectra coherence properties for different waist core diameters in As-S MFs. Reprinted with the permission from [104] © IEEE. (d) Spectral expansion at increasing peak power in a polymer-protected As2Se3/As2S3 MF pumping at 3 μm. The inset shows the mode profile in the MF section (3-μm core diameter) at various wavelengths. Reprinted with the permission from [105] © The Optical Society. (e1) Spectrum of SC generation around 3 μm from the As2S3 tapered MF compared with the OPO output and the simulation result (e2) RF spectrum of the interference of a c.w. laser at 1564 nm with the second harmonic (SH) of the tapered MF output. Reprinted with the permission from [106] © The Optical Society.
Figure 7. Supercontinuum generation and frequency comb generation in mid-IR MF. (a) SC spectra from amplified pump pulses using a 1.6-micrometer diameter As2Se3 MF for various amplified pump power levels. Inset: the autocorrelation trace of the amplified pump pulse. Reprinted with the permission from [102] © The Optical Society. (b) Measured SC spectra generated from 12-centimeter long As-S tapered MFs with diameters varying from 4.3 and 7.9 μm pumped at 3.25 μm. Reprinted with the permission from [103] © The Optical Society. (c) Simulated SC spectra coherence properties for different waist core diameters in As-S MFs. Reprinted with the permission from [104] © IEEE. (d) Spectral expansion at increasing peak power in a polymer-protected As2Se3/As2S3 MF pumping at 3 μm. The inset shows the mode profile in the MF section (3-μm core diameter) at various wavelengths. Reprinted with the permission from [105] © The Optical Society. (e1) Spectrum of SC generation around 3 μm from the As2S3 tapered MF compared with the OPO output and the simulation result (e2) RF spectrum of the interference of a c.w. laser at 1564 nm with the second harmonic (SH) of the tapered MF output. Reprinted with the permission from [106] © The Optical Society.
Photonics 08 00497 g007
Figure 8. Mid-IR MF-based sensors. (a1) Schematic illustration of gas sensing based on cascaded ChG MFs. (a2) Detection limit of methane concentration for cascaded MFs sensing at different L1. Hereon, L2 = 9.5 cm and P0 = 0.1 W. Reprinted with the permission from [125] © The Optical Society. (b1) Schematic illustration of graphene on-Si slot fiber structure. (b2) Reflection spectra of Bragg gratings based on graphene on-Si slot fiber with different NO2 gas concentrations. Reprinted with the permission from [126] © IEEE. (c1) Experimental setup of Ge15Sb15Se70 tapered MF sensor. (c2) Dependence of the absorbance at 9.78 μm on the methanol aqueous solution concentration for the asymmetric and symmetric structure tapered fiber sensors. Reprinted with the permission from [127] © IEEE.
Figure 8. Mid-IR MF-based sensors. (a1) Schematic illustration of gas sensing based on cascaded ChG MFs. (a2) Detection limit of methane concentration for cascaded MFs sensing at different L1. Hereon, L2 = 9.5 cm and P0 = 0.1 W. Reprinted with the permission from [125] © The Optical Society. (b1) Schematic illustration of graphene on-Si slot fiber structure. (b2) Reflection spectra of Bragg gratings based on graphene on-Si slot fiber with different NO2 gas concentrations. Reprinted with the permission from [126] © IEEE. (c1) Experimental setup of Ge15Sb15Se70 tapered MF sensor. (c2) Dependence of the absorbance at 9.78 μm on the methanol aqueous solution concentration for the asymmetric and symmetric structure tapered fiber sensors. Reprinted with the permission from [127] © IEEE.
Photonics 08 00497 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cai, D.; Xie, Y.; Guo, X.; Wang, P.; Tong, L. Chalcogenide Glass Microfibers for Mid-Infrared Optics. Photonics 2021, 8, 497. https://doi.org/10.3390/photonics8110497

AMA Style

Cai D, Xie Y, Guo X, Wang P, Tong L. Chalcogenide Glass Microfibers for Mid-Infrared Optics. Photonics. 2021; 8(11):497. https://doi.org/10.3390/photonics8110497

Chicago/Turabian Style

Cai, Dawei, Yu Xie, Xin Guo, Pan Wang, and Limin Tong. 2021. "Chalcogenide Glass Microfibers for Mid-Infrared Optics" Photonics 8, no. 11: 497. https://doi.org/10.3390/photonics8110497

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