Open Access
1 March 2024 Optical microfiber or nanofiber: a miniature fiber-optic platform for nanophotonics
Jianbin Zhang, Hubiao Fang, Pan Wang, Wei Fang, Lei Zhang, Xin Guo, Limin Tong
Author Affiliations +
Abstract

An optical micro/nanofiber (MNF) is a quasi-one-dimensional free-standing optical waveguide with a diameter close to or less than the vacuum wavelength of light. Combining the tiny geometry with high-refractive-index contrast between the core and the surrounding, the MNF exhibits favorable optical properties such as tight optical confinement, strong evanescent field, and large-diameter-dependent waveguide dispersion. Meanwhile, as a quasi-one-dimensional structure with extraordinarily high geometric and structural uniformity, the MNF also has low optical loss and high mechanical strength, making it favorable for manipulating light on the micro/nanoscale with high flexibility. Over the past two decades, optical MNFs, typically being operated in single mode, have been emerging as a miniaturized fiber-optic platform for both scientific research and technological applications. In this paper, we aim to provide a comprehensive overview of the representative advances in optical MNFs in recent years. Starting from the basic structures and fabrication techniques of the optical MNFs, we highlight linear and nonlinear optical and mechanical properties of the MNFs. Then, we introduce typical applications of optical MNFs from near-field optics, passive optical components, optical sensors, and optomechanics to fiber lasers and atom optics. Finally, we give a brief summary of the current status of MNF optics and technology, and provide an outlook into future challenges and opportunities.

1.

Introduction

An optical micro/nanofiber (MNF) is a kind of quasi one-dimensional (1D) free-standing optical waveguide with a diameter close to or less than the vacuum wavelength of the transmitted light. The earliest report of glass MNFs with diameters around 1 µm could be traced back to 1887 when Boys fabricated glass threads by drawing molten minerals at a high speed[1]. However, at that time, those MNFs were not used for optical waveguiding. Instead, owing to their excellent elasticity and small resilience of these threads, they were used as elastic springs or suspension wire for measuring a very small force or torsion[2,3]. In 1910, Hondros and Debye reported the first theoretical model for waveguiding light along a dielectric cylinder, and showed that electromagnetic waves could be confined and propagated in a lossless dielectric cylindrical waveguide, with a diameter below the wavelength[4]. However, it was not until the 1950s that these waveguiding modes in cylindrical waveguides, named as “surface waves,” began to receive much attention[58]. In 1951, O’Brien and van Heel proposed covering low-index cladding on the surface of cylindrical waveguides to reduce the crosstalk between the waveguides[9], and in the following years, Hopkins and Kapany demonstrated sub-micrometer multiple fibers coated with low-index cladding for image transmission[1012]. In 1966, Kao and Hockham proposed the possibility of developing low-loss glass fibers[13], opening up the era of optical fiber communication, as well as low-loss fiber optics and technology. From then on, taper drawing a standard silica fiber became a routine approach to fabricating an MNF (typically called the fiber taper) connected with a silica fiber through a conical taper, and the shape of the taper and the diameter of the waist (i.e., MNF) could be controlled much better than before[1418]. Relying on the surface waves (i.e., waveguided evanescent fields) of these MNFs[19,20], a variety of applications were proposed or developed including optical filters[21,22], couplers[23,24], evanescent field amplification[25], sensors[26,27], and supercontinuum generation[28], while the diameters of the MNFs used or assumed were mostly larger than the vacuum wavelength of the guided light.

In 2003, Tong and Mazur experimentally demonstrated that subwavelength-diameter silica nanofibers taper drawn from silica fibers could be used for low-loss optical waveguiding[29], opening an opportunity for guiding light in MNFs with smaller sizes and stronger “surface waves,” which in turn, bestowed the MNFs with tight optical confinement, strong evanescent fields, and highly engineerable waveguide dispersion[30]. Shortly after, such MNFs were experimentally proved to have losses (e.g., 1.4 dB/mm[31]) lower than all other subwavelength-width optical waveguides in visible and near-infrared (NIR) spectral ranges[32]. Since then, a number of improved MNF fabrication techniques, based on flame-heated[3340], electrically heated[4145], and CO2 laser-heated[4648] taper drawings of standard optical fibers, have been demonstrated[49,50]. So far, waveguiding loss of silica MNFs can typically be below 0.1 dB/m, with the lowest value of 0.03 dB/m[37] (i.e., three orders of magnitude lower than that of a planar waveguide with similar width[51]). Also, with a high-precision real-time diameter-monitoring technique[38,39,45,50,52], the diameter of the MNF can be obtained on-demand, with the best precision of ±2  nm[39]. Moreover, by keeping the tapering angle below the mode-transition critical angle, the single-mode light field in a standard-sized fiber can be squeezed into an MNF mode almost losslessly (typical loss <0.0015  dB), with the highest fiber-MNF-fiber overall transmittance exceeding 99.9% (i.e., insertion loss <0.005  dB)[53]. The excellent fiber compatibility not only offers an efficient and compact input/output scheme, but also facilitates the handling and manipulation of MNFs in experiments. More recently, based on the ultra-low optical loss of pristine MNFs, high-power (>10  W) continuous-wave (CW) optical waveguiding in a subwavelength-diameter silica MNF at 1550-nm wavelength has been experimentally realized, with the smallest MNF diameter down to 410 nm (i.e., λ/3.8)[54], opening an opportunity for MNF-based high-power optical applications. In addition, besides the silica glass, many other types of glass (e.g., phosphate, tellurite, and chalcogenide glass[5557]) have also been drawn into MNFs, which greatly enriches the category of optical MNFs.

Generally, as the fiber diameter decreases to the subwavelength scale, the MNF exhibits fascinating optical properties that are different from standard optical fibers, including a strong evanescent field, tight optical confinement, surface field enhancement, and diameter/wavelength-dependent large waveguide dispersion[30,5860]. Also, benefitting from nearly perfect structural and geometric uniformities, glass MNFs possess remarkable mechanical properties, such as high tensile strength (e.g., higher than 10 GPa[61]) and excellent elasticity[6163], which enable robust and flexible manipulation of freestanding MNFs in various surroundings (e.g., vacuum, gas, or liquid). These favorable optical and mechanical properties, as well as high physical and chemical stabilities of silica glass, make the MNFs a versatile platform for studying light-matter interaction on the micro/nanoscale and developing related photonic technologies [see Fig. 1]. Firstly, the strong evanescent field of the MNF is ideal for studying the near-field interaction between the waveguide mode and matter [e.g., molecules, nanoparticles (NPs), and two-dimensional (2D) materials] located close to the fiber surface, and developing highly efficient near-field coupling techniques (e.g., in/out-coupling of nanowaveguides), as well as functionalization of the MNF itself (e.g., evanescently coupled active or nonlinear materials). Secondly, the tightly confined high-fraction evanescent fields of optical MNFs make the waveguiding field highly sensitive to the refractive-index change Δn of the surrounding environment and/or coupled samples, offering special advantages for high-sensitivity optical sensing on the micro/nanoscale[6466]. Thirdly, when being squeezed adiabatically from a larger-area mode of a standard optical fiber, the tightly confined field of an MNF mode (i.e., with a very small mode area), is highly favorable for enhancing nonlinear optical effects in the MNF or coupled materials[6769], as has been widely investigated for harmonic generation[7072], Brillouin scattering (BS)[73,74], four-wave mixing (FWM)[75], stimulated Raman scattering (SRS)[76,77], and supercontinuum generation[31,78,79]. Fourthly, the waveguide dispersion of optical MNFs is strongly dependent on the diameter or wavelength, offering a compact, flexible, low-loss, and fiber-compatible scheme for dispersion management in nonlinear optics[78,80], pulse compression[81], and fiber lasers[8284]. Fifthly, by selecting an appropriate diameter-to-wavelength ratio (D/λ), a large field gradient in the evanescent field can be generated in the vicinity of the MNF surface, which has been exploited for manipulating micro/nanoparticles[85] and cold atoms[8688]. Finally, owing to their small mass, optical MNFs can exhibit a sensitive optomechanical response (typically manifested as mechanical vibrations) to the momentum change of the waveguided light fields[8991], which has been adopted for studying optoacoustic interactions[73,74,92,93] and optomechanical technology[94]. So far, optical MNFs have been attracting broad interest from near-field optics, nonlinear optics, atom optics, and optomechanics to optical nanowaveguides, micro-couplers, resonators, sensors, and lasers.

Fig. 1

Overall description of optical MNFs in terms of characteristics and applications.

PI_3_1_R02_f001.png

Overall, the past two decades have witnessed an encouraging development of optical MNFs. As a miniaturized fiber-optic platform, optical MNFs not only retain the intrinsically outstanding waveguiding properties of conventional optical fibers, but also offer favorable nanophotonic behaviors. Parallel to the flourishing of fiber-optic technology and nanotechnology in the early 21st century, high-quality glass MNFs have been developed as low-dimensional ultralow-loss optical waveguides on the (sub)wavelength scale, and have been inspiring abundant research interest in multidisciplinary fields from time to time. Previously, MNF optics and applications have been reviewed in many articles[49,50,5860,6567,69,8488,95,96], with most of them focusing on a certain specific research area such as fabrication[49,50], nonlinear optics[67,69], optomechanics[85], atom optics[8688], passive components[60], optical sensors[6466], and fiber lasers[84]. There are several comprehensive review articles[58,59] or a monograph[97], but they all have been published for more than 10 years. Now it is the time to give an updated and comprehensive review, to include the latest progress and new insights in this field.

In this regard, we review the development of optical MNFs over the past two decades, with emphases on their fabrication, properties, and applications. The article contains the following six sections: (1) Introduction—a brief introduction to the historical background and a summary of properties and potentials of glass MNFs; (2) Fabrication of Optical MNFs—recent advances in the fabrication of optical MNFs, especially an advanced fabrication technique with real-time high-precision diameter control; (3) Optical Waveguiding Properties of the MNFs—theoretical and experimental waveguiding properties of optical MNFs, including linear and nonlinear optical properties; (4) Mechanical Behavior—elastic and plastic deformation, with emphases on elastic strain and tensile strength of glass MNFs; (5) MNF-Based Applications—from near-field optical coupling, passive optical components, optical sensors, and optomechanics to fiber lasers and atom optics; (6) Conclusion and Outlook—a brief summary of the current status of MNF optics and technology, and an outlook of future challenges and opportunities.

2.

Fabrication of Optical MNFs

The materials of optical MNFs are generally divided into glassy or crystal materials. Silica glass, which presents a number of advantages of excellent homogeneity, broadband optical transparency (around 250–2800 nm), high physical and chemical stability, low thermal expansion, temperature-dependent viscosity, and easy remolding, is a typical glassy material for optical MNFs. Historically, since the first glass MNF was drawn from molten quartz by a flying arrow in the late 19th century[1], with the advancement of technology in the first half of the 20th century, silica MNFs were feasible to fabricate from glass in a stable and efficient manner, and their optical properties (e.g., birefringence and bending loss) were investigated[98,99]. After the 1960s, along with the development of standard glass fibers, glass MNFs were routinely taper drawn from glass fibers heated by flame or laser beams[24,100,101], which also facilitated optical launching and handling of the MNFs that were naturally connected to the glass fibers. Meanwhile, mechanized or automated stretching systems were invented to better control the tapering shape and improve the repeatability[18,24]. The emergence of subwavelength-diameter MNFs in the 2000s, especially for waveguide dispersion management and cold atom manipulation, puts forward higher requirements on the diameter accuracy, and mechanized stretching systems with real-time feedback for high-precision diameter control were thus developed.

For glass MNF, high-temperature drawing is the main fabrication technique. When heated, glass becomes viscous[102], making it possible to be drawn into a fiber with a diameter down to nanometer scale. Meanwhile, the high-temperature drawing process bestows the MNF a pristine molten-frozen surface, with an intrinsic surface roughness down to sub-nanometer scale[103105]. Therefore, compared with other approaches (e.g., chemical etching[106] and excimer laser ablation[107]), the high-temperature drawing technique offers glass MNFs with unparalleled geometric uniformity and surface smoothness (e.g., 0.2  nm in surface roughness), which are essential for low-loss optical waveguiding[29,33,34,56]. With the assistance of a scanning heating system (i.e., “flame brush”[18,49,108] and later the “scanning fiber” with a motionless flame[40]), the length of the uniform-diameter MNFs can reach tens of centimeters[40].

Inspired by glass MNFs, other types of optical MNFs have been fabricated from functional materials other than glass. For example, a variety of polymers [e.g., polymethyl methacrylate (PMMA)[109], polystyrene (PS)[110,111], and polyacrylamide (PAM)[112]] or biomaterials (e.g., silkworm silk[113], spider silk[114,115], lotus root silk[116], and Escherichia coli[117]) have been fabricated into MNFs by physical (solution/melt) drawing[109112], chemical/biological synthesis[113118], or electrospinning[119], and used to waveguide light for a variety of purposes. In addition, recently, crystal micro/nanoscale wires made of semiconductors[120123], dielectrics[124,125], or even ice[126] have also been called optical “nanofibers” or “microfibers” when they are used for optical waveguiding, just as we named waveguiding wire-like sapphire single crystals as sapphire optical fibers some 30 years ago[125,127], although they are typically fabricated by bottom-up crystal growth processes[128132].

This section focuses on high-temperature drawing techniques of glass optical MNFs and the corresponding diameter control and characterization techniques. The fabrication of other types of MNFs can be found in many review articles elsewhere[128136].

2.1.

Taper Drawing of Silica Fibers

The taper-drawing technique is a top-down approach that physically tapers and extends a structure based on its viscosity at a certain temperature or concentration. This technique works for glassy materials (e.g., glass, polymer, or even metals) and can reduce the cross-sectional size of the taper down to nanometer scale, while maintaining an ultra-low surface roughness. Figure 2(a) shows a structural diagram of taper drawing a standard glass fiber by stretching both sides of the heated fiber. When the diameter goes down to the target size, an MNF is obtained with both ends connected to standard fibers through conical transition regions, and this kind of MNF is usually called a “biconical” MNF. The typical taper drawing process and the geometry of the transition region of an MNF can be predicted by a tapering model proposed by Birks and Li[18]. Relying on a flame brush technique, which will be introduced in detail in Section 2.1.1, one can fabricate MNFs with various profiles (e.g., linear, exponential, and sinusoidal) of the transition region for different purposes.

Fig. 2

(a) Structural diagram of a biconical optical MNF. SMF, single-mode fiber; Ω, tapering angle of the transition region; DT, fiber diameter in the transition region. (b) Tapering angle of a typical MNF (black squares) and calculated critical angle (red dots) as functions of local taper diameter at 1550-nm wavelength[40].

PI_3_1_R02_f002.png

The optical loss of the transition region is mostly concerned. As a bridge connecting with a standard fiber pigtail and the MNF, the transition region allows the compression of the propagation light field from 10-μm to sub-1-μm sizes. When the propagation light goes through the transition region, the fundamental mode is progressively waveguided at the silica-air interface as the higher-order modes, leading to continuous weak coupling and interference between the fundamental mode and high-order modes. If the taper angle of the transition region is too steep [Ω in Fig. 2(a)], a certain fraction of the mode energy will be transferred from the fundamental mode to high-order modes and radiation modes, resulting in a radiation loss and thus reduction of the optical transmittance[137]. To prevent the mode leakage as far as possible, the waveguiding mode should adiabatically couple from the core to the cladding and couple back to the core, suggesting that all the energy remains in the fundamental mode during the mode evolution[16,18,137,138]. The adiabatic criterion for the transition region is[18,138,139]

Eq. (1)

DT/2tanΩ2πβ1β2,
where the left-hand formula represents the characteristic length of the fiber diameter variation with a tapering angle Ω [see Fig. 2(a)], while the right-hand formula represents the beat length between the fundamental mode and one of the high-order modes. DT is the fiber diameter of the transition region. β1 and β2 are propagation constants of the fundamental mode and the first excited high-order mode (i.e., HE12 mode), respectively. Figure 2(b) shows the profile of the transition region of a typical adiabatic MNF. The red dots are the calculated critical angle of the HE12 mode, while the black dots, always located below the red dots, represent the actual tapering angles Ω. Under such an adiabatic criterion, the mode-coupling loss can be minimized during the MNF fabrication.

For a standard silica fiber, the softening temperature and annealing temperature are about 1665°C[140] and 1140°C[141], respectively. To draw a silica fiber into an MNF, the heating temperature should be higher than the annealing temperature. Since the viscosity of the fiber is temperature-dependent, the optimal drawing speed for a low-loss MNF varies with the heating temperature. So far, the reported heating temperature for drawing a silica fiber falls between 1160°C and about 1500°C, with a drawing speed ranging from 0.04 mm/s to about 0.5 mm/s, correspondingly[3746,49,50]. As a reference, for a drawing speed of 0.1 mm/s, the typical heating temperature of the heating source is ranging from 1200°C to 1300°C.

In real fabrication, the precision of mechanical motion components and the stability of the thermal field for fiber heating are crucial to obtain MNFs with excellent diameter uniformities and extraordinary surface smoothness. Also, the choice of heating methods is particularly important for taper drawing different glass materials, as well as for obtaining MNFs with different properties. Currently, there are three heating methods: flame, electric, and laser-heated methods, as introduced below.

2.1.1.

Flame-heated taper drawing of glass fibers

The typical fabrication system of the “flame-heated taper drawing” is shown in Fig. 3(a). A fixed gas nozzle, cooperating with a mass flow controller, provides a stable flame (e.g., oxyhydrogen flame[37,142,143] and isobutane-oxygen flame[33]) for heating a standard fiber. Once a bare fiber located at the heating zone reaches a stable temperature, two high-precision translation stages on both sides smoothly draw the fiber into an MNF. Usually, in the case of a fixed hot zone and simply drawing a fiber to both sides, it is challenging to achieve a uniform waist length exceeding 1 cm. As an alternative, the flame brush technique, first used for fabricating fiber couplers[144], has been commonly adopted to fabricate fiber tapers[18] and optical MNFs[58] with larger lengths. Using this fabrication technique and the fabrication system in Fig. 3(a), we can obtain a long-length optical MNF. Figure 3(b) gives the schematic (upper, not to scale) and experimentally measured (bottom) fiber diameter evolution of a typical biconical silica MNF (930±2  nm in diameter and 9 cm in uniform length), in which the scanning electron microscope (SEM) image [see Fig. 3(b), bottom inset] reveals an extremely smooth surface of the MNF.

Fig. 3

Typical flame-heated taper-drawing fabrication system. (a) Photograph of a flame-heated taper-drawing system for fabricating silica MNFs. Inset: close-up image of the flame nozzle. (b) Schematic (upper, not to scale) and experimentally measured (bottom) fiber diameter evolution of a biconical drawn silica MNF along the fiber length[54]. The MNF has a diameter of about 930 nm and a uniform length of 9 cm. The diameter evolution of the tapering region (red circles) was measured by an optical microscope (upper insets), while that of the MNF (blue circles) was measured by a scanning electron microscope (bottom inset).

PI_3_1_R02_f003.png

Typically, the flame brush technique has two configurations. The first is that the flame moves to and fro to heat and scan an optical fiber, while two translation stages move outward to draw the fiber[145,146]. Note that the hydrogen flame may be fluttered by the airflow during the reciprocating motion, resulting in an uneven heating of the fiber. Alternatively, using an electric heater or a CO2 laser heater can be much more stable in the scanning process, which is also called as “modified flame brush technique”[58]. In the second type the flame is motionless, while the fiber is scanned and stretched on top of the flame, driven by the translation stages[40]. It is worth mentioning that the excellent stability and smoothness of the moving translation stages are critical to removing uncertainties and obtaining a high-quality MNF in the taper-drawing process. In recent years, with the development and widespread adoption of high-precision, high-stability, and highly controllable translation stages, the uncertainties in the MNF fabrication, especially in the second configuration, have been effectively suppressed.

With these improved fabrication systems, MNFs with low loss, large length, and high diameter uniformity can be routinely obtained. For instance, in 2014 Hoffman et al. achieved an optical transmittance as high as 99.95±0.02% in a single-mode MNF (500 nm in diameter and 5 mm in length) at 780-nm wavelength[37], corresponding to a loss as low as 2.6×105  dB/mm. In 2020, Yao et al. reported a 99.4%-transmittance low-loss MNF with a diameter of 1.2 µm, diameter uniformity of 107, and a length of 10 cm at 1550-nm wavelength[40].

Typical optical and electron microscope images of as-fabricated MNFs are shown in Fig. 4. Figure 4(a) shows a bright-field optical microscopic image of a 550-nm-diameter silica MNF. Although measuring the accurate diameter of such a thin MNF is beyond the capability of conventional optical microscopy, the diameter uniformity and the defect-free surface of the MNF can be clearly observed. To examine the MNF with a higher resolution, electron microscopes are typically adopted. Figures 4(b)4(e) present SEM images of as-fabricated silica MNFs, manifesting excellent diameter uniformities with fiber diameter down to 30 nm [see Fig. 4(b), a bundle of MNFs with diameters of 140, 510, and 30 nm, respectively], extraordinary surface smoothness [see Fig. 4(c)], long length [see Fig. 4(d)], and outstanding mechanical strength [see Fig. 4(e)]. To investigate the surface roughness of optical MNFs, a higher-magnification transmission electron microscope (TEM) can be used. Figure 4(f) gives a TEM image of a 330-nm-diameter MNF[29], obtaining a root-mean-square roughness of about 0.2 nm, which is much better than all other types of subwavelength-diameter/width optical waveguides. The electron diffraction pattern [see the inset of Fig. 4(f)] manifests that the silica MNF is amorphous.

Fig. 4

Structural characterization of silica MNFs. (a) Optical microscope image of a 550-nm-diameter silica MNF. SEM images of (b) self-supporting bundle of MNFs assembled with silica MNFs with diameters of 140, 510, and 30 nm[97], (c) 790-nm-diameter silica MNF with a smooth surface, (d) coiled 260-nm-diameter silica MNF with a total length of about 4 mm[29], and (e) 360-nm-diameter silica MNF with a bending radius of 3 µm[95]. (f) TEM image of the surface of a 330-nm-diameter silica MNF[29]. Inset: electron diffraction pattern of the MNF.

PI_3_1_R02_f004.png

However, the flame-heated configuration has some limitations: (1) the flame will cause airflow disturbance to the MNF, especially to the MNF with a diameter of a few hundred nanometers or less; (2) the combustion byproducts may adhere to the surface of MNFs, resulting in the surface contamination. To address these issues, one may use other types of heating sources.

2.1.2.

Electrically heated taper drawing of glass fibers

Compared with the flame, the electric heater not only provides a more stable temperature environment with minimal airflow disturbance and contamination, but also offers an opportunity to draw fiber in the atmosphere other than air (e.g., argon and nitrogen), and can thus effectively isolate the glass from oxygen or OH when needed. Also, by controlling the supply current, the heating temperature can be controlled much more precisely and conveniently than the flame. Figure 5 shows photographs of typical electric heaters used for drawing glass MNFs. For high-temperature operation (e.g., >1100°C), the optional materials for heating elements are silicon nitride, silicon carbide, platinum, and metal alloy wires (e.g., FeCrAl). Besides silica fibers, the electric heating method is also suitable for taper drawing MNFs from other types of glass fibers, especially soft glass with relatively low softening temperatures [see Fig. 5(b)][55,147,148]. More recently, using a wide-zone electric heater, a parallel fabrication technique of silica MNFs has been reported[149], which enables simultaneous drawing of multiple MNFs with almost identical geometries. However, it should be mentioned that compared to the flame, the electric heating element will inevitably suffer from aging when operating at high temperatures (e.g., >1100°C) in the long term, and thus usually has a limited lifespan (e.g., hundreds of hours).

Fig. 5

Typical electric heaters in fiber-drawing systems. Photographs of (a) a ceramic heater for drawing silica MNFs (NTT-AT, CMH-7022) and (b) a self-designed U-type copper heater for drawing soft-glass MNFs.

PI_3_1_R02_f005.png

2.1.3.

Laser-heated taper drawing of glass fibers

In the far-infrared spectrum, silica glass exhibits significant optical absorption. Therefore, a 10.6-μm-wavelength CO2 laser can be used for pollution-free heating and drawing fibers[150,151]. Previously, a direct laser-heating approach was adopted for fabricating silica MNFs with relatively large diameters (e.g., several micrometers). For example, in 1999 Dimmick et al. used a CO2 laser beam [13 W in power and 820 µm at half-maximum (FWHM) in a focused spot size] to fabricate MNFs with diameters down to 4.6 µm[24]. However, due to the Mie scattering, the effective absorption decreases with a reduced fiber diameter. Consequently, when the fiber diameter goes down to a certain value at which the effective heating temperature of the MNF drops below the allowable drawing temperature, the drawing process is forced to stop[24,150]. Typically, with a reasonable laser power (e.g., less than 1 kW), the minimum diameter of the MNF obtained by this direct laser-heating approach can hardly go below 1 µm. To draw thinner MNFs with a laser-heating system, an indirect laser-heating approach can be used. In 2004, using a sapphire tube to absorb the CO2 laser beam and create a stable high-temperature zone for heating a fiber, Sumetsky et al. successfully fabricated MNFs with diameters less than 100 nm[152]. Such an indirect laser-heating configuration offers the flexibility to adjust the heating temperature by the laser power and temperature distribution by the tube geometry. More recently, using micro-sized plasmonic heaters (i.e., pieces of metal plates), Jia et al. obtained nonadiabatic optical MNFs with steep tapering angles and ultra-short transition regions (e.g., a few tens of micrometers)[153]. Owing to the compactness of the fabrication system, the taper-drawing process can be carried out inside an SEM chamber and in-situ monitored with a high resolution.

2.1.4.

High-precision diameter control and measurement

High-precision diameter control is essential to avoid radiation loss in the tapering transition region, as well as to field distribution and waveguide dispersion management in the uniform MNF for applications ranging from near-field optical coupling[154], optical sensing[65,66], and harmonic generation[71] to pulse compression[81,84] and atom trapping[8688]. To date, the most frequently adopted method is real-time feedback control by monitoring the intermodal interference and high-order-mode cutoff during the fabrication process[38,39,45]. In 2014, Yu et al. first reported a diameter-control method with feedback by monitoring the cutoff of the LP02 mode in the intermodal interference[45], showing a diameter error of less than 2%. Later in 2017, Xu et al. demonstrated a feedback control of the MNF diameter by monitoring the drops of high-order modes at 785-nm wavelength, stopping the drawing process after a certain time according to the constant hot-zone model, and obtaining a precision better than 5 nm for fiber diameters in the range of 800–1300 nm[38]. More recently, Kang et al. improved the diameter-control technique by employing a white-light source as the probing light, real-time monitoring, and immediately stopping the drawing process at the cutoff of the TE01 mode[39], and realized a precision of about ±2  nm with fiber diameters from 360 to 680 nm. Figure 6 shows the real-time diameter control of the taper drawing a 360-nm-diameter MNF by monitoring the TE01-mode cutoff at 532-nm wavelength. Once the abrupt intensity drop for the TE01 mode is observed, the MNF pulling process is terminated instantly.

Fig. 6

Schematic diagram of the diameter-control technique in the fabrication of a silica MNF based on the mode-cutoff feedback[39].

PI_3_1_R02_f006.png

With the above-mentioned high-precision diameter control technique, MNFs with accurate diameters and excellent diameter uniformity can be fabricated with good reproducibility, which have found applications in MNF Fabry-Pérot (FP) cavities[53] and high-efficiency harmonic generation[54]. Besides the high-order-mode cutoff, the scattering intensity of waveguiding evanescent fields has also been used for real-time diameter control[52]. Also, the diameter uniformity and control accuracy can be further improved by optimizing the heating condition, feedback system, and mechanical performance of the fabrication system.

In addition, it is worth mentioning that, driven by the demand for in-situ diameter measurement of the MNF, a number of non-destructive optical methods have been reported over the last 20 years. For example, in 2004 Warken and Giessen demonstrated that the profile of subwavelength-diameter MNFs could be measured by the diffraction pattern, with a resolution of ±50  nm over a length of 5 cm[155]. Similarly, a series of optical microscopy techniques have also been reported for measuring the MNF diameter with a resolution of below 50 nm[156159]. In 2006, Sumetsky et al. proposed a fiber-assisted scanning detection technique for probing the surface and bulk distortions of optical MNFs[160]. By scanning a partly stripped 125-μm-diameter fiber along an MNF and detecting the optical transmission of the MNF, sub-nanometer measurement accuracy of the MNF diameter was realized. Besides, many other techniques based on external gratings[161,162], near-field probing[52,163165], stress-strain analysis[62], Rayleigh scattered light imaging[166,167], optical backscattering reflectometry technique[168], short-time Fourier transform analysis on the modal evolution[169], second- and third-harmonic generation[170], and forward BS[171], with a resolution from 15 nm to 40 pm, have been reported in recent years.

2.2.

More Fabrication Techniques

The aforementioned technologies primarily use standard fibers as preforms for fabricating MNFs with mechanized systems. However, not all materials can be obtained in fiber forms, especially those actively functionalized glass or polymer materials. Therefore, several other fabrication techniques, including drawing MNFs directly from bulk pieces of glass[56], polymer solution[112,172174], or melt[175177], have been reported in recent years. Meanwhile, the fabrication of glass MNFs using chemical etching[106] and electric arc[178180], and polymer MNFs using chemical synthesis[118], electrospinning[119], and nanolithography[181], has also been demonstrated. In addition, bottom-up synthesized techniques, originally for growing crystalline whiskers[182], have also been improved for growing highly uniform single-crystal photonic semiconductor micro/nanowires (also called MNFs recently)[136,183,184], oxide nanowires[185187], or even ice microfibers[126], for low-loss optical waveguiding. As most of these techniques have been reviewed elsewhere[128136], limited by the space of this article, we will not go into detail.

3.

Optical Waveguiding Properties of the MNFs

For guiding light on the micro/nanoscale, optical waveguiding behaviors are the most concerned properties of the MNFs. Since the first waveguiding model of the dielectric cylinders[4], so far, comprehensive theoretical models have been established for both linear and nonlinear optical waveguiding in MNFs. Incorporated with versatile numerical calculation software products (e.g., MATLAB, Mathematica, COMSOL, and Ansys Lumerical FDTD), waveguiding behaviors of complicated MNF-based structures (e.g., bending loss in bent MNFs[188,189], near-field coupling between multiple MNFs[190194], output endface patterns[195], and strong mode-coupling-induced ultra-confined optical fields[196198]) can now be obtained with a high precision. Meanwhile, along with the fast-developed experimental techniques for the manipulation and characterization of individual MNFs, the optical properties of MNFs have been measured and engineered experimentally. The first half of this section aims to provide a comprehensive understanding of the waveguiding behaviors in subwavelength-diameter MNFs based on analytical and numerical approaches. In the latter half of the section, we mainly discuss and summarize recent advances in optical waveguiding properties.

3.1.

Fundamental Waveguide Theory of Optical MNFs

Due to the large index contrast between the core of subwavelength-diameter MNFs (e.g., n=1.44 for silica) and surrounding (e.g., n=1.0 for air), it is difficult to calculate the propagation light field using weakly guiding approximation. The most frequently used method is solving for the propagation constants based on exact solutions of Maxwell’s equations and numerical calculations[30]. Generally, an as-fabricated optical MNF possesses an extremely uniform diameter, smooth surface [see Figs. 4(a), 4(c), and 4(f)], and perfect circular cross section [see Fig. 7(a)], which are also well-defined in the theoretical model. The length of the MNFs (typically >10  μm) is large enough to establish the spatial steady state, and the MNF diameter (D, typically >10  nm) is not very thin so that the macroscale parameters of permittivity (ε) and permeability (μ) can be used to describe the responses of a dielectric medium to an incident electromagnetic field. The MNF is assumed to have an infinite air-clad with a step-index profile (i.e., two-layer structure), which is expressed as

Eq. (2)

n(r)={n1,0<r<D/2n2,D/2r<},
where n1 and n2 are the refractive indices of the MNF core and the surrounding, respectively. It is worth noting that for MNFs with more layers of index profiles, a multiple-layer-structured cylindrical waveguide model should be employed[199201]. For non-dissipative and source-free MNF materials, Maxwell’s equations can be reduced to the following Helmholtz equations[30,202]:

Eq. (3)

(2+n2k2β2)e=0,(2+n2k2β2)h=0,
where k=2π/λ, λ is the vacuum wavelength of the transmitted light and β is the propagation constant. Given the perfect circular cross section of a subwavelength-diameter MNF, Eq. (3) can be analytically solved in the cylindrical coordinate and the eigenvalue equations for the HEvm and EHvm modes are obtained as[30,203]

Eq. (4)

{Jv(U)UJv(U)+Kv(W)WKv(W)}{Jv(U)UJv(U)+n22Kv(W)n12WKv(W)}=(vβkn1)2(VUW)4;
for the TE0m modes:

Eq. (5)

J1(U)UJ0(U)+K1(W)WK0(W)=0;
and for the TM0m modes:

Eq. (6)

n12J1(U)UJ0(U)+n22K1(W)WK0(W)=0,
with the normalized transverse wave numbers U and W given by

Eq. (7)

U=D2(k2n12β2)12,W=D2(β2k2n22)12,
with the V-number given by

Eq. (8)

V=D2k(n12n22)12,
where Jv is the Bessel function of the first kind and Kv is the modified Bessel function of the second kind.

Fig. 7

Calculation of waveguiding modes in optical MNFs. (a) SEM image of a 400-nm-diameter tellurite glass MNF with a circular cross section[56]. (b) Calculated propagation constant (β) of waveguiding modes in an air-clad silica MNF at a wavelength of 633 nm[30]. Solid line: fundamental mode. Dotted lines: high-order modes. Dashed line: critical diameter for single-mode operation. (c) Electric fields of several waveguiding modes in a 600-nm-diameter silica MNF at 633-nm wavelength. (d) SEM image of a 900-nm-diameter CdS nanowire with a hexagonal cross section[208]. d1 and d2 are the diagonal-circle approximation diameter and circular-area-equivalence diameter for nanowires with different cross sections, respectively.

PI_3_1_R02_f007.png

By numerically solving the eigenvalues of Eqs. (4(5))–(6), the propagation constants of waveguiding modes supported by a circular-cross-section MNF can be obtained [see Fig. 7(b)]. The diameter-dependent waveguiding properties of air-clad silica MNFs at 633-nm wavelength are presented. One can clearly see that when the MNF diameter reduces to a certain value [denoted as a dashed vertical line in Fig. 7(b), corresponding to the V-number to be 2.405], only the fundamental mode (i.e., HE11 mode) exists and higher-order modes are effectively suppressed. In other words, the V-number can be used to evaluate whether an MNF is operated at a single mode for a given wavelength. In principle, for the case of single-mode operation in a subwavelength-diameter MNF, the electric field of the HE11 mode is quasi-linearly polarized. Specifically, the transverse component of the field is linearly polarized in time at each fixed local point, while the total electric field vector rotates elliptically with time, in a plane parallel to the MNF axis[88,204]. The transverse component inside the MNF is not only linearly polarized in time but also almost linearly polarized in space, while the orientation angle of the transverse component outside the MNF varies in space. To maintain the polarization state in the waveguiding mode field of a subwavelength-diameter MNF, several methods have been developed including using scattering imaging[205,206] and directional coupling[207]. Additionally, it is widely known that standard fiber falls in the regime of paraxial or weakly guiding, and thus the description of waveguiding modes can be greatly simplified using degenerated linearly polarized (LP) modes. However, due to the large index contrast, the waveguiding modes in an optical MNF are non-degenerate with complex polarization properties. To observe the light field distribution intuitively, Fig. 7(c) shows electric fields of several waveguiding modes in a 600-nm-diameter silica MNF including HE11, TE01, HE21, and TM01 modes at 633-nm wavelength. It is worth mentioning that 1D crystalline nanowires, usually polygonal in cross section [see Fig. 7(d)], are also widely used as nano-waveguides. To quickly and accurately estimate their waveguiding properties (e.g., propagation constants), Bao et al. proposed a circular-area-equivalence scheme to treat a polygonal-cross-section nanowire as a circular MNF with an equivalent cross-section area[208], and showed its advantages of simplicity, intuition, high accuracy, and versatility in numerical calculation, even for cross sections with fewer sides. Following the fundamental waveguide theory above, we can judiciously investigate and design the optical waveguiding properties of optical MNFs.

3.2.

Basic Optical Waveguiding Properties

3.2.1.

Evanescent field and optical confinement of waveguiding modes

When the fiber diameter goes down to the subwavelength scale, the light waveguided in an MNF contains an evanescent wave carrying a significant fraction of the power, which penetrates a certain distance outside the core. The power distribution in an optical MNF can be written by the time-averaged Poynting-vector component along the z-axis[30,69,202]:

Eq. (9)

Sz=12(E×H*)·uz,
where E and H represent spatial distributions of electric and magnetic fields, respectively, and uz denotes a unit vector in the z-direction.

For the fundamental mode, the z-component of the Poynting vector is obtained as inside the core (r<2/D):

Eq. (10)

Sz,in=12(ε0μ0)12kn12βJ12(U)[a1a3J02(UR)+a2a4J22(UR)+1F1F22J0(UR)J2(UR)cos(2ϕ)],
and outside the core (r2/D):

Eq. (11)

Sz,out=12(ε0μ0)12kn12βK12(W)U2W2[a1a5K02(WR)+a2a6K22(WR)12ΔF1F22K0(WR)K2(WR)cos(2ϕ)],
with

Eq. (12)

a1=F212,a2=F2+12,a3=F112,a4=F1+12,a5=F11+2Δ2,a6=F1+12Δ2,F1=(UWV)2[b1+(12Δ)b2],F2=(VUW)21b1+b2,Δ=n1n2n1,b1=12U{J0(U)J1(U)J2(U)J1(U)},b2=12W{K0(W)K1(W)+K2(W)K1(W)}.

The fractional power inside the core (i.e., the percentage of the confined light power inside the core) is defined as

Eq. (13)

η=0D/202πSz,indA0D/202πSz,indA+D/202πSz,outdA,
with dA=r·dr·dϕ.

For reference, Fig. 8(a) presents the power distribution of the HE11 mode at 633-nm wavelength in silica MNFs with diameters of 800, 400, and 200 nm[32,59], clearly showing the increasing fractional power outside the fiber core with decreasing fiber diameter. While an 800-nm-diameter MNF exhibits good optical confinement at 633-nm wavelength with major energy inside the core, a 200-nm-diameter MNF leaves a large amount of light (>90% in power) outside the core as evanescent waves. For optical and photonic applications, a tight optical confinement is beneficial for reducing the modal diameter and increasing the integrated density of optical circuits with less crosstalk, while a large evanescent field is helpful for near-field energy exchange between the MNFs and other 1D optical waveguides within a short interaction length, as well as for improving the sensitivity of the MNF-based sensors and optomechanical devices.

Fig. 8

Optical waveguiding properties of silica MNFs. (a) z-direction Poynting vectors of the fundamental mode in silica MNFs with different diameters at 633-nm wavelength in 3D view (upper row) and 2D view (lower row)[30,32]. (b) Diameter dependence of the waveguide dispersion of fundamental modes in silica MNFs at the wavelengths of 633 nm and 1.5 µm, respectively[30]. (c) Calculated longitudinal electric-field intensity distribution of two evanescently coupled parallel 350-nm-diameter silica MNFs at 633-nm wavelength[154]. The overlapping length between two MNFs is 4.8 µm. (d) Schematic diagram of two evanescently coupled identical parallel MNFs (upper row)[190]. The lower row shows calculated cross-sectional electric-field intensity in the x- and y-polarizations of fundamental modes in the MNF j (j=1 or 2). The radius a of two identical MNFs is 200 nm, and the input light wavelength is 800 nm. (e) Schematic illustration of the crosstalk in two intersecting silica MNFs[213]. (f) Calculated output patterns of 400-nm-diameter silica MNFs with flat, 30°-tapering-angled, and 60°-tapering-angled endfaces in air[195]. The input light wavelength is set to be 633 nm. The white-line rectangles in (c) and (f) map the topography profile of the MNFs.

PI_3_1_R02_f008.png

3.2.2.

Engineerable waveguide dispersion

As the subwavelength-diameter MNF enables a considerably high fraction of the mode power outside the core, a strongly diameter-dependent waveguide dispersion can be achieved with a magnitude much larger than that in a standard fiber. The group velocity dispersion (GVD, Dw) of an optical MNF is defined as

Eq. (14)

Dw=d(vg1)dλ,
with group velocity given by

Eq. (15)

vg=cn12·βk·112Δ(1η).

Based on Eqs. (14) and (15), Fig. 8(b) gives the dispersion of the fundamental mode in air-clad MNFs as a function of the MNF diameter at the wavelengths of 633 nm and 1.5 µm[30]. One can clearly see that the GVD in the normal dispersion regime can go up to ns/(nm·km) level, while that of the weakly guided glass fibers is on the order of tens of ps/(nm·km). Moreover, the large waveguide dispersion in an MNF can also dominate over the material dispersion. To be specific, Dw of an 800-nm diameter MNF is about 1.4  ns/(nm·km) at 1.5-μm wavelength, which is 70 times larger than that of the material dispersion. Therefore, we can finely tailor the total dispersion (combined material and waveguide dispersions) to be zero, positive, or considerably negative by controlling the MNF diameter and length. In practice, effective dispersion management using subwavelength-diameter MNFs has been widely applied for supercontinuum generation[78,80], FWM[75], ultrafast fiber lasers[82,83,209,210], and quantum optics[211,212].

3.2.3.

Near-field coupling and crosstalk between two MNFs

The aforementioned waveguiding properties are basically appropriate for single optical MNFs. The investigation of the mutual interaction of waveguiding modes between two optical MNFs is also important for photonic applications. When two parallel MNFs are in close contact, the evanescently coupled MNFs with high index contrast cannot be treated as weakly coupled systems. In this case, it is difficult to perform the calculation analytically and thus numerical methods are more appropriate. In 2007, Huang et al. investigated evanescent coupling between two air-clad parallel MNFs by means of the finite-difference time-domain (FDTD) simulation[154]. Figure 8(c) presents the longitudinal electric-field intensity distribution of two parallel evanescently coupled silica MNFs (350 nm in diameter) at 633-nm wavelength. For the case of two parallel MNFs with a certain distance [see Fig. 8(d), upper row], Le Kien et al. investigated the coupling of fundamental waveguiding modes in two identical parallel MNFs based on the coupled mode theory[190]. The lower row of Fig. 8(d) shows the cross-sectional profiles of the electric-field intensity distributions at 800-nm wavelength in the x- and y-polarized fundamental modes of a single MNF (400 nm in diameter), which are symmetric with respect to the principal axes x and y. On this basis, Le Kien et al. studied the optical force and optical trap of an atom around the middle between two coupled identical parallel MNFs over the next few years[191193]. The optical forces between the MNFs were attractive for the fields of symmetric modes (i.e., even modes) and repulsive for the fields of antisymmetric modes (i.e., odd modes). For a ground-state cesium atom around the middle of two coupled identical parallel optical MNFs, a net trapping potential with a significant depth of about 1 mK, a large scattering-limited coherence time of several seconds, and a large recoil-heating-limited trap lifetime of several hours can be obtained by properly choosing realistic parameters. Such a twin-MNF structure can also be used for highly efficient single-photon collection, with optimal coupling efficiency >50%[194]. When the mode overlapping and coupling between two nanowaveguides exceed a certain degree, strong mode coupling occurs. In 2022, in a coupled nanowire pair (CNP), Wu et al.[196] reported strong mode coupling induced slit waveguiding mode, in which an ultra-confined optical field was generated with an optical confinement down to sub-1-nm level (λ/1000) and a peak-to-background ratio of 30  dB. Shortly after, Yang et al. proposed a waveguiding scheme to generate such a sub-nanometer-confined optical field using a tapered optical fiber[197], showing great flexibilities for narrow linewidth, broadband tunability, and ultrafast pulsed operation. More recently, Yang et al. reported a similar mode-coupling-induced nano-slit mode in a coupled glass MNF system, showing the possibility of generating a sub-nm-thick blade-like optical field[198]. With the same scale as a small molecule, this kind of confined optical field is promising for super-resolution nanoscopy, atom/molecule manipulation, and ultra-sensitive detection.

Besides parallel MNFs, coupling between two intersecting MNFs has been investigated. In 2019, Li et al. investigated the crosstalk between two intersecting optical MNFs[213], as schematically shown in Fig. 8(e). When the intersection angle is large enough (e.g., >60°), the crosstalk of two single-mode MNFs can be better than 20  dB. For a perpendicular intersection (i.e., intersection angle of 90°), the crosstalk is minimized to be better than 33  dB. Meanwhile, the crosstalk is not only intersection-angle-dependent, but also related to the MNF diameter and separation of two MNFs. Following these results, it is possible to design a close arrangement of the optical MNFs with acceptable crosstalk.

3.2.4.

Endface output patterns

As a kind of quasi-1D optical waveguide structure on the micro/nanoscale, glass MNFs with diameters smaller than the light wavelength are promising for tailoring endface output patterns, which is highly desirable in many photonic applications like subwavelength-dimension light beams[214], optical probes[215], and point sources[216,217]. In 2007, Ma et al. exploited a near-field scanning optical microscope (NSOM) to scan the endfaces of glass MNFs (including silica and tellurite MNFs) and obtained the endface output patterns of the MNFs on a MgF2 substrate[218]. To further investigate the longitudinal-field-intensity distribution of the MNFs, Wang et al. numerically calculated the output patterns of the MNFs at 633-nm wavelength by means of a three-dimensional finite-difference time-domain (3D-FDTD) approach[195]. For reference, Fig. 8(f) shows the calculated longitudinal field-intensity distribution of 400-nm-diameter MNFs with flat and tapered endfaces. Compared with standard optical fibers, glass MNFs possess much lower reflectance (e.g., with a flat endface, the reflectance is about 2% in glass MNFs and 4% in standard optical fibers). For the MNFs with tapered endfaces, the light output from the tapered tip spreads out symmetrically along the propagation direction. Interestingly, the smaller tapering angle yields a larger divergence of the output from the tapered tip, which is well suitable for the generation of point sources.

3.3.

Optical Loss

For standard single-mode fibers, the minimum transmission loss (e.g., 0.14 dB/km at 1550-nm wavelength[219]) is generally determined by the fundamental scattering and intrinsic absorption of silica glass. As the diameter decreases to the subwavelength scale, due to the decreased mode area, a high-fraction evanescent field, and normally open-air clad of the MNFs, the optical transmittance is very sensitive to a number of factors, including surface roughness, diameter fluctuation, surface contamination, and micro-bending, resulting in an overall loss much higher than that of a standard silica fiber. Fortunately, in the past 20 years, benefitting from the great progress in the fiber fabrication and protection techniques, the optical loss of an MNF has been significantly reduced.

3.3.1.

Scattering loss

For a subwavelength-diameter optical MNF with a clean surface, the scattering loss mainly arises from the surface roughness. To be specific, the molten-drawn process of the MNFs thermally activates capillary waves that are frozen onto the surface at the glass transition temperature, inevitably leading to quasi-long-range correlations in the surface height (i.e., surface roughness)[103105,220222]. By a roughness-induced current model, in 2007 Zhai and Tong numerically calculated the roughness-induced scattering loss of silica, tellurite, phosphate, and silicon MNFs with sinusoidally perturbed surfaces[223]. As shown in Fig. 9(a), with the same roughness amplitude, the loss coefficient exhibits an oscillating dependence on the perturbation period, indicating that the distribution of the perturbation period (i.e., the correlation length) should be determined for accurate estimation. Shortly after, by treating the perturbed surface as a random field and determining the radiated power with averaging over the perturbation ensemble, Kovalenko et al. investigated the radiation losses of optical MNFs with random rough surfaces in the assumption of Gaussian statistics, and estimated the loss value based on an inverse-square perturbation power spectrum[224]. Around the same time, by considering the nonadiabatic intermodal transition, Sumetsky investigated the radiation loss of a subwavelength-diameter MNF introduced by a tiny intrinsic nonuniformity, and showed an exponential dependence of the loss on the MNF diameter[225], with an important conclusion that the allowed minimum diameter of an optical MNF is about 1/10 the vacuum wavelength of the guided light. Subsequently, with an experimental test on a tapered fiber, Sumetsky et al. demonstrated that a subwavelength-diameter MNF can exhibit a low loss only if its diameter is larger than a threshold value, which is primarily determined by the wavelength and the characteristic length of the long-range nonuniformity[226]. Later, Hartung et al. determined that a threshold diameter of a 10-mm-long MNF appears to be D=0.24λ[227]. In addition, considering that in some cases the MNFs are supported on a substrate, the radiation loss caused by the substrate has also been investigated[228,229]. Generally, to effectively alleviate or eliminate the optical loss induced by the substrate, the MNF should not be too thin and the refractive index of the substrate should be lower than the effective index of the MNF. It is also important to mention that the surface contamination of the MNF will give rise to a further increase in radiation loss[35]. Typically, the contamination originates from electrostatic and diffusive adhesion of ambient microparticles and molecules. Therefore, an isolation of optical MNFs to atmospherical contamination is necessary for practical use.

Fig. 9

Optical losses and absorption of optical MNFs. (a) Roughness-induced radiation losses in air-clad MNFs versus the perturbation period[223]. The amplitude of the surface roughness is assumed to be 0.2 nm and the wavelength of the input light is 1550 nm. (b) Mathematical simulation model of a circular 90° bent MNF[188]. Inset: topography profile of the bent MNF. (c) Electric-field intensity distributions in the xz plane (y=0) of a 450-nm-diameter MNF at a wavelength of 633 nm, with bent radii of (I) 5 µm and (II) 1 µm. The output mode profiles of the 5-μm and 1-μm bent MNFs at the P1 transverse cross planes in (b) are shown in (III) and (IV), respectively. The black solid lines map the topography profiles of the MNFs. (d) Bending losses of a 350-nm-diameter silica MNF (I-line, squares), 350-nm-diameter PS MNF (II-line, circles), and 270-nm-diameter ZnO MNF (III-line, triangles) at 633-nm wavelength (quasi-x and quasi-y polarizations) as functions of the bending radius. (e) Schematic diagram of combined effects of the fiber heating, mechanically tapering, and pulsed laser guiding processes on the structural changes of siloxane rings[234]. Heating and mechanical stretching processes break the six-membered rings into highly strained three-membered rings. Given the bandgap value of silica (9  eV), using a laser with photon energy in the range of 48  eV can break the highly strained three-membered rings and generate oxygen-dangling bonds. (f) Defect-absorption-induced temperature rise of a 1.2-μm-diameter silica MNF as a function of the waveguiding power around 1550-nm wavelength[54].

PI_3_1_R02_f009.png

3.3.2.

Bending loss

As freestanding waveguided structures, optical MNFs are required to be bent in some cases. As a benefit of the high-refractive-index contrast between the MNF and surrounding, waveguiding modes can be tightly confined inside the MNF, enabling a low bending loss through relatively sharp bends (e.g., a few micrometers in bending radius). In 2009, Yu et al. numerically calculated the bending loss of optical MNFs with circular 90° bends by means of 3D-FDTD simulations [see Fig. 9(b)][188]. It can be seen in Figs. 9(c) and 9(d) that when waveguiding a 633-nm-wavelength light, a bent silica MNF with a diameter of 350 nm and bending radius of 5 µm has an acceptable bending loss of 1 dB/90°. However, as the bending radius reduces to 1 µm, a serious energy leakage occurs around the bending region [see Fig. 9(c)]. In such a sharp bend, to reduce the bending loss, one has to replace the silica MNF with higher-refractive-index MNFs such as PS and ZnO MNFs [see Fig. 9(d)]. Another possible approach proposed by Yang et al. recently is to rearrange the mode field by placing Au NPs around the inner side of the bent MNF[189]. Owing to the localized surface plasmon resonance (LSPR), a considerable fraction of energy can be confined to the interface of the Au NPs and the optical MNF, leading to a reduction in the bending loss. In addition, it is found that, compared with suspended MNF, in a substrate-supported MNF, the bending radius should be larger to guarantee a low propagation loss[230].

3.3.3.

Optical absorption

As is well known, the intrinsic absorption of silica glass is very low in its optically transparent window. However, in the high-temperature taper-drawing process, the six-membered rings of the amorphous silica are likely to deform to three-membered rings (i.e., strained Si-O-Si bonds), as the precursor sites for Si and O dangling bonds [see Fig. 9(e)][231234]. It has been demonstrated that the surface dangling bonds of the MNFs will be photoactivated by a pulsed laser/high-power CW laser, and act as point defects[54,233,234]. Such point defects can be identified by photoluminescence (PL) emission spectra. The point defects on the MNF surface will give rise to an increase in surface roughness and optical absorption, resulting in additional optical losses. It is worth mentioning that the photothermal effect induced by optical absorption is prominent when the MNFs are operated under high optical power[54]. As shown in Fig. 9(f), an evident temperature rise in a high-power waveguiding MNF was measured using a knot resonator (e.g., the temperature of the MNF operating at 5 W was 152°C).

3.3.4.

Progresses in loss reduction in optical MNFs

With the improvement of the fabrication and characterization techniques, the measured waveguiding losses of as-fabricated MNFs, especially of subwavelength-diameter MNFs, have been effectively reduced over the last 20 years. As summarized in Fig. 10, since the first reported experimental loss of 0.1  dB/mm at 1550-nm wavelength in a subwavelength-diameter silica MNF in 2003[29], the loss has been reduced quickly to a level of 0.001 dB/mm[31]. In 2014, Hoffman et al. reported the lowest loss of 2.6×105  dB/mm at 780-nm wavelength in a 500-nm-diameter silica MNF[37], which is orders of magnitude lower than all other available nanowaveguides [see Fig. 11]. For long MNFs, in 2020 Yao et al. demonstrated ultra-long subwavelength-diameter MNFs (e.g., 15 cm in length) with optical transmittance as high as 97.5% at 1550-nm wavelength[40]. In the same year, Ruddell et al. reported silica MNFs with diameters below 400 nm and transmission higher than 99.9% at 852.3-nm wavelength[53], which were used for constructing high-finesse fiber FP cavities. Besides silica, optical MNFs made of other glass materials have also been extensively studied. For example, in 2006, Brambilla et al. reported the fabrication of compound-glass (e.g., lead-silicate and bismuth-silicate) subwavelength-diameter optical MNFs[55], with measured losses of 100 to 101  dB/mm at 1550-nm wavelength. In 2007, Mägi et al. reported an overall tapering loss of 3 dB for a biconical As2Se3 fiber taper (1.2 µm in diameter, 18 mm in MNF length, and 164 mm in total length) around 1550-nm wavelength[57]. In 2012, Baker and Rochette measured a propagation loss of 5.2×102  dB/mm at 1530-nm wavelength in a PMMA-cladded chalcogenide MNF (450 nm in core diameter, 13.7 µm in PMMA-cladding diameter)[235]. More studies on optical loss in chalcogenide MNFs have been reported[236].

Fig. 10

Optical losses of silica and As2Se3 MNFs over the last 20 years[29,31,33,37,4143,53,143,211,235237].

PI_3_1_R02_f010.png

Fig. 11

Optical propagation losses of typical optical micro/nanowaveguides with corresponding effective mode areas[32,37,235,238240].

PI_3_1_R02_f011.png

3.3.5.

High-power optical waveguiding

Although low-loss optical MNFs have been well-developed since their first experimental demonstration, most of them are operated in low-power regions, for example, <0.1W in CW or averaged power. To further enhance the interaction between the waveguiding mode and matters within the mode field, the most straightforward method is increasing the waveguiding power in the MNFs. Recently, Zhang et al. reported high-power CW optical waveguiding in high-quality silica MNFs around 1550-nm wavelength[54]. As seen in Fig. 12(a), the measured output power (Pout) of a 1.1-μm-diameter changes quite linearly with the input power (Pin), and the MNF maintains a high transmittance (>95%) with waveguided power up to 13 W and a fiber diameter down to 410 nm (i.e., 1/4 of the vacuum wavelength). At such high power, the maximum power density inside the MNF can be higher than 23  W/μm2 [see Fig. 12(b)]. The ultralow absorption of the silica fiber (used as the preform), high precision in the taper-drawing process, and high cleanliness of both taper-drawing and testing environment enable 10-W-level optical waveguiding in silica MNFs. Furthermore, there are no predominant single scattering points on the surface of MNFs when waveguiding a high-power CW light, indicating that the upper limit of the waveguiding power in an MNF will be higher. By measuring the power-dependent temperature rise in the MNF, a damage threshold of 70 W was predicted [see Fig. 9(f)].

Fig. 12

High-power CW optical waveguiding in subwavelength-diameter silica MNFs[54]. (a) High-power optical transmittance of a 1.1-μm-diameter MNF around 1550-nm wavelength, with a CW waveguided power from 0 to 13 W. (b) Calculated diameter-dependent maximum power density in the MNFs at a waveguided power of 1 W. Insets: cross-sectional power density distribution of 0.5-μm-diameter and 1.1-μm-diameter silica MNFs.

PI_3_1_R02_f012.png

3.3.6.

Packages of optical MNFs

As we discussed previously, the surface contamination of the MNFs (e.g., dust and particulate) will introduce additional optical propagation loss. Generally, the degradation is more prominent as the MNF becomes thinner. For the purpose of long-term usage and stability of glass MNFs, adequate protection is highly desirable. Typically, the package of glass MNFs can be categorized into three types: polymer embedding, air-tight sealing, and vacuum packaging.

  • (1) Polymer embedding

    Low-refraction polymer is widely exploited to embed and protect glass MNFs from optical degradation. In 2007, Vienne et al. proposed to embed MNFs in a low-index fluoropolymer matrix, and experimentally demonstrated polymer-embedded MNF resonators with Q-factors of 12000[241]. In 2008, Xu and Brambilla compared the effectiveness of packaging of the MNFs in a perfluoro polymer (i.e., Teflon)[242]. While a bare MNF decreased with a rate of 0.2  dB/h in optical transmittance, the embedded MNF remained basically unchanged within 6 days. After that, a variety of low-loss, low-refractive-index materials have been employed for MNF packaging, including polydimethylsiloxane (PDMS)[243247], Teflon[248,249], hydrophobic aerogel[250], ultraviolet (UV) curable polymer[251,252], and high-substitution hydroxypropyl cellulose[253,254]. Figure 13(a) illustrates a silica MNF embedded in a PDMS film[247], showing the advantages of a small footprint, high flexibility, and conformability to non-flat surfaces (e.g., a human hand).

  • (2) Air-tight sealing

    Sealing the MNFs inside an air-tight box has also been commonly adopted. For example, in 2021 Bouhadida et al. carried out long-term and repeatable measurements of optical transmittance of a 1-μm-diameter MNF at 1.5-μm wavelength[255]. They found that when the MNF was put in an airtight box, the degradation rate of the optical transmittance was as low as 6.3×103%/day in several months. According to the degradation rate, they predicted an acceptable decrease of 10% in the optical transmittance after more than 4 years. Recently, Zhang et al. transferred an as-fabricated 1.2-μm-diameter MNF into a 3D-printed acrylic box protected by high-purity nitrogen gas [see Fig. 13(b)][54]. The encapsulated MNF shows an ability to waveguide a high-power CW light around 1550-nm wavelength for long-term operation. Specifically, the MNF was tested at 12-W power continuously for 20 min every day for 2 months. Over the whole test period of 2 months, the optical transmittance of the MNF remained around 95%, as shown in Fig. 13(c). Apart from an airtight box, other sealed chambers have also been applied to satisfy various experimental requirements, for example, capillary[256258] and stainless steel tube[92,259].

  • (3) Vacuum treating

    For atom optics, silica MNFs are usually installed in a vacuum chamber. During the installation procedure, a flux of argon gas is maintained to protect the MNF surface from pollution[260]. After the introduction of atom vapor, one of the major challenges is that the warm atoms will accumulate on the MNF surface, leading to optical scattering of waveguide modes and consequently, a drastic degradation in optical transmission. For instance, Lai et al. reported that the transmission of a sub-500-nm-diameter silica MNF degraded to 1.5% of its initial value in rubidium vapor in a full 2700-s run[261]. In this scheme, the base pressure and temperature in the vacuum chamber were 107  Torr and 60°C, respectively. To alleviate the atom accumulation, they designed a heating unit to raise the surface temperature of the MNF, enabling the stable preservation of the MNF with a relatively high optical transmittance (30%) surrounded by a high-density rubidium vapor. Afterwards, Lamsal et al. employed metastable xenon as a promising alternative to rubidium[262,263], in which a sub-500-nm-diameter silica MNF resided in a vacuum chamber backfilled with xenon gas. Using this low-pressure system (around 30 mTorr), they demonstrated a complete lack of optical transmission degradation in a 350-nm-diameter silica MNF over several hours[263].

    In addition, safe and convenient handling of the MNFs without breaking is also required in experiments. For this purpose, a U-shaped bracket is typically employed for detaching an as-fabricated MNF from the taper-drawing system [see Fig. 13(d)] and transferring it to other places, such as an air-tight box for surface protection [see Fig. 13(e)].

Fig. 13

Encapsulation of optical MNFs. (a) Schematic of an MNF embedded in a PDMS film on a glass substrate[247]. Inset: photograph of an MNF-embedded PDMS patch attached to a human hand. (b) Photograph of an optical MNF sealed in an airtight 3D-printed acrylic box, filled with high-purity nitrogen gas[54]. (c) Long-term optical transmission of the MNF presented in (b) around 1550-nm wavelength. The waveguided power of the MNF is 12 W. (d) Photograph of an as-fabricated MNF mounted on a U-shaped bracket, with two standard fiber pigtails fixed on both sides of the bracket through the glue. (e) Photograph of an MNF sealed in an air-tight box.

PI_3_1_R02_f013.png

3.4.

Nonlinear Optical Properties

The above-mentioned optical properties are in the realm of linear optics, in which the induced polarizability P has a linear relationship with the electric-field intensity. This is valid only when the power density of the waveguiding modes is low. For a high-power optical waveguiding MNF (in pulsed or high-power CW), the nonlinear effect should be considered, in which the optical response could be described by the relationship[264]

Eq. (16)

P=ε0[χ(1)E+χ(2)E2+χ(3)E3+],
where ε0 is the vacuum permittivity, χ(1) is the linear susceptibility, and χ(2) and χ(3) are the second- and third-order nonlinear optical susceptibilities, respectively. In the nonlinear optical processes, χ(2) is associated with second-order nonlinear effects such as SHG, sum-frequency and difference-frequency generation (SFG and DFG), and optical parametric oscillation. χ(3) is related to third-order nonlinear effects including supercontinuum generation, THG, BS, SRS, FWM, and optical Kerr effect.

As we reviewed in Section 3.2, the tight optical confinement, long interaction length, and large diameter-dependent dispersion make optical MNFs excellent candidates for nonlinear optical processes. For example, the power density of the evanescent field can be enhanced to 108  W/cm2 in a 340-nm-diameter silica MNF waveguiding a 1-W 780-nm-wavelength light[20]. Figure 14(a) shows the wavelength-dependent GVD in silica MNFs. For a given MNF diameter, there is a maximum value of the GVD (on the order of thousands of ps·nm1·km1) in the normal dispersion region. Generally, the optical nonlinearity of photonic waveguides is described by the nonlinear parameter γ as[68]

Eq. (17)

γ=n2ω0cAeff,
where n2 is a nonlinear-index coefficient (for silica, n2=2.7×1020  m2/W[71]), ω0 is the angular frequency, c is the speed of light in vacuum, and Aeff is known as the effective mode area of the waveguiding modes, which is expressed as[68]

Eq. (18)

Aeff=(|E(x,y)|2dxdy)2|E(x,y)|4dxdy.

Fig. 14

Nonlinear optical properties of optical MNFs. (a) Wavelength dependence of the GVD with different MNF diameters. (b) Nonlinear coefficient of silica MNFs versus the fiber diameter at 532-nm wavelength. Spectra of the (c) SHG and (d) THG in a silica MNF (779 nm in diameter, 7 cm in length) pumped by a 5-W-power CW light[54]. The optimal phase matching of the SHG and THG is achieved at wavelengths of 1558.2 and 1572.5 nm, respectively. SH, second harmonic; TH, third harmonic. Insets of (c) and (d) show optical microscope images of output spots of the SH and TH signals at the output end of the standard fiber connected with the MNF, respectively. (e) Supercontinuum generation in a silica MNF pumped by 532-nm-wavelength ns pulses[31], with output far-field patterns from the MNF at (I) low and (II) maximum powers. The pattern in (II) was passed through 10-nm bandpass filters at the center wavelengths of (III) 633, (IV) 589, and (V) 450 nm.

PI_3_1_R02_f014.png

Figure 14(b) plots the diameter-dependent nonlinear parameter of silica MNFs at a wavelength of 532 nm. It is observed that the 350-nm-diameter MNF has a maximum nonlinear parameter of 1500  W1km1, and the dispersion coefficient falls in the vicinity of the zero-dispersion region. For comparison, the zero-dispersion wavelength of the silica MNFs can be much shorter than that of a standard optical fiber (around 1310-nm wavelength). Moreover, the zero-dispersion wavelength is also diameter-dependent, which provides great convenience for dispersion management in nonlinear processes. In this subsection, we will provide an overview of the nonlinear optical properties in optical MNFs.

3.4.1.

Pulse propagation

Generally, for nonlinear applications, ultrashort pulses with high peak power are used to achieve a high nonlinear conversion efficiency. To study the femtosecond pulse propagation in a subwavelength-diameter optical MNF, in 2004 Kolesik et al. proposed a theoretical model for simulation according to a corrected nonlinear Schrödinger equation[265]. Based on the theoretical framework, supercontinuum generation in optical MNFs can be accurately modeled[79,266]. In 2005, Foster et al. experimentally demonstrated the soliton-effect self-compression of ultrafast pulses from 70 to 6.8 fs[81], corresponding to a compression factor of 10.29 and a quality factor of 0.73. Using a nonlinear-envelope equation, they predicted that self-compression could be further down to single-cycle duration. In 2012, Lægsgaard developed a full-vectorial nonlinear propagation formalism for studying the spectrum evolution of short broadband pulses in fiber tapers[267]. In this method, they proposed a perturbative scheme for interpolating fiber parameters along the taper, which provides higher accuracy. Benefitting from the large effective nonlinearity and broad region of anomalous GVD of the MNFs, one can utilize optical MNFs with specific diameters to compress the pulse.

3.4.2.

SHG and THG

Harmonic generation including SHG and THG is a kind of nonlinear optical process, in which new signal sources with short wavelengths can be generated. With the rapid development of fiber technology, fiber-based harmonic generation has received much attention, showing the advantages of small footprint, long interaction length, good stability, and low cost. Generally, the SHG is very weak in a silica fiber due to the central-symmetry nature of an amorphous silica. However, when the fiber diameter decreases to the subwavelength scale, the surface dipole and bulk multipole nonlinearities will contribute evidently to the second-order nonlinearity. The theoretical framework of the SHG in MNFs has been well-studied since 2010[268270]. For efficient SHG, intermodal phase matching and high mode overlapping are highly desirable. In the small-signal limit, the SHG process can be modeled by the following equation:

Eq. (19)

dA2dziρ2A12exp(iΔβz)=0,
where A1 and A2 are the amplitudes of the fundamental and SHG modes, Δβ=β22β1 is the propagation constant mismatch between the SHG signal and the pump, and ρ2 represents the overlap between the fundamental and the second-harmonic modes, which is expressed as

Eq. (20)

ρ2=ω22A12dre2*·P(2)Redr[e2*×h2]z,
where ω2 is the angular frequency of the SHG signal, P(2) is the second-order nonlinear polarization (including both dipole contributions from the surface and multipole contributions from the bulk), r is the vector normal to the surface, and the electric and magnetic fields of the SH modes are expressed as E(r,ω2)=A2e2(r,ω2)exp[i(β2zω2t)] and H(r,ω2)=A2h2(r,ω2)exp[i(β2zω2t)].

In 2013, Gouveia et al. experimentally exploited a 700-nm-diameter silica MNF to realize the intermodal phase-matching SHG, with conversion efficiency up to 2.5×109 when pumped with 1550-nm-wavelength pulses (90 W in peak power)[72]. By assembling an MNF into a loop resonator that incorporated the phase-matched region of the MNF, they further improved the efficiency by 5.7 times. Very recently, Zhang et al. proposed to accomplish perfect intermodal phase matching and maximum mode overlapping between the fundamental and the harmonic modes by precisely controlling the fiber diameter and finely tuning input light wavelength. Using an 11.3-W 1558.2-nm-wavelength CW light [see Fig. 14(c)], they demonstrated CW-pumped SHG, with conversion efficiency up to 8.2×108, which is higher than those pumped by short pulses[54]. Besides a single MNF structure, other MNF-based structures have been reported for the enhancement of the SHG effect. For example, in 2013 Luo et al. proposed a slot MNF configuration to enhance surface power density and nonlinearity, showing a 25-fold increase in the SHG conversion efficiency as compared to a circular-cross-section MNF[271]. In 2018, Wu et al. employed an optical MNF coupler to achieve quasi-phase matching by coupling compensation. High-efficiency SHG was experimentally demonstrated, with conversion efficiency four orders of magnitude higher than that of each individual MNF[272].

Compared with the SHG, THG is more widely studied in optical MNFs. In 2003, Akimov et al. demonstrated the THG in a 2.6-μm-diameter silica MNF using 30-fs 1250-nm-wavelength pulses[70]. In the following 20 years, a series of theoretical and experimental studies have been carried out to optimize the conversion efficiency in nonlinear optical processes[71,273280].

For an air-clad optical MNF with a step-index profile, the third-order nonlinear susceptibility χ(3) is assumed to be z-independent, whose value is constant within the cross section of the MNF while zero outside of the MNF. Then the THG process can be modeled by the following coupled-mode equations[71]:

Eq. (21)

A1z=in2k[(J1|A1|2+2J2|A3|2)A1+J3A1*2A3eiδβz],A3z=in2k[(6J2|A1|2+3J5|A3|2)A3+J3*A13A3eiδβz],
where A1 and A3 are the amplitudes of the fundamental and THG modes, δβ=β33β1 is the mismatch of the propagation constant between the third-order harmonic signal and the pump, and Ji is nonlinear overlap integrals (as defined in Ref. [71]). Among them, J1 and J5 govern the self-phase modulation of the pump and the harmonic signal, J2 is related to the cross-phase modulation, and J3 represents the overlap between the pump and the third-harmonic modes.

To achieve efficient THG in an optical MNF, the intermodal phase mismatch (i.e., the propagation-constant mismatch between the high-order modes of the third-order harmonic signal and the fundamental mode of the pump) should be as small as possible, and the overlap integral J3 should be sufficiently large. In 2005, Grubsky and Savchenko theoretically predicted that with an appropriate MNF diameter, an ideal conversion efficiency could be achieved as high as 50% in a 1-cm-long silica MNF[71]. However, it is difficult to obtain such high efficiency because the inherent nonuniformity of optical MNFs may adversely affect the perfect phase matching. In recent years, several effective approaches to realizing quasi-phase matching between the fundamental pump mode and the third-harmonic modes have been reported by Jiang et al. and Hao et al., such as the employment of a counter-propagating pulse train[277], nonlinear phase modulation[278,279], and mechanical strain[280]. Indeed, the THG conversion efficiency excited by short laser pulses reported so far is basically on the order ranging from 107 to 104. Additionally, pumped with a high-power CW light, high-efficiency THG could also be observed in a 779-nm-diameter MNF (1572.5 nm in fundamental light wavelength), as presented in Fig. 14(d). When the waveguided power of the MNF was increased to 11.3 W, the conversion efficiency was measured as 4.9×106, falling within the range of typical results obtained with short pulses. Since silica MNFs have the potential to waveguide a higher-power light (both in CW and pulsed), the higher conversion efficiency of the harmonic generation is foreseeable in the future.

3.4.3.

Four-wave mixing

The FWM process is a third-order optical nonlinear process containing the creation of signal and idler photons (ωs and ωi) and the annihilation of two pump photons (ωp1 and ωp2) simultaneously. This process is governed by the conservation of the energy and momentum: ωs+ωi=ωp1+ωp2, and βp1+βp2=βs+βi. The tailorable dispersion characteristics of optical MNFs are beneficial for efficient FWM. In 2012, Li et al. reported for the first time the cascaded FWM in optical MNFs pumped by two synchronized picosecond lasers (around 850-nm wavelength)[75]. They showed that the spectrum range could span from several hundreds of nanometers to almost one octave, depending on the MNF diameter, pump power, and wavelength detuning of the two pumps. In practice, the FWM in an optical MNF has been widely applied in quantum optics and nonlinear optics. For example, in 2013 Cui et al. utilized spontaneous FWM in a 15-cm-long MNF to generate correlated photon pairs[211]. Likewise, in 2019 Kim et al. realized the generation of photon pairs via a spontaneous FWM process using a 12-cm-long 615-nm-diameter MNF[281]. In 2021, Delaye et al. reported the emission of photon pairs by the FWM in a silica MNF, with a high coincidence to accidental ratio and high pair emission rate[212]. On the other hand, Abdul Khudus et al. demonstrated parametric amplification through the FWM in silica MNFs with a wavelength band over 1000 nm in 2016[282].

3.4.4.

Brillouin scattering

BS is known as a third-order nonlinear process that involves the interaction between an electromagnetic wave and an acoustic wave. An incident photon at frequency ω is scattered into an up- or downshifted photon with frequencies ω±ΩB depending on the phase-matching condition, where frequency ΩB is the frequency of the acoustic field. Since its discovery in optical fibers in 1972[283], the BS has been extensively studied in both fundamental and applied sciences[284]. Over the past decade, the advent of optical MNFs has driven immense scientific interest in studying the photon-phonon interaction in subwavelength-dimension structures. Compared with standard optical fibers, optical MNFs with low dimension and hard mechanical boundary conditions enable the enhancement of optical nonlinearity and tightly confined modes of both photons and phonons. It has been reported by Beugnot et al. in 2014 that in a subwavelength-diameter MNF, two kinds of acoustic modes will be excited during the backward BS process: hybrid acoustic wave (HAW) and surface acoustic wave (SAW) modes[73]. The SAWs propagate at a velocity between 0.87 and 0.95 of a shear wave velocity VS (for fused silica, VS=3400  m/s), leading to new optical sidebands down-shifted from 6 GHz in the spectrum. The HAWs propagate at an intermediate speed between shear and longitudinal waves with acoustic frequencies of 9  GHz. Interestingly, the SAWs in the MNFs are inherently sensitive to surface features and defects, which offers attractive potential for optical sensing, detection, strain measurement, and optomechanics[63,74,92,285]. Another novel effect, the so-called BS self-cancellation, was demonstrated experimentally by Florez et al. in 2016[286]. Specifically, they scanned the MNF diameter to precisely control the acoustic and optical mode profiles, where the photo-elastic and moving-boundary effects canceled out exactly. In 2018, Chow et al. utilized a phase correlation distributed Brillouin approach to experimentally demonstrate the presence of surface and hybrid acoustic waves at distinct fiber locations[287]. Apart from the backward BS, the forward BS in silica MNFs has been investigated over the last few years[171,288,289]. In 2018, Jarschel et al. constructed a pump and probe forward BS setup to excite the fundamental torsional-radial acoustic mode in a silica MNF[171]. The frequency shift of the torsional-radial acoustic mode can be used to characterize the MNF diameter, as has been mentioned in Section 2.1.4. After that, two theoretical frameworks were reported by Cao et al. on the inter-mode forward BS[288] and Brillouin gain characteristics[289] in silica MNFs. Very recently, Xu et al. realized strong forward intermodal optomechanical interactions in a few-mode optical MNF, with long phonon lifetimes (>2  μs) and strong coupling (>400  W1m1)[93].

3.4.5.

Stimulated Raman scattering

The SRS is an inelastic third-order nonlinear process in which energy is transferred from input photons to molecular vibrations. Unlike the stimulated BS with a photon-to-acoustic phonon process, the SRS involves a photon-to-optical phonon interaction. In the SRS process, a Stokes (anti-Stokes) photon at a longer (shorter) wavelength with respect to the pump wavelength will be generated. As discussed in Section 3.2, subwavelength-diameter optical MNFs possess a high-fraction evanescent field with high optical intensity. Such a unique evanescent field can create a strong interaction with external materials (with high Raman gain). By this way, in 2013 Shan et al. theoretically investigated[76] and experimentally achieved[77] the SRS in a liquid (e.g., ethanol and mixture of toluene and ethanol) pumped by the evanescent fields of silica MNFs. In 2019, Bouhadida et al. developed the external Raman conversion efficiency to 60% with high reproducibility by optimizing the MNF diameter[290]. Besides the liquid surroundings, the evanescent-wave SRS can also be achieved with a silica MNF immersed in the gas. It has been reported by Qi et al. in 2019 that the SRS efficiency of an MNF surrounded by hydrogen gas is orders of magnitude higher than that in a hollow-core photonic crystal fiber[291]. Relying on the SRS spectroscopy, a novel optical sensing and detection technique with fast response, high sensitivity, and wide dynamic range is on the rise.

3.4.6.

Supercontinuum generation

Supercontinuum generation is a kind of complex optical nonlinear process that involves self-phase modulation, Raman scattering, and FWM. Since its first experimental observation in borosilicate glass in 1970[292], the supercontinuum generation has received extensive attention. In 1976, Lin and Stolen demonstrated supercontinuum generation in a 19.5-m silica fiber using a 20-kW 10-ns dye-laser pulse[293]. In principle, to generate a supercontinuum source with a wide spectral range in an optical waveguide, the high optical power density of waveguiding modes, high optical nonlinearity of the waveguide, long nonlinear interaction length, and flat zero-dispersion region are required. The flexibly controllable dispersion and high nonlinear parameter of subwavelength optical MNFs just satisfy these demands[265,294] [see Figs. 14(a) and 14(b)]. In 2004, Leon-Saval et al. reported that supercontinuum spectra with a 400-nm broadening could be observed from 20-mm-long sub-micrometer optical MNFs pumped by 532-nm-wavelength ns pulses[31]. When the peak power is sufficiently high, a significant broadening of the supercontinuum spectrum occurs and the supercontinuum white-light source can be extracted from the output endface of a standard fiber connected with a silica MNF [see Fig. 14(e)]. Afterwards, supercontinuum generation in silica MNFs has been studied at the central wavelengths of 800[79,295] and 1064 nm[78]. From the numerical simulation, Hartung et al. also reported the possibility of using optical MNFs with normal dispersion behavior for coherent supercontinuum generation at deep-UV wavelengths[80]. The short length of the employed silica MNFs allows a low loss in the UV-wavelength edge, making them ideal candidates for deep-UV supercontinuum generation. Meanwhile, chalcogenide glasses have broadband intrinsic transparency (0.5 to 25 µm) and high optical nonlinearity (for the As2Se3, n2=1.1×1017  m2/W[296]) for supercontinuum generation from visible to infrared ranges.

Compared with silica MNFs, the chalcogenide MNFs taper drawn from standard chalcogenide fibers show greater potential in nonlinear optics. Besides the UV and visible spectrum regions, NIR and mid-infrared (MIR) supercontinuum sources generated from highly nonlinear chalcogenide-glass MNFs have received a lot of attention in the past 20 years. In 2008, Yeom et al. first demonstrated low-threshold supercontinuum generation around 1550-nm wavelength using an As2Se3 MNF. In the following years, several interesting investigations on the NIR supercontinuum generation in chalcogenide-glass MNFs have been reported[297299]. For the MIR supercontinuum generation, Marandi et al. achieved a spectrum broadening from 2.2 to 5 µm at 40 dB below the peak in 2012[300]. In 2014, Al-kadry et al. demonstrated two-octave MIR supercontinuum generation from 1.1 to 4.4 µm at 30  dB[301]. In 2015, using a 15-cm-length As2Se3-As2S3 chalcogenide MNF (1.9 µm in core diameter), Sun et al. realized the supercontinuum generation spanning from 1.5 µm to beyond 4.8 µm at 20  dB[302]. In 2017, Hudson et al. coupled 4.2-kW-power pulses into a polymer-protected As2Se3/As2S3 MNF and successfully demonstrated a spectrum spanning from 1.8 to 9.5 µm at 20  dB points (2.4 octaves)[303]. In the same year, Wang et al. reported a broadband spectrum spanning from 1.4 to 7.2 µm in chalcogenide tapered fibers pumped in the normal dispersion regime[304]. The generated MIR supercontinuum sources have found important implications in the applications of molecular spectroscopy, MIR frequency comb, early cancer diagnosis, and remote sensing[147,305,306].

4.

Mechanical Behavior

As an important property of an optical MNF, the mechanical behavior, basically depending on the material used, is different from that of standard optical fibers with typical diameters around 125 µm. Generally, similar to mechanical behaviors of many other 1D nanostructures that have been extensively investigated[307310], owing to the decreased defect size and density with decreasing fiber diameter, the silica MNF demonstrates enhanced mechanical properties regarding tensile strength and allowed strain, which is beneficial for flexible manipulation, characterization, and functionalization of the MNF. Meanwhile, the low geometric dimension, high material purity, and structural uniformity of the MNF make it an ideal platform for investigating the deformation behaviors of such covalently bonded nanostructures, as well as offering a bridge between experiments and molecular dynamics simulations. This section reviews the mechanical properties of both elastic and plastic deformations of silica MNFs, with emphasis on elastic modulus, tensile strength, and linear strain, which are within the elastic limit in most applications.

4.1.

Elastic Deformation

For an optical MNF with a diameter close to the wavelength of the waveguided light, the allowed minimum bending radius RBmin is determined by the bending loss[152,188]. Typically, for low-loss waveguiding, RBmin is much larger than the allowed minimum elastic bending radius REmin. Thus, in optical applications, almost all deformation of the silica MNF is elastic.

4.1.1.

Elastic modulus

The elastic modulus, which reflects the stiffness of solid materials, is a crucial indicator for describing the elastic properties, and it varies significantly among different materials. For amorphous silica, the elastic modulus measured in bulk glass is typically around 72 GPa. The question of whether the elastic modulus of amorphous silica changes with different physical sizes is a topic of interest, and a number of simulations and experiments have been conducted on low-dimensional structures, including films, micro/nanoparticles, and MNFs[311313]. Among these structures, the MNF is a widely investigated quasi-1D structure due to its excellent diameter uniformity, high surface quality, and convenience for measurement of mechanical properties. There is sufficient experimental evidence that silica MNFs with diameters larger than 50 nm still exhibit the elastic modulus close to bulk amorphous silica. For example, in 2004 Chen et al. determined the elastic modulus of MNFs in the diameter range of 23.3–133 µm by resonance vibration measurements, and obtained a mean value of (70±6)  GPa[314]. In 2006, using a scanning probe microscope, Silva et al. directly measured the elastic modulus of silica MNFs with diameters ranging from 280 to 1950 nm, which varied from 68 to 76 GPa[315]. In the same year, utilizing an atomic force microscope (AFM), Ni et al. measured the elastic modulus of ultrathin silica MNFs with diameters ranging from 50 to 100 nm, and obtained a modulus of (76.6±7.2)  GPa[316].

4.1.2.

Elastic strain and tensile strength

The allowed elastic strain is a measure of the mechanical strength of a structure. In the case of MNFs, elastic tensile strain and tensile strength are two key factors that attract considerable attention. There are two main reasons: firstly, in practical applications, it is challenging to directly apply axial compression strain or stress to MNFs given their quasi-1D structure. Instead, bending and stretching are the most typical forms of strain loading. Secondly, in terms of fracture, the fracture of MNFs usually occurs when the tensile strength exceeds the failure limit, and the allowed compression strength is typically higher than the tensile strength. The relationship between the elastic strain ε and corresponding tensile strength σ can be expressed as

Eq. (22)

σ=ζε,
where ζ is the elastic modulus of the fiber material, which is around 72 GPa for amorphous silica. Under elastic bending, the tensile strength can be estimated as

Eq. (23)

σ=ζD2RB,
where D is the MNF diameter and RB is the bending radius.

In 2003, by bending MNFs to the point of fracture, Tong et al. estimated that the fracture strengths of silica MNFs were typically higher than 5.5 GPa[29]. In 2009, Brambilla et al. directly measured the fracture strength of silica MNFs with diameters from 120 to 600 nm by loading vertical tension. As shown in Fig. 15(a), the measured fracture strength exceeds 10 GPa for most MNFs, with maximum values in MNFs with diameters of 100–200 nm. The strength of the MNF is higher than that of standard communication optical fibers (about 5 GPa)[317]. It is worth mentioning that under large elastic strain, similar to that observed in silica glass[318,319], nonlinear elasticity can also be observed in an MNF. For example, in 2019 through the backward BS in silica MNFs (660 and 930 nm in diameter, respectively), Godet et al. experimentally verified the third-order nonlinear behavior of elasticity when the tensile strain is larger than 2%[63].

Fig. 15

Mechanical properties of optical MNFs. (a) Dependence of fracture strengths of silica MNFs on the MNF radius r[61]. (b) SEM image of a plastically bent silica MNF (800 nm in MNF diameter)[321]. The sharp bent radius is less than 1 µm. (c) SEM image of a 170-nm-diameter tellurite glass MNF with sharp plastic bends[56]. (d) Maximum plastic elongation of silica MNFs. The horizontal dashed line (purple) indicates a reference line of 1%[323].

PI_3_1_R02_f015.png

Meanwhile, the strength of silica MNFs is dependent on their fabrication method. For instance, the strength of the MNF obtained by hydrogen flame heating seems relatively low compared with that obtained by electric heating or laser heating. This is because the water molecules produced by the hydrogen flame will partially enter the silica network through the chemical reactions (Si-O-Si+H2OSiOH+HOSi), forming defects that reduce the strength of MNFs[320]. Similarly, when the MNFs are exposed to the air for a prolonged period, water molecules and other molecules will attach to the fiber surface and may lead to an accumulated decrease in strength. Therefore, effective packaging and isolating an MNF from environmental contamination is also beneficial to maintain its excellent mechanical properties.

4.2.

Plastic Deformation

As an amorphous glass structure, a glass MNF can be plastically deformed under certain conditions. So far, two approaches have been studied for plastic deformation of MNFs: high-temperature annealing under elastic deformation and room-temperature deformation in ultrathin MNFs. Annealing is a versatile approach to plastically deforming glass MNFs with all diameters. When an elastically deformed MNF is annealed under high temperature, the strain can be released and the deformation can be transformed into permanent plastic deformation. In 2005, Tong et al. annealed an elastically bent silica MNF for 2 h at 1400 K in a vacuum (2×103  Pa) and obtained a permanent plastic bend[321]. As shown in Fig. 15(b), sharp plastic bends (the bending radius of less than 1 µm) can be achieved by performing the annealing-after-bending process. The annealing approach is also applicable to other types of glass MNFs. For example, Fig. 15(c) shows sharp plastic bends made on a 170-nm-diameter tellurite glass MNF[56]. Plastic deformation realized by such annealing-after-bending process can avoid long-term fatigue and fracture of sharply bent glass MNFs, and the geometry of the deformation can be designed and assembled before annealing.

For ultrathin silica MNFs, it is found that they can undergo plastic deformation at room temperature, and can be employed as an ideal platform for studying the deformation behaviors of these covalently bonded nanostructures. In such cases, low-density electron beams and slower strain rates are commonly used. For instance, in 2010 Zheng et al. showed that under room temperature (300  K) and a strain rate >104 per second, and moderate exposure to a low-intensity (<1.8×102  A/cm2) electron beam, a superplastic elongation >200% in tension was achieved in a 36-nm-diameter silica nanofiber[322]. In 2016, Luo et al. found that when the MNF diameter was below 18 nm, it would undergo a brittle-to-ductile transition at room temperature, and large tensile plastic elongation (up to 18%) could be realized at a strain rate ranging from 102 to 104 per second, as shown in Fig. 15(d)[323]. These studies may help understand the mechanical behaviors of low-dimensional amorphous structures.

5.

MNF-Based Applications

As a miniature fiber-optic platform, optical MNFs have been extensively studied over the past two decades. Their remarkable optical properties, including ultralow loss, tight optical confinement, high-fraction accessible evanescent fields, and engineerable waveguide dispersion, incorporated with a small footprint and excellent mechanical properties, have inspired a variety of possibilities from near-field coupling, passive optical components, optical sensing, and fiber lasers to nonlinear optics, optomechanics, and atom optics. Since a complete coverage of these possibilities will be miscellaneous and lengthy, in this section we introduce typical MNF-based applications in different categories. Also, as MNF-based nonlinear optics has been introduced in Section 3.4, we will not go into further detail here.

5.1.

Near-Field Optical Coupling

As has been discussed in Section 3.3, optical MNFs have the special advantages of low-loss waveguiding with tightly confined high-fraction evanescent fields, which makes them favorable for compact and high-efficiency near-field optical coupling with external structures on the micro/nanoscale. Assisted with micro/nanomanipulation, a variety of MNF-based coupling structures, including 1D optical waveguides, 2D materials, and 3D micro-cavities, have been demonstrated, as introduced in this section.

5.1.1.

1D optical waveguides

1D optical waveguides with (sub)wavelength-scale cross sections are the mostly used micro/nanoscale waveguiding structures. Benefitting from the strong evanescent field, the waveguiding modes of an optical MNF can be efficiently coupled into a 1D photonic or even plasmonic waveguide via near-field coupling. Figure 16(a) shows the near-field coupling of two tellurite glass MNFs with diameters of 350 nm (top arm) and 450 nm (bottom arm), respectively[56]. With a 633-nm-wavelength light launching into the 450-nm-diameter MNF from the bottom left arm, a certain portion of the light is coupled into and waveguided along the 350-nm-diameter MNF (top arm). By changing the coupling length and MNF diameters, the coupling efficiency can be optimized (e.g., >95% with a coupling length on 2-μm level[154]). The efficient evanescent-field coupling technique is also applicable to plasmonic nanowaveguides. In 2009, Guo et al. reported direct coupling of photonic (e.g., silica MNFs or ZnO nanowires) and plasmonic nanowaveguides (Ag nanowires) via near-field coupling [see Fig. 16(b)] with coupling efficiency up to 80%, and further demonstrated hybrid “photon-plasmon” functionalized components such as polarization splitters, Mach-Zehnder interferometers (MZIs), and micro-ring resonators[324]. By optimizing the coupling conditions, in 2013 Li et al. demonstrated a coupling efficiency of up to 92% in a silica MNF-Ag nanowire structure[325]. Similarly, a number of other types of 1D waveguides, including polymer MNFs [see Fig. 16(c)][110], semiconductor nanowires (e.g., CdS, CdSe, ZnO, and CdTe) [see Fig. 16(d)][326330], ice MNFs[126], niobium nitride (NBN) superconducting nanowire[331333], and on-chip silicon waveguides[334338], have been evanescently coupled using waveguiding MNFs to obtain efficient optical coupling.

Fig. 16

Near-field optical coupling with 1D micro/nanowaveguides using silica MNFs. (a) Optical microscope image of optical coupling of a 633-nm-wavelength light between two tellurite glass MNFs with diameters of 350 (top arm) and 450 nm (bottom arm), respectively[56]. Optical microscope images of a silica fiber taper coupled with a (b) 200-nm-diameter silver nanowire[324], (c) 450-nm-diameter polyacrylamide MNF doped with fluorescein sodium salt (FSS-PAM MNF)[110], (d) 170-nm-diameter CdS nanowire[330], and (e) 4.4-μm-diameter ice MNF[126]. The wavelength of the light launched from the left side in (b)–(e) is 633, 355, 473, and 500 nm, respectively. In particular, an obvious PL signal around the 550-nm wavelength of the FSS-PAM MNF is observed in (d). (f) Schematic of an MNF-coupled SNSPD for NIR wavelengths[332]. (g) Optical microscope image of an SU8 capped tapered fiber placed on the fork silicon-nitride-waveguide (SiN WG) coupler for low-loss, high-bandwidth fiber-to-chip coupling[337]. (h) Optical microscope image of a fiber-nanowire-silicon-waveguide cascade structure for efficient fiber-to-chip coupling[338]. The operation wavelength ranges from 1520 to 1640 nm.

PI_3_1_R02_f016.png

It is worth mentioning that MNF-assisted near-field optical coupling also works at low temperatures. For reference, Fig. 16(e) shows the coupling of a blue light (500 nm in wavelength) from a silica MNF to an ice MNF at 70°C[126]. The evident output from the right end of the ice MNF confirms the efficient launching of the waveguiding mode of the ice MNF. The low-temperature MNF-based coupling has also been used in superconducting nanowire single-photon detectors (SNSPDs). In 2017, You et al. reported a cascaded “standard fiber-MNF-superconducting nanowire” coupling structure[331,332], as shown in Fig. 16(f). Compared with the conventional planar-waveguide-based coupling structure, the MNF-based structure could offer higher coupling efficiency (up to 90%) between the standard fiber and the NbN superconducting nanowire, enabling high-performance fiber-compatible SNSPDs. Specifically, at a temperature of 2.2 K, the overall detection efficiency reached 50% and 20% at the input light wavelengths of 1064 and 1550 nm, respectively. Later in 2019, by further optimizing the coupling condition, Hou et al. improved the detection efficiency to 66% and 45% at the wavelengths of 785 and 1550 nm, respectively, offering the possibility for ultra-wideband weak-light detection with quantum-limit sensitivity[333].

Another notable example is MNF-assisted optical coupling between standard single-mode fibers and on-chip planar waveguides, which is one of the main challenges in silicon photonics[339]. In 2007, Zhang et al. reported end-fire coupling from a subwavelength-diameter silica MNF to a silicon waveguide, with coupling efficiency higher than 40% over a wide wavelength range of 1300–1700 nm[334]. Later in 2011, Shen et al. proposed optical coupling of a silica MNF and tapered silicon waveguide, showing improved coupling efficiency higher than 80% in the wavelength range of 1300–1700 nm[335]. Similarly, in 2017 Chen et al. reported vertical near-field coupling of a 976-nm-wavelength CW light from an Er3+/Yb3+-co-doped tellurite glass MNF to a silicon racetrack resonator, and realized waveguided luminescence at the telecommunication band[336]. In 2020, Khan et al. employed a capped, terminating, adiabatic tapered fiber to couple with a fork silicon nitride (SiN) waveguide [see Fig. 16(g)], demonstrating a coupling loss as low as 1.4 dB around 1550-nm wavelength and a 3-dB bandwidth of 90 nm[337]. More recently, Jin et al. designed a cascaded coupling structure of a “fiber-nanowire-silicon waveguide” for effective-index matching in each coupling area [see Fig. 16(h)], enabling bidirectional coupling efficiency as high as 90% and a 3-dB bandwidth in excess of 100 nm (1520 to 1640 nm in wavelength) for both TE and TM polarizations[338]. In addition, in some cases where sharp fiber tapers are required (e.g., for ultra-compact optical coupling), one can consider a scheme proposed by Wu et al. recently[340]. A short-length fiber taper (150 µm in length) with a nonlinearly shaped profile was fabricated on the top of a cleaved SMF tip using the direct laser writing method, allowing a relatively high optical transmittance (77%) at 1550-nm wavelength.

5.1.2.

2D materials

In recent years, owing to their broadband optical response, fast relaxation, high nonlinearity, and controllable optoelectronic properties, 2D materials have shown great potential for nanophotonics[341,342]. However, due to their atomically thin structures, light absorption is typically weak, and enhancing light-matter interaction is an essential step towards high-efficiency photonic applications in many situations. The tightly confined high-fraction evanescent fields in optical MNFs open a route for the enhancement of interaction between waveguiding modes and 2D materials. Generally, two kinds of methods including optical deposition[343] and micro/nanomanipulation-enabled dry/wet transfer[344,345], can be used to realize the integration of optical MNFs and 2D materials. To date, a variety of 2D-material-integrated MNF photonic devices have been demonstrated in a wide range of applications such as all-optical signal processing[345348], nonlinear optics[349], mode-locked fiber lasers[343,348,350], optical sensing[351], and quantum optics[352]. For example, Li et al. realized the near-field coupling of graphene films and silica MNFs (1  μm in diameter) in 2014 [see Fig. 17(a)], and demonstrated an all-optical modulator with a response time of 2.2 ps and modulation depth of 38%. Such a graphene-clad-MNF structure can also serve as a saturable absorber to realize ultrafast mode-locked pulse lasers, with advantages of low saturation intensities, ultrafast recovery times, and wide wavelength ranges[350]. In 2019, Chen et al. reported a tungsten disulfide (WS2)-clad-MNF structure and realized the enhancement of PL and SHG through the evanescent-field coupling[353]. In 2020, Jiang et al. integrated silica MNFs with few-layer gallium selenide (GaSe) nanoflakes and achieved high-efficiency second-order nonlinear processes including the SHG and SFG [see Fig. 17(b)][349]. In 2022, Yap et al. deposited MoS2 nanosheets onto an optical MNF and realized a volatile-organic-compound sensor at room temperature [see Fig. 17(c)], with high sensitivity of 0.0195, 0.0143, 0.0072, and 0.0058 nm/ppm to acetone, ethyl acetate, cyclohexane, and isopropyl alcohol, respectively[351]. Figure 17(d) shows the schematic diagram of coupling quantum emitters by use of a hexagonal boron nitride (hBN)-integrated MNF[352]. Excited by 532-nm-wavelength pulses or CW lasers, the emitters at a wavelength of 666 nm from the hBN were coupled into and waveguided along the MNF, with a coupling efficiency of 10%. This MNF-based coupling scheme provides convenience for efficiently exciting and collecting quantum emitters. More recently, Xiao et al. reported a miniature waveguide photoactuator by embedding an optical MNF in a PDMS/Au nanorod-graphene oxide photothermal film[354]. With a 635-nm-wavelength light coupled into the optical MNF, the photothermal-effect-induced temperature rise led to a significant bending of the photoactuator, with large bending angles (>270°), fast response (1.8 s for 180° bending), and low energy consumption (<0.55  mW/°)[354]. Using the MNF-2D-material-integrated photoactuators, they demonstrated soft grippers for capturing, moving, and releasing small objects with different shapes.

Fig. 17

Near-field optical coupling with 2D materials using optical MNFs. Schematic diagrams of silica MNFs coated with (a) thin layer of graphene[345], (b) few-layer GaSe[349], (c) MoS2 nanosheets[351], and (d) hBN flakes[352].

PI_3_1_R02_f017.png

5.1.3.

Optical micro-cavities

Optical micro-cavities, taking forms of microscale FP structures, dielectric spheres, disks, cylinders, rings, and other complicated structures, are basic functional structures of photonic applications ranging from nonlinear optics, optical sensing, and microlasers to cavity quantum electrodynamics and optomechanics. Coupling of light into and out of the micro-cavities is the initial step for their characterization and application. Historically, optical MNFs (known as fiber tapers) have long been used for in/out coupling of the micro-cavities, especially for whispering gallery mode (WGM) cavities (in the form of microcylinders, microspheres, microdisks, microbottles, or microbubbles), which have small mode volumes, high Q factors, and small footprints. Compared with other typical coupling methods including free-field coupling, prism coupling, planar-waveguide coupling, and angle-polished-fiber coupling, the MNF-coupling method has advantages including the highest coupling efficiency, low insertion loss, small footprint, and fiber compatibility. For example, in 1997 Knight et al. demonstrated a phase-matched excitation of WGM resonances in a microsphere[101]. By adjusting the diameter of a silica MNF to match the propagation constant of the waveguiding modes with that of the WGMs, they achieved coupling efficiency up to 90%. In 2003, Spillane et al. demonstrated nearly lossless coupling between silica MNFs and silica microspheres[355] with coupling efficiency higher than 99.97%, showing a Q factor in excess of 108. The ultrahigh-Q factor makes microspheres favorable for a huge range of applications including optical sensing[356], optomechanics[357], nonlinear optics[358], and microlasers[359]. As a case in point, Fig. 18(a) schematically shows a thulium-erbium-ytterbium (Tm-Er-Yb) co-doped silica microsphere cavity coupled with a silica MNF[359]. Pumped by a 975-nm-wavelength CW light at room temperature, an upconversion white light (i.e., three primary RGB lights) can be extracted from the output end of the MNF. In 2005, Dong et al. investigated the coupling between a 2-μm-diameter tapered fiber and a microcylinder resonator (i.e., SMF-28 fiber after thermal treatment) [see Fig. 18(b)], demonstrating a high Q factor up to 1.4×107 of the resonator[360]. In 2016, Lu et al. employed a tapered-fiber-coupled silicon microdisk (4.68×105 in Q factor at 1534.4-nm wavelength) [see Fig. 18(c)] to generate a single-photon source from spontaneous FWM in the microdisk[361]. In 2018, Gorajoobi et al. proposed the coupling of a silica MNF with a Yb3+-doped microbottle resonator [see Fig. 18(d)] to excite a low-threshold, high-efficiency, tailorable microbottle laser[362]. Despite the versatility and flexibility of the MNF-micro-cavity coupling system, the robustness of the system is relatively low since optical MNFs are susceptible to environmental perturbations (e.g., airflow disturbance and mechanical vibrations). For practical use, an effective package for the coupling system is highly desirable, such as vacuum treatment and surface protection. Alternatively, Farnesi et al. proposed a robust coupling structure based on a thick MNF (15-18 µm in diameter) with a pair of long-period fiber gratings written in the standard fiber pigtails, and demonstrated total coupling efficiency up to 60% with microspheres or microbubbles[363,364].

Fig. 18

Optical MNF as an invaluable tool for evanescent-wave coupling with micro-cavities. Schematic diagrams of the optical MNF-coupled (a) rare-earth-doped microsphere[359], (b) silica microcylinder[360], (c) silicon microdisk[361], and (d) rare-earth-doped microbottle resonators[362].

PI_3_1_R02_f018.png

5.2.

MNF-Based Passive Optical Components

Benefitting from their superior mechanical behavior and pliability, the freestanding optical MNFs with large available lengths can be handled for high-precision, flexible micro/nanomanipulation, such as fine tailoring, knotting, splicing, positioning, transferring, and assembly. To facilitate the manipulation, optical MNFs are usually placed on a clean surface of a certain substrate (e.g., a silicon, sapphire, MgF2 wafer, or a glass slide). Electrochemically sharpened tungsten probes with tip sizes of tens to hundreds of nanometers are usually mounted on high-precision moving stages to perform the micro/nanomanipulation under an optical microscopy[34,110,135,243,321]. Some basic manipulations of the optical MNFs have been reported in 2005[321], showing that an optical MNF can be pushed, cut, bent, twisted, picked up, transferred, and positioned on a substrate. Indeed, sharp fiber taper probes with relatively high stiffness can also be used as an effective micro/nanomanipulation tool, offering convenience for high-efficiency direct coupling of the input light from standard fibers to optical MNFs during the manipulation[365]. To bestow the as-fabricated MNFs with more functionalities, optical MNFs can be assembled into a variety of geometries, with typical passive components introduced below.

5.2.1.

MNF-based resonators

Owing to the efficient evanescent coupling between touched MNFs, an MNF-assembled resonator can be fabricated by simply folding or knotting an MNF into a knot, loop, ring, or coil geometry. Depending on the ring size and geometry, the Q factor of an MNF resonator typically ranges from 102 to 106.

The MNF loop resonator, usually formed by twisting an MNF, is the simplest structure among MNF resonators. In 2004, by bending and coiling free silica MNFs into a self-touching microloop structure, Sumetsky et al. developed a self-coupling microloop optical interferometer[152]. Shortly after, by improving the self-coupling condition, the same group successfully demonstrated a loop resonator assembled by a 660-nm-diameter MNF [see Fig. 19(a)], showing a Q factor higher than 1.5×104 and a finesse of 10 around 1.5-μm wavelength[366]. In 2006, Sumetsky et al. achieved an intrinsic Q factor of 6.3×105 by improving the coupling efficiency and reducing the optical loss of the MNF[367]. To improve the robustness, in 2007 Guo et al. demonstrated a copper-rod-supported loop resonator by wrapping a 2.8-μm-diameter MNF around a 460-μm-diameter copper rod[368]. In addition, an MNF can also be assembled into a Sagnac loop[369], based on which an all-fiber FP resonator can be obtained.

Fig. 19

MNF-based photonic components. Optical microscopic images of MNF-based passive optical components including (a) loop[366], (b) knot[54], and (c) ring[378] resonators. (d) Optical microscopic image of an MZI assembled with two 1-μm-diameter silica MNFs[395]. (e) SEM image of a Bragg grating inscribed on a 1.8-μm-diameter silica MNF[401]. (f) SEM image of a plasmonic-photonic cavity with several Au nanorods deposited on a 2.2-μm-diameter silica MNF[413].

PI_3_1_R02_f019.png

In contrast to the self-touching loop structure that is maintained by the van der Waals and electrostatic forces, a knot structure has much higher robustness, especially operating in liquids[370374]. In 2006, Jiang et al. assembled a free-standing MNF into a knot and demonstrated its high stability in a liquid environment, exhibiting a Q factor as high as 3.1×104 and a finesse of 13 around 1570-nm wavelength[370]. Such an MNF knot has only one input/output end connected with a standard fiber, while the other end typically relies on evanescent coupling to the output/input port. To simplify the input/output coupling, in 2011 Xiao and Birks developed a “knot-stretch” approach to assembling MNF knot resonators connected to standard fibers at both sides[371], which also increased the overall robustness of the knot structure [see Fig. 19(b)]. Using high-precision translation stages, they obtained a knot resonator (1 µm in MNF diameter and 570 µm in knot diameter) with a Q factor of 9.7×104 and a finesse of 73 around 1550-nm wavelength. In 2017, by attaching an MNF knot (5 µm in MNF diameter and 1 mm in knot diameter) on a 100-nm-thick gold film, Li et al. reported a hybrid plasmonic MNF knot resonator with a Q factor higher than 5.2×104 around 1550-nm wavelength[373]. Later in 2020, the same group improved the Q factor of this hybrid plasmonic MNF knot resonator (2.4 µm in MNF diameter and 1.1 mm in knot diameter) to 7.9×104[374].

Furthermore, an MNF can be fabricated into a free-standing closed-loop ring resonator through fusion splicing, which possesses a higher mechanical robustness than the above-mentioned MNF-based resonators. For example, in 2008 Pal and Knoxa used a CO2 laser beam to splice MNFs, showing a splicing loss below 0.3%[375]. Shortly after, by fusion-splicing the coupling region of an MNF loop, they obtained a high-stability loop resonator with a Q factor of 2.5×104 around 1550-nm wavelength[376]. In 2009, by fusion-splicing two ends of an MNF (3.8 µm in diameter), Wang et al. successfully obtained a 2.4-mm-diameter closed-loop resonator, with a Q factor of 1.1×105 and a finesse of 15.3 around 1.5-μm wavelength[377]. In 2011, by fusion-splicing a phosphate glass MNF, Li et al. fabricated a free-standing closed-loop 1.34-mm-diameter ring resonator [see Fig. 19(c), green part], with typical splice losses of 0.2  dB and a Q factor of 2.5×105 around 1.6-μm wavelength[378].

To obtain a high Q-factor 3D microcavity, in 2004 Sumetsky theoretically proposed a stacked MNF coil resonator[379]. Different from the WGMs in loop/knot resonators, the light confinement in such an MNF coil resonator is achieved by self-coupling between turns rather than by the presence of a closed optical path. In 2007, by wrapping a 1.7-μm-diameter MNF around a 1-mm-diameter rod, Sumetsky et al. experimentally demonstrated a two-turn MNF coil resonator, showing a Q factor as high as 6.1×104 around 1530-nm wavelength[380]. In the same year, Xu and Brambilla fabricated an MNF coil resonator by wrapping a 1.5-μm-diameter MNF around a 560-μm-diameter rod in two to four turns, with a Q factor of about 1×104 around 1530-nm wavelength[381]. To optimize the resonators with higher Q factors, several options of geometry modification have been investigated[382,383]. For example, in 2010 Jung et al. reported a uniform cylindrical MNF coil resonator with an improved Q factor of 2.2×105 around 1550-nm wavelength[384].

So far, benefitting from their advantages of high Q-factors, tunable resonance, high robustness, and fiber compatibility, MNF-assembled resonators have been explored for applications from optical sensors[385389], filters[390], and lasers[391393] to nonlinear optics[72] and atom optics[394].

5.2.2.

MNF-based MZIs

A Mach-Zehnder interferometer (MZI), which typically has an isolated reference arm and a sensing arm, is one of the most common structures used in optical sensors, modulators, and filters. Benefitting from their high-efficiency near-field coupling, MNFs can be assembled into MZIs with small footprints and high flexibility. In 2008, Li and Tong experimentally assembled silica and tellurite glass MNFs into highly compact MZIs [see Fig. 19(d)], with footprints of tens to hundreds of micrometers and extinction ratios of 10  dB[395]. In 2012, Wo et al. demonstrated a simple and robust MNF-based-MZI structure assembled by a 2-μm-diameter silica MNF for sensing applications[396]. To enhance mechanical stability, MNF-based MZIs can also be embedded into low-index polymer, with a slight degradation in extinction ratio[251]. It is worth mentioning that MNF-based MZIs can also be assembled by the MNFs made of different materials[397]. For example, in 2013, by coupling a Ag nanowire with a silica MNF, Li et al. reported a hybrid photon-plasmon MNF-based MZI, exhibiting a Q factor of 6×106 and an extinction ratio up to 30 dB around 1550-nm wavelength[325].

5.2.3.

MNF Bragg gratings

Similar to the Bragg gratings inscribed on standard fiber, they can also be fabricated on an MNF, making it an MNF Bragg grating (MNFBG) with a much smaller size. So far, MNFBGs have been fabricated by femtosecond laser pulse/CW irradiation[398,399], focused ion beam (FIB) milling[400402], etching commercial FBG[403,404] or UV irradiated FBG[403,405]. For example, in 2005 Liang et al. reported a chemically etch-eroded MNFBG with a diameter of 6 µm and demonstrated its application in sensing refractive indices of different liquids[403]. In 2011, Liu et al. obtained a 518-μm-long and 1.8-μm-diameter MNFBG by the FIB milling [see Fig. 19(e)][401], with a reflection peak/transmission dip located at 1538-nm wavelength. When being used for optical sensing, such an MNFBG exhibited a sensitivity as high as 660 nm per refractive index unit.

Meanwhile, many other types of gratings, including long-period gratings[179,406,407], Type IIa Bragg gratings[408], chirped Bragg gratings[409], and MNFBGs arrays[404], have also been demonstrated and applied to high-sensitivity optical sensing. Additionally, it is worth noting that, benefitting from the efficient near-field coupling, microscale gratings can also be formed by simply attaching and coupling an MNF to an external grating structure[410], which offers additional flexibility to MNF-based gratings.

5.2.4.

MNF-based plasmonic-photonic cavity

By strongly coupling plasmonic modes of metal nanostructures (e.g., metal NPs[411,412]) with WGMs of an optical MNF, a new type of hybrid plasmonic-photonic resonator can be realized. For example, in 2015 Wang et al. constructed a plasmonic-photonic cavity by depositing single Au nanorods on the surface of a silica MNF [see Fig. 19(f)][413]. Owing to the strong coupling between the LSPR modes of the Au nanorod and the WGMs of an MNF, a significant reduction in the LSPR spectral width (from 50 to 2 nm) of the Au nanorod was observed. Such a strong-coupled hybrid plasmonic-photonic resonating scheme has opened a variety of opportunities from improving spatial resolution or sensitivity in optical sensing[414,415] and miniaturizing strong-coupling systems[416] to enabling single-nanorod-based photon-plasmon lasing[417] and enhancing nonlinear optical effects[418420].

5.3.

Optical Sensors

Owing to their favorable waveguiding properties, especially tightly confined high-fraction evanescent fields, optical MNF is one of the most promising choices for optical sensing on the micro/nanoscale. Typically, the changes of surrounding media or samples will change the waveguided light in an MNF via scattering, absorption, dispersion, emission, or other processes, and change the transmitted light in intensity, phase, polarization, or spectral features, which can be used for retrieving the information of the measurands. Compared with other fiber-optic sensing approaches[421425], MNF-based sensors have the advantages of high sensitivity, small footprints, and fast responses. So far, various MNF-based structures, including straight MNF, MNF-assembled structures, surface functionalized MNF structures, and polymer-embedded MNF structures, have been employed for optical sensing. Over the past few decades, there have been numerous review articles focusing on MNF-based optical sensors or similar devices[6466,387,426437]. In this section, we will briefly summarize typical MNF-based sensors and provide an update on the recent advances in this field, categorized with different structures.

5.3.1.

As-fabricated straight MNFs

As-fabricated MNFs, usually biconically connected to standard optical fibers, are the simplest optical sensing structures based on surface absorption or scattering. For instance, in 2007 Warken et al. proposed a molecule sensing scheme based on the absorption spectrum of the waveguided light in a 500-nm-diameter silica MNF, and demonstrated the detection of sub-monolayers of 3,4,9,10-perylene-tetracarboxylic dianhydride (PTCDA) molecules[438]. Later in 2014, Yu et al. achieved the single-NP detection and sizing in an aqueous environment using a pair of 500-nm-diameter MNFs[439]. As shown in Fig. 20(a), when an NP is attached to the MNF surface, a downward step in the transmission occurres [see Fig. 20(a), lower panel]. Such a sensing structure can also be extended to MNF arrays, enabling faster and more efficient detection[440].

Fig. 20

Typical MNF-based photonic sensors. (a) Schematic illustration of a single NP detection system, where a pair of identical MNFs is used (upper panel)[439]. A diode laser with a wavelength of 680 nm is employed as the probe light. The transmitted light is finally detected by a 125-MHz photodetector and monitored by an oscilloscope. Typical optical transmission of an MNF during a time interval of 10 s when the PS NPs are binding to the surface of the MNF one-by-one (lower panel). Each data point is the average of 250 values of measured transmitted power during 20 ms, and the red curve is for guiding the eyes. (b) Schematic illustration of a microchannel-supported polymer-MNF-based gas sensor[112]. A laser with a wavelength of 532 nm is coupled into a 250-nm-diameter PANI/PS MNF with fiber tapers. The time-dependent absorbance of the sensor to cyclic NO2/nitrogen exposure with NO2 concentration from 0.1 to 4 ppm is shown in the lower panel. Inset: dependence of the absorbance over the NO2 concentration ranging from 0.1 to 4 ppm. (c) Schematic illustration of a gelatin layer coated silica MNF for relative humidity sensing[499]. The transmitted light intensity of the sensor at 1550 nm wavelength in the range of 9%–94% relative humidity, and the typical time-dependent transmittance of the sensor when relative humidity jumps from 75% to 88% are shown in the lower panel. (d) Plasmonic-nanostructure-activated MNF biosensor[510]. Images in the lower panel demonstrate that the sensor can not only detect cancer cells, but also treat cells through cellular photothermal therapy. (e) Schematic illustration of a skin-like wearable MNF-based sensor[247], constituting of an 80-μm-thickness PDMS film, a 980-nm-diameter MNF, and a glass slide. The response to the pressure of 2.1, 1.3, 0.2, and 0.1 Pa, and the temporary response to forced oscillation frequencies of 1, 4, and 20 kHz are shown in the lower panel. (f) Optical detection of cardiovascular vital signs (upper panel) based on the PDMS-packaged-MNF pulse-wave signal sensing principle shown in the lower panel[517].

PI_3_1_R02_f020.png

For biochemical sensing, the MNFs are usually immersed in a liquid, in which the effective refractive index of the MNFs should be larger than the refractive index of the liquid. In 2011, by integrating a 900-nm-diameter silica MNF into the microfluidic chip, Zhang et al. demonstrated efficient sensing of bovine serum albumin using a 633-nm-wavelength probe light[244], with a detection limit of 10 fg/mL and a probe light power down to 150 nW. Meanwhile, the selective detection of target analytes in biochemical samples can be achieved by immobilizing specific signal-responsive receptors[441443].

The intermodal interference in a multimode MNF has also been explored for optical sensing[444,445]. Commonly, a large proportion of the light energy from the fundamental mode of the untapered region will couple into the HE11 and HE12 modes in the tapered region, resulting in inter-mode interference. The feature of such interference is highly sensitive to the change of the surrounding environment, and is well-suited for sensing applications. In 2006, Kieu et al. reported a displacement sensor assembled by an 8-μm-diameter MNF, showing an accuracy of 100 nm[444]. A refractive-index sensor capable of measuring Δn (1.42×105) and a temperature monitor with sensitivity ΔT (1°C) were also demonstrated. Recently, a number of MNF sensors based on multimode interference have been reported for measuring the refractive index[445448], temperature[449], strain[449,450], and magnetic fields[451].

Compared to glass MNFs, polymer MNFs have special advantages for optical sensing, including low cost, great flexibility, infrared (long-wavelength) transparency, excellent biocompatibility, permselective feature to gas molecules, and hospitality for a variety of dopants. For instance, in 2008, relying on the spectral response of a NO2 concentration-dependent oxidation degree of polyaniline (PANI) mixed in a 250-nm-diameter PS MNF, Gu et al. developed a NO2 sensor with a low detection limit (<0.1  ppm) and a response time of 7  s [see Fig. 20(b)][112]. In 2010, on the basis of the surface passivation of QD emission in a 480-nm-diameter CdSe/ZnS QD-doped PS MNF, Meng et al. realized a miniaturized optical humidity sensor with a response as fast as 90 ms and an ultra-low optical power of about 100 pW[111]. Most recently, in 2023 Yang et al. reported eco-friendly polymer MNFs from natural lotus silks, and found their applications in sensing pH value and bacterial activity[116].

5.3.2.

MNF-assembled structures

In principle, most of the MNF-assembled passive optical structures mentioned in Sections 5.1 and 5.2 can be employed for optical sensing, as introduced below.

For an MNF-based coupling structure, the coupling efficiency is highly dependent on the refractive indices of the MNF and the surrounding medium, as well as the coupling length, which is available for optical sensing. For example, in 2015 Luo et al. proposed a compact magnetic-field sensor using an MNF coupler enclosed in a magnetic fluid[452]. The magnetic-field-induced change in the refractive index of the magnetic fluid changed the coupling efficiency of the MNF coupler, resulting in a maximum sensitivity of 191.8 pm/Oe in wavelength shift. In 2016, based on two 1.4-μm-diameter coupled MNFs, Li et al. achieved a refractive index sensor working near the turning point of the effective group index difference between the even supermode and odd supermode, and obtained a sensitivity as high as 39541.7 nm/RIU with an ambient refractive index of 1.3334[453]. In addition, the loss of the MNF coupler-based sensor can be reduced by introducing a Sagnac loop, which is constructed by connecting two standard optical fibers on one side of the coupler and acts as a reflector[454,455].

The MNF-based MZI with phase-sensitive detection is another widely used highly sensitive structure. In 2005, Lou et al. theoretically predicted that an MZI assembled with two MNFs could offer a sensitivity one order of magnitude higher than those of conventional waveguide MZIs[456], owing to the accessible high-fraction evanescent waves in the MNFs. To date, a number of MNF-based MZIs have been utilized for phase-sensitive optical sensing. For example, in 2012, based on an MZI comprised of two 2-μm-diameter silica MNFs, Jasim et al. detected the current flowing in a copper wire with a sensitivity of 60.17 pm/A2[457]. In 2015, Luo et al. achieved an ultrahigh refractive index sensitivity of 10777.8 nm/RIU near the dispersion turning point of a multimode MNF-based MZI[458].

Since the resonance spectrum of an MNF resonator is highly sensitive to the change of resonating structure or environmental conditions (e.g., refractive index, temperature, and strain), the MNF resonators have also been widely studied for optical sensing. In 2006, relying on a high-Q-factor MNF loop resonator (with an intrinsic Q factor of 6.3×105), Sumetsky et al. realized the temperature detection with a resolution of 0.1 mK and a response time on the order of microseconds[367]. Compared with the loop structure, an MNF knot possesses higher robustness. In 2010, Wu et al. reported an accelerometer based on a 386-μm-diameter knot resonator assembled with a 1.1-μm-diameter MNF in a micro-electromechanical system (MEMS), exhibiting a sensitivity of 624.7 mV/g and a dynamic range of ±20  g[459]. Meanwhile, the stacked 3D MNF coil resonator has also been explored for measuring the refractive index[389], current[460462], acoustics[463], and absorption[464,465]. For example, in 2010 benefitting from the Faraday rotation, Belal et al. presented a 25-turn MNF-coil current sensor with a responsivity of 16.8±0.1  μrad/A, which could also be used to sense high-frequency currents or magnetic fields (e.g., 2 GHz in principle)[460].

Up to now, a variety of MNF-assembled sensing structures have been reported for sensing force[454,466], strain[467470], temperature[471477], humidity[385387,478], refractive index[458,479483], electric current[484], magnetic field[485,486], biochemical compositions[487], and gases[325]. More details can be found elsewhere[426,430,431,434,437].

5.3.3.

Surface functionalized MNF structures

Owing to its strong surface waves (i.e., waveguided evanescent fields), an MNF can be readily functionalized by modifying either the surface structure or the dielectric environment near the surface. Here, we introduce three typical types of surface-functionalized MNF sensors: MNFBG sensors, functionalized-coating MNF sensors, and plasmonic-nanostructure-activated sensors.

As mentioned in Section 5.2.3, an MNFBG can significantly enhance the sample-light interaction using abundant evanescent fields, and has been investigated for sensing the refractive index[398,401,488,489], temperature[400,490], strain/force[490,491], gas[492], biochemical compositions[493495], and acoustic waves[496]. For example, in 2010 Fang et al. demonstrated a refractive sensor based on a 2-μm-diameter MNFBG[398]. Typically, an MNFBG with a thinner diameter and higher-order mode resonance exhibited larger refractive sensitivity, and a maximum sensitivity of 231.4 nm/RIU at a refractive index of 1.44 was achieved. In 2022, by integrating an 800-nm-diameter ZnO MNFBG on the tip of an optical fiber taper, Li et al. demonstrated a compact label-free nanosensor for real-time in-situ early monitoring of cellular apoptosis in individual living cells[495]. In 2023, Song et al. developed a near-infrared MNFBG operated at 785-nm wavelength[497], and employed it to monitor the axial tension and the bending with a responsivity of 211 nm/N and 0.18 nm/deg, respectively. Additionally, other designs of MNF-based gratings such as long-period gratings[179,406,407], chirped Bragg gratings[409], and MNFBGs arrays[404] can also be employed for high-sensitivity optical sensing of measurands from the refractive index and force to acoustic waves.

Similar to inscribing gratings on the MNF surface, coating the MNF surface is another efficient approach to functionalizing an MNF for optical sensing with high compactness and sensitivity, and can have a simpler fabrication process and higher selectivity. In 2005, Villatoro et al. developed a hydrogen sensor by coating a 4-nm-thick palladium film on the surface of a 1.3-μm-diameter MNF[498]. According to the hydrogen-concentration-dependent attenuation at 1550-nm wavelength, they obtained a detection limit down to 0.05% and a response time of about 10 s. In 2008, relying on a gelatin-coated (80 nm in thickness) 680-nm-diameter MNF, Zhang et al. developed a humidity sensor with a fast response (70 ms) within a wide humidity range (9%–94% relative humidity) [see Fig. 20(c)][499]. Additionally, the usage of graphene atomic/graphene oxide layers for the detection of adsorbed gas molecules has been attracting much attention due to its high sensitivity and low detection limit[492,500505]. In 2021, Huang et al. demonstrated a multilayer-nanoparticle-modified graphene oxide-coated MNF for sensing ethanol with a low detection limit (5.25 ppm) and a fast response (118 ms)[506]. In addition, many other functional films, such as black phosphorus[507], WS2 layer[508], and ternary cross-linked film (PVA-APTES-ICA)[509], have been coated on MNFs for optical sensing.

Besides the above-mentioned surface functionalization, plasmonic nanostructures have also been deposited or integrated on the surface of an MNF, such as nanorods[414], NPs[411,412], and nanohybrids[510], for locally enhancing the MNF-based light-matter interaction, which offers an opportunity for hybrid photon-plasmon sensing with enhanced performances. For example, in 2015 Gu et al. integrated Pd NPs on an MNF surface and generated strong coupling between LSPRs and the MNF whispering-gallery modes (WGMs)[415]. Based on the ultranarrow resonances (a measured FWHM of 3.2 nm at 622.7-nm resonant wavelength), the Pd-nanoantenna-MNF cavity system can detect hydrogen gas with an enhanced sensitivity of about 5.11 nm/%VOL. Later in 2019, Zhou et al. demonstrated a gold-nanorod-MNF coupled system for relative humidity sensing, and achieved a spatial resolution of 1.5 mm[414]. Meanwhile, such photon-plasmon coupled MNF systems have also served as an effective platform for biosensing[441,511,512]. For example, in 2019 Li et al. functionalized the surface of a 7.1-μm-diameter MNF with a plasmonic nanointerface consisting of black-phosphorus-supported gold nanohybrids[510]. As shown in Fig. 20(d), this MNF-based sensor was not only able to detect an epidermal growth factor receptor (ErbB2) at concentrations from 10 zM to 100 nM at the single-molecule level (detection limit 6.72  zM), but also able to treat cancer cells through cellular photothermal therapy. Two optical images in Fig. 20(d) illustrate the comparison before and after cellular photothermal therapy, clearly showing the removal of the cancer cell.

5.3.4.

Polymer-embedded MNF structures

To isolate the MNF from surface contamination, enhance the mechanical stability, or make the MNF sensor wearable, MNF sensing structures can be embedded in polymer substrates or thin films. Typically, in such structures, the change of the refractive index or geometry of the polymer film will lead to the change of the optical transmittance of the MNF embedded inside, making it suitable for sensing a variety of measurands including tension, pressure, bending, strain, temperature, and humidity, which is highly desired in applications including health monitoring[246,513,514], human-machine interaction[515,516], and intelligent robots. For instance, in 2018, by embedding a hybrid plasmonic MNF knot resonator into a PDMS film, Li et al. developed a pressure sensor with a sensitivity of 0.83  kPa1 and a detection limit of 30 Pa, for sensing of wrist pulse, respiration, and finger pulse[470]. In 2020, based on an MNF-embedded PDMS patch, Zhang et al. obtained a skin-like wearable optical sensor with a pressure detection limit of 7 mPa and a response time of 10 µs [see Fig. 20(e)][247]. In 2022, Li et al. realized a wearable MNF-based sensor chip for precise vital signs monitoring and cardiovascular health assessment[517]. Figure 20(f) shows the physical image of the sensor chip (upper panel) and the schematic diagram of pulse wave signal sensing (lower panel).

Moreover, discrete single sensors can be integrated to implement complex functions. For instance, in 2020 Zhang et al. demonstrated a five-sensor-integrated optical data glove for monitoring the flexion and extension of the joints of fingers[247]. In 2022, Ma et al. used an MNF array (five parallel MNFs) to fabricate an optically driven wearable human-interactive textile with a high sensitivity (65.5  kPa1) and a fast response (25 ms) for touch sensing, and demonstrated a remote-control robotic hand and a smart interactive doll based on such optical smart textiles[515]. More recently, relying on soft PDMS-embedded MNFBGs attached to different body locations (e.g., chest, wrist, and neck), Zhu et al. reported a spatiotemporal hemodynamic measurement technique to monitor hemodynamic parameters (e.g., systemic pulse transit time, heart rate, blood pressure, and peripheral resistance), with high sensitivity, electromagnetic immunity, and temporal synchronization between multiple remote sensor nodes[514].

In addition, by flexible geometry design of the PDMS cladding and the embedded MNF structure, these kinds of sensors have also shown great versatility for sensing of hardness[518], flow rate[519,520], strain[246,466,521], slip[522], and acoustics[523].

5.3.5.

More possibilities

Besides the aforementioned MNF-based optical sensors, there are also many other sensing structures, such as in-line MNF-MZI[474], cascaded MNF knot resonators[524,525], and MNF-WGM coupling structure[526,527], which have been demonstrated for physical, chemical, and biological applications. Due to the limited space, we will not go into detail here.

Essentially, as a miniature fiber-optic sensor, the MNF sensor is outstanding for high-sensitivity, fast-response, electromagnetic-immunity, and high-flexibility optical sensing on the micro/nanoscale. From our point of view, the current trend of MNF sensing is pushing the detection limit, achieving multiparameter sensing, and developing intelligent wearable devices, as well as targeting future biomedical sensing applications. Rapid progress on MNFs with new functional structures and materials, as well as new mechanisms or effects for optical sensing, will continue to bestow MNF-based optical sensors with new opportunities.

5.4.

Optomechanics

5.4.1.

Optomechanics in a single MNF

It is well known that optomechanical force arises from a photonic momentum exchange between light and matter. When acting on optical MNFs, a fast, evident optomechanical response can be observed due to their small mass/weight (for a 200-nm-diameter 10-μm-long silica MNF, on the order of 1015  kg) and low stiffness (5.44 fN/nm). In 2008, She et al. investigated the optomechanical behavior of silica MNFs triggered by a 650-nm-wavelength long pulse (1/5 s in pulse interval)[89]. When the light traveled from the free end of a 450-nm-diameter MNF to the air, an inward push force on the free endface was observed, supporting Abraham’s momentum of light in a transparent dielectric. Also, the MNF responded to the pulse on and off instantly by observable deformation. Based on this optomechanical effect, an MNF-based all-optical switch was demonstrated, possessing a turnoff time of 500  ms and recovery time of 760  ms[94]. In 2015, Luo et al. presented a detailed theoretical model of the optomechanical effect in a silica MNF Bragg grating[528]. The light-induced strain along the grating introduced an optically reconfigurable chirp in the grating period and optical delay at a shorter wavelength, which is promising for all-optical switching and tunable optical delay lines. In 2019, Zheng et al. further investigated the optomechanical motion of a suspended silica MNF on the top of a glass substrate[90]. When a 1458-nm-wavelength pump light was waveguided in the MNF, a portion of light was evanescently coupled into the substrate, which induced a momentum transfer from the MNF to the substrate. The momentum change introduced a repulsive optical force on the MNF, which pushed the MNF away from the substrate. Assisted with a white-light-interference measurement method (0.356 nm in resolution), they obtained an optomechanical efficiency of the MNF-based nano-optomechanical system as 20.5 nm/μW. Subsequently, based on the nano-optomechanical system, a broadband (up to 208 nm) and low-power (down to 624.13 µW) light-control-light technique was achieved, as schematically illustrated in Fig. 21(a)[91]. With a high-power pump light launched into the MNF, the repulsive force exerted on the MNF caused a micro-bend (i.e., a gap between the MNF and the substrate), which reduced the coupling loss between the MNF and the substrate. At this time, the signal light could propagate through the MNF with a lower optical loss. By utilizing this scheme, the minimum optomechanical force exerted on the MNF was measured to be 380.8 fN.

Fig. 21

Optical MNF optomechanical systems. (a) Schematic illustration of the light-control-light process in an MNF nano-optomechanical system[91]. The evanescent-field coupling of the pump light (at a wavelength of λp) to the substrate generates a repulsive optical force to push the MNF far away from the substrate, allowing the signal light (at a wavelength of λs) to pass through with less loss. (b) Time-sequential optical microscope images of 3-μm-diameter PS particles propelled along a 950-nm-diameter MNF in the deionized water at 1-s intervals[532]. The wavelength of the input light from the left side is 1047 nm. (c) Schematic illustration and time-lapse-compilation images of light-induced rotation of 3-μm-diameter PS particles around a 700-nm-diameter MNF in the deionized water[540]. The counter-propagating lights from both sides come from the same source at 1064-nm wavelength, with a helicity parameter σ=+1. (d) Sequencing optical microscope images showing optical propelling of an oil droplet (11  μm×10  μm ellipsoid) along a silica MNF (1 µm in diameter) at an interval of 1.2 s[54]. The input CW light (0.7 W in power) is coupled and waveguided along the MNF from left to right. (e) Schematic of pulling a hexagonal gold plate (5.4 µm in side length and 30 nm in thickness) up on a tapering-profile silica MNF (6° in cone angle) near the tapered end[541]. The photophoretic pulling force originates from the light-induced thermal effect. (f) Schematic of optical selection and sorting of single nanodiamonds along an optical MNF in pure water[546]. When two different-wavelength lasers counter-propagate along the MNF, a nanodiamond can be trapped by the gradient force and transported by the absorption and scattering forces. The scattering forces can be cancelled out by choosing applicable laser power and in this case, the movement of the nanodiamond depends on the absorption cross section.

PI_3_1_R02_f021.png

5.4.2.

MNF-based optomechanical trapping and propelling of particles

When acting on the surrounding media (e.g., micro/nanoparticles), the evanescent field around subwavelength-diameter MNFs can trap micro/nanoparticles by an optical gradient force, and propel them along the length of the MNF by a scattering force. Compared with other optomechanical manipulation tools like free-space optical tweezers and on-chip waveguide tweezers, the freestanding MNF-based optomechanical system exhibits advantages of high efficiency, high precision, low loss, long propelling length, high flexibility, and fiber compatibility. To quantificationally describe the optomechanical forces exerted on the micro/nanoparticles around the MNFs, several theoretical models have been proposed[529,530]. The most frequently used model is to integrate the time-independent Maxwell stress tensor T over a closed surface S surrounding the particles, as given by[530,531]

Eq. (24)

F=ST·nSdS,
where ns is a normal vector pointing to the outward direction from the surface S, and the elements of the Maxwell stress tensor Ti,j can be expressed as

Eq. (25)

Ti,j=ε0εrEiEj+μ0μrHiHj12δi,j(ε0εr|E|2+μ0μr|H|2),
where subscripts i and j are the indices running from x, y to z in Cartesian space, ε0 and μ0 are the vacuum permittivity and permeability, εr and μr are the relative permittivity and relative permeability of the medium, E and H are the electric-field and magnetic-field vectors, Ei and Ej are the i-th and j-th components of the E vector, while Hi and Hj are the i-th and j-th components of the H vector, respectively, and δi,j is Kronecker’s delta.

Early experimental work on the optomechanical manipulation of microparticles using silica MNFs was carried out by Brambilla et al. in 2007[532]. Figure 21(b) shows consecutive images of propelling 3-μm-diameter PS particles along a 950-nm-diameter MNF in deionized water, which were captured by the CCD camera at 1-s intervals. For a 1047-nm-wavelength, 400-mW input light, the velocities of particles A and B were measured as 9.0 and 7.0 µm/s, respectively. In 2012, Xu et al. demonstrated size-dependent optical trapping and propelling of submicrometer-diameter PS particles in water by launching a 532-nm-wavelength light into a 600-nm-diameter silica MNF[533]. At the same input power, the particles with a larger diameter were more easily trapped and delivered. For example, at an input power of 10 mW, the measured average delivery velocity of 400-nm-diameter particles was 24 µm/s, while that of 700-nm-diameter particles was 63 µm/s. In the same year, Lei et al. realized bidirectional optical transportation of 713-nm-diameter PS particles by coupling 980-nm-wavelength light into the two ends of a silica MNF (with different powers)[534]. When the power of two counter-propagating lasers was adjusted to the same, the transported particles halted on the MNF surface. Furthermore, by employing two beams of counter-propagating light with different wavelengths (e.g., 808 and 1310 nm), Zhang and Li achieved continuous particle sorting in a subwavelength-diameter MNF[535]. In this regime, the PS particles in the two sizes (i.e., 600 nm and 1 µm in diameter, respectively) could be transported in opposite directions along an 800-nm-diameter MNF. For biological applications, in 2013 Xin et al. demonstrated stable optical trapping and transport of the Escherichia coli bacteria using a silica MNF placed in a microfluidic channel[536].

Note that the above optomechanical manipulation is based on the fundamental mode in a silica MNF. Considering the larger extension of evanescent fields in high-order modes, Maimaiti et al. realized optically propelling particles by exploiting the evanescent fields of high-order modes in an MNF. Compared with the case of fundamental-mode propelling, the evanescent field in high-order modes provided a larger optomechanical force, allowing for a higher transportation velocity of the microparticles[537,538]. For reference, at the same power of 25 mW, the velocity of a 3-μm-diameter particle driven by a quasi-LP11 mode (i.e., 72.5 µm/s) was eight times faster than that driven by a quasi-LP01 mode (i.e., 8.5 µm/s)[537].

Besides the linear propelling trace, the microparticle can also be optomechanically rotated around the MNF. In 2019, following the early theoretical work of Le Kien and Rauschenbeutel[539], Tkachenko et al. experimentally demonstrated optically rotating a 3-μm-diameter PS particle around a 660-nm-diameter silica MNF when waveguiding elliptically polarized fundamental modes [see Fig. 21(c)][540]. The orbit behavior can be attributed to the azimuthal optical force, and the orbit frequency of the particle is proportional to the helicity parameter σ of the waveguided light. This finding provides a new degree of freedom for optomechanical manipulation.

It is worth pointing out that the above-mentioned cases of optomechanical manipulation of micro/nanoparticles are carried out in a liquid environment to alleviate inevitable factors such as surface adhesion force and gravity. As has been mentioned in Section 3.3.5, with much higher waveguided power and thereby much larger optomechanical force in a silica MNF, higher-speed optical propelling of microparticles can be achieved in air or vacuum. For example, Fig. 21(d) shows time-sequential optical microscope images of optically propelling a silicone oil droplet along a 1-μm-diameter silica MNF in the air[54]. Driven by a 0.7-W-power 1.55-μm-wavelength CW light, the droplet moved along the MNF with a velocity of 158 µm/s. When the waveguided power of the MNF was increased to 2.2 W, the droplet velocity could be increased to 2.1 mm/s.

5.4.3.

Photophoretic force

Besides the optical gradient/scattering force, the photophoretic force (originated from the optical field-induced thermal effect) also provides an alternative approach to manipulating light-absorbing objects in air (usually shown as a pulling effect), which has intrigued great research interest in recent years[541545]. In 2017, by tailoring the optical scattering force and photophoretic force, Lu et al. realized the optical pulling and pushing of a hexagonal gold plate up on a tapering-profile silica MNF in the air[541]. As shown in Fig. 21(e), the micrometer-sized gold plate could be pulled against the direction of light propagation from the taper end where the photophoretic force dominated. Then the gold plate was pushed back by an optical-scattering-dominated force after it was pulled to the middle section of the taper. Later, more optomechanical systems for gold-plate manipulation have been demonstrated[542544]. In addition, relying on the synergistic working of heat-induced expansion, friction, and contraction, Linghu et al. demonstrated a continuous and controllable wriggle of single gold nanowires along silica MNFs in air, showing advantages of sub-nanometer positioning accuracy, low actuation power, and self-parallel parking[545]. For a single light-absorbing NP (e.g., nanodiamond), the photophoretic force can be used to balance the optical scattering force, facilitating optical selection and sorting of individual NPs [see Fig. 21(f)]. A concrete scheme was demonstrated by Fujiwara et al. in 2021[546], which is also applicable to the high-precision optical sorting of nanocrystals, quantum dots, and molecular NPs based on their resonant absorption properties.

5.5.

Fiber Lasers

Miniaturization of fiber lasers is always of significant interest for smaller footprint and shorter-cavity applications. As mentioned previously, owing to the tightly confined waveguiding mode and small allowable bending radius, an MNF can be assembled into a micro-ring resonator with an overall size of less than 1 mm, offering an opportunity to develop a micro-ring laser with a sub-mm size. Moreover, the large diameter-dependent waveguide dispersion and very low insertion loss with standard fibers make the MNF ideal for low-loss short-length dispersion management in mode-locked fiber lasers. This section will focus on these two types of MNF-assisted lasers.

5.5.1.

MNF micro-ring lasers

As has been mentioned in Section 2.2, an active MNF can be directly drawn from a bulk glass doped with rare earth ions. By assembling such an MNF into an active cavity (e.g., knot, ring, or loop), a highly compact MNF laser can be realized. In 2006, based on theoretical calculations, Li et al. proposed a compact laser configuration relying on the resonance of the pump and signal light in a rare-earth-doped MNF ring resonator, and anticipated that the size of an Yb3+-doped MNF ring laser could go down to 50 µm[547]. Shortly after, Jiang et al. experimentally demonstrated a 1.5-μm-wavelength micro-ring laser using an Er3+/Yb3+-doped phosphate-glass MNF knot [see Fig. 22(a)][391], with a ring size of 2 mm, a lasing threshold of 5  mW, and output power higher than 8 µW. In 2007, by immersing an MNF knot into a rhodamine 6G dye solution, Jiang et al. demonstrated an MNF knot laser with a size of 350 µm[392]. In 2012, by assembling an MNF drawn from an Er3+/Yb3+ co-doped phosphate glass fiber into a double-knot resonator, Fan et al. successfully realized a single-frequency MNF laser around 1536-nm wavelength, with a linewidth as narrow as 2 kHz[417]. Besides the above-mentioned configurations, in recent years, many other types of micro-ring lasers have been reported based on rare-earth-doped MNFs[378,548550]. It is worth mentioning that much smaller lasing cavities can be realized using WGMs of active MNFs, usually dye-doped polymer MNFs. For example, in 2013 Ta et al. demonstrated single-mode and multi-mode WGM lasing at room temperature under optical pumping in a 32-μm-diameter dye-doped PMMA MNF, with a linewidth of lasing mode narrower than 0.09 nm[551]. Later, a great deal of MNF-based WGM lasers have been reported[552554]. More recently, in a Au nanorod-coupled dye-doped polymer MNF structure, Zhou et al. experimentally observed the lasing action of the hybrid photon-plasmon mode [see Fig. 22(b)][417]. Benefitting from the strong mode coupling-enabled loss reduction of the WGM in thin MNFs, laser output was observed with MNF diameter down to 2 µm.

Fig. 22

Optical MNF-based fiber lasers. (a) Single-longitudinal-mode laser emission in an Er3+/Yb3+-doped phosphate glass MNF knot[391]. Inset: optical microscope image of the MNF knot. Clear green upconverted photoluminescence is excited by a 975-nm-wavelength light. (b) Output spectrum of the hybrid photon-plasmon lasing emission in a Au-nanorod-coupled dye-doped polymer MNF structure[417]. The insets show the optical microscope image (left) and SEM image (right) of the lasing structure (2 µm in MNF diameter). (c) Schematic of a mode-locked Yb3+-doped ultrafast fiber laser integrated with silica MNFs inside and outside the laser cavity (upper panel)[83]. WDM, wavelength division multiplex; ISO, isolator; PBS, polarization beam splitter; λ/4 (λ/2), quarter-(half)-wave plate. The middle panel shows the output spectra of fiber lasers with (red solid line) and without (blue solid line) the dechirping MNF outside the cavity. For reference, the output spectrum of the fiber laser without the intracavity MNF is shown in the black dashed line. The bottom panel shows interferometric autocorrelation signals of fiber lasers with (red) and without (blue) the dechirping optical MNF. (d) Schematic of a high-repetition-rate ultrafast mode-locked laser based on a hybrid plasmonic MNF resonator (upper panel)[374]. Insets: optical microscope image of the employed MNF knot resonator and SEM image of the MNF. The output spectrum of the fiber laser in the middle panel manifests that the generated pulses have a high repetition rate of 144.3 GHz around 1550-nm wavelength. The bottom panel shows the corresponding autocorrelation trace.

PI_3_1_R02_f022.png

5.5.2.

MNF-based ultrafast mode-locked pulse lasers

Typically, the diameter-dependent waveguide dispersion of an MNF can be two orders of magnitude higher than that in a standard optical fiber, and can be either positive or negative for compensating material dispersion at almost any wavelength. Meanwhile, the adiabatic transition of waveguiding modes between the standard fiber and the MNF enables a compact and fiber-compatible dispersion management with an insertion loss below 0.1 dB. In 2006, Rusu et al. integrated a 20-cm-long optical MNF (1.8 µm in diameter) into the cavity of an Yb3+-doped mode-locked fiber laser to offset the intracavity normal chromatic dispersion, and demonstrated a reduction of pulse duration from 8 to 3 ps[209]. To obtain a shorter pulse duration, in 2018 Wang et al. employed a 25-cm-long, 1-μm-diameter optical MNF for intracavity dispersion compensation in an Yb3+-doped mode-locked fiber laser and another MNF outside the laser cavity for output dechirping, and realized a single-pulse output with a pulse duration of 110 fs, repetition rate of 120 MHz, and output power of 60 mW [see Fig. 22(c)][83]. Shortly after, based on the dispersion-management technique, Yang et al. developed an ultrafast Yb3+-doped all-fiber laser, with a shorter pulse duration of 65 fs, repetition rate of 66.1 MHz, and output power of 28  mW[555]. Through effective dispersion and nonlinearity management, the broadband noise-like pulse could also be extracted from an MNF-integrated Yb3+-doped fiber laser, exhibiting an optical spectrum spanning from below 1 µm to beyond 1.6 µm[210]. In 2020, Li et al. used a 10-cm-long 1-μm-diameter silica MNF to compensate for the intracavity anomalous dispersion and demonstrated a Tm3+-doped dissipative soliton fiber laser at 2-μm wavelength, with a pulse duration of 195 fs, a repetition rate of 49 MHz, and an output power of 25 mW[82]. In the same year, relying on a hybrid plasmonic MNF knot resonator, Ding et al. reported a high-repetition-rate ultrafast mode-locked laser[374]. As shown in Fig. 22(d), the all-fiber laser can deliver pulses with a repetition rate as high as 144.3 GHz around 1550-nm wavelength. Compared with silica MNFs, chalcogenide-glass (e.g., As2Se3) MNFs possess large nonlinearity and thus can serve as efficient nonlinear media in laser cavities. In 2015, Al-Kadry et al. introduced a passively mode-locked fiber laser with nonlinear polarization rotation based on a 10-cm-long As2Se3-PMMA MNF[556]. The fiber laser could generate wavelength-tunable soliton pulses from 1530 to 1562 nm and noise-like pulses with a central wavelength of 1560 nm. In addition to the fiber lasers mentioned above, there is a kind of ultrafast fiber laser enabled by low-dimension-nanomaterial-integrated MNFs including carbon nanotubes, graphene, and transition-metal dichalcogenides, where the modified MNFs serve as saturable absorbers. This kind of ultrafast fiber laser has been widely studied in the past decade, showing advantages of fast response, high damage threshold, and all-fiber structure. For more details, one can refer to the reviews reported previously[84,348].

5.6.

Atom Optics

Exploring the light-matter interaction at the atomic level is of primary interest in quantum optics. Over the last decade, tremendous progress has been accomplished in optical controlling individual atoms by exploiting nanoscale optical structures[557561], in which MNF-supported atom trapping has considerable merit such as large potential depth, high flexibility, low loss, and high compatibility with fiber systems. With tightly confined evanescent fields around the surface, subwavelength-diameter MNFs have been proven a very promising platform for developing cold-atom-based quantum-optics techniques. To confine cold atoms near the surface of optical MNFs, in 2004 Balykin et al. suggested a theoretical scheme to provide an attractive potential around a silica MNF by waveguiding a red-detuned light along the MNF[557]. The red-detuned light far from resonance generated an intensity-dependent gradient force, which could be used to balance the centrifugal force of a moving atom. It was predicted that for a silica MNF (400 nm in diameter) that waveguided a 27-mW-power 1.3-μm-wavelength light, cesium atoms could be trapped at a temperature of less than 0.29 mK and guided along the MNF. Subsequently, Le Kien et al. proposed an improved scheme by waveguiding two-color evanescent fields (i.e., red-detuned 1.06-μm-wavelength and blue-detuned 700-nm-wavelength lights) along a 400-nm-diameter silica MNF[558], demonstrating a stable trapping potential for cesium atoms with a trap depth of 2.9 mK. In 2008, Fu et al. proposed using a red-detuned light excited in the superposition of the HE11 and TE01 modes and a blue-detuned light excited in the HE11 mode to produce 1D lattice potentials with trapping depths larger than 363.1 µK[562]. In 2010, Vetsch et al. experimentally realized MNF-supported atom trapping[559]. By launching a red-detuned light (1064 nm in wavelength) and a blue-detuned light (780 nm in wavelength) in a 500-nm-diameter silica MNF, they created a trapping potential for localizing cold cesium atoms in a 1D optical lattice close to (200  nm) the MNF surface [see Fig. 23(a)]. To visualize the trapped atoms, they used a probe laser resonant with the atomic transition (852 nm in wavelength) to excite the fluorescence light of the trapped atomic ensemble. Figure 23(b) presents the transmission spectrum of the probe light with respect to its detuning, showing a strong absorption after loading the MNF, which can be attributed to a significant growth in the number of trapped atoms in the evanescent field. After that, a variety of optically trapping technologies have been proposed, allowing in-depth study on the quantum properties of the MNF-trapped-atom system[563571].

Fig. 23

MNF-based atom optics. (a) Schematic of the MNF-based atom trapping in the evanescent field (upper panel)[559]. Fluorescence image of a trapped ensemble of cesium atoms (lower panel). (b) Transmission spectrum of a probe beam waveguided along the MNF after loading the trap (black squares)[559]. Green line: the measured spectrum of a magneto-optical trap (MOT) cloud. Red line: theoretical fit. (c) Schematic of storage of MNF-guided light based on the EIT in an evanescent-field configuration[573]. An ensemble of cold cesium atoms is spatially overlapped with a silica MNF (400 nm in diameter). The signal pulse to be stored is waveguided inside the MNF while the control light propagates outside the MNF with an angle of 13°. (d) Transmitted pulses with different control powers in (I). The reference is measured in the absence of atoms. (II) Storage and retrieval processes. In the absence of the control field, the blue and purple points give the transmitted pulses without and with atoms. The red data indicate the memory sequence, showing leakage and retrieval. The black line represents the control timing. After the end of the input pulse, the reference and absorption curves are superimposed and correspond to the background noise level. (III) Normalized efficiency versus the control linear polarization angle. The zero-angle corresponds to a vertical polarization. (e) Schematic of a fiber ring cavity containing an MNF section for collective strong coupling of cold atoms with a cavity mode[576]. DM, dichroic mirror; APD, avalanche photo diode. (f) Normalized transmission of the cavity as the probe laser frequency is scanned across the atomic resonance with input powers of 30 pW and 2.3 nW. The blue circles show data for a cavity in the absence of atoms with a Lorentzian fit. The red crosses correspond to an ensemble of atoms interacting with the cavity mode, with the theoretical fit shown as a red solid line.

PI_3_1_R02_f023.png

Meanwhile, in a fiber-coupled atom system, benefitting from the strong interaction between the waveguiding evanescent field and the atoms nearby, it is possible to realize the electromagnetically induced transparency (EIT) process when the EIT condition is met, as predicted by Patnaik et al. in 2002[572]. In 2015, using a dynamic EIT protocol, Gouraud et al. experimentally demonstrated reversible quantum storage of the MNF-guided light using a cloud of cold cesium atoms [see Fig. 23(c)][573], in which the storage and retrieval processes are shown in Fig. 23(d). One can see that the laser pulses at the single-photon level (0.6 in mean photon number per pulse) are stored in and retrieved after around 650 ns, with an efficiency of 10%±0.5% and signal-to-noise ratio of 20. To ensure a stable storage process, controlling the polarization in the MNF is crucial. Following this approach, Sayrin et al. studied the propagation of a probe pulse under EIT conditions and demonstrated slow light with a group velocity of 50 m/s[574]. Meanwhile, they stored the optical pulses at the single-photon level and retrieved them after 2 µs, with an overall efficiency of 3.0%±0.4%.

It is worth noting that the strength of light-atom interaction is expected to be further enhanced in MNF cavities. In 2015, Kato and Aoki first reported the strong coupling between a trapped single cesium atom and an MNF FP cavity[575]. Well-resolved vacuum Rabi splitting was observed in the cavity transmission spectrum when an atom was trapped in a state-insensitive MNF trap. In 2017, to achieve collective strong coupling with cold cesium atoms in the weak driving limit, Ruddell et al. constructed an alternative all-fiber ring cavity[576]. As shown in Fig. 23(e), for a low input power (e.g., <1  nW), clear splitting of the cavity resonance is obtained owing to a collective enhancement by an ensemble of atoms interacting with the cavity mode [see Fig. 23(f)]. With further increasing input power, the splitting is gradually reduced and eventually disappears (e.g., with a power of 2.3 nW).

In addition, the adiabatic evolution of waveguiding modes between the MNF and standard optical fiber enables effective collection, coupling, and low-loss propagation of single-photon sources from MNF-coupled quantum emitters to standard optical fibers. In recent years, A wide range of MNF-based functionalized structures have been demonstrated for highly efficient single-photon collection, such as nanoscale cavities[577579], 2D-material-integrated MNFs[352], and twin MNF[194] structures. These high-efficiency fiber-coupled single-photon sources may play a role in quantum techniques.

As a miniature fiber-optic platform for atom optics, MNF-based quantum photonics and technology have experienced rapid development in recent years. These high-efficiency fiber-coupled single-photon sources may play a role in quantum techniques. We apologize that we cannot cover the whole content of this topic; a more comprehensive introduction can be found in recent review papers[8688].

5.7.

More Applications

Besides the applications mentioned above, in recent years, there are many other versatile MNF-based photonic devices or techniques that have been reported, ranging from optical filters[390,580582], couplers[453,454,583586], modulators[345,347,587591], optical autocorrelators[326329], and spectrometers[592] to far-field subwavelength focusing[593] and super-resolution imaging[594]. For example, in 2009 Wang et al. proposed to focus optical beams with subwavelength resolution in the far field using an MNF array[593]. In 2013, Hao et al. exploited the evanescent waves of silica MNFs to illuminate the sample in the near field and demonstrated super-resolution imaging at the far field in a single snapshot with a spatial resolution of tens of nanometers[594]. More recently, Relying on the leaky modes generated from non-adiabatic optical MNFs, Cen et al. developed a low-cost, scalable spectrometer with a picometer resolution and sub-millimeter footprint[592].

6.

Conclusion and Outlook

Over the past two decades, we have witnessed rapid progress in MNF optics and related technology. As a unique one-dimensional cylinder with a highly symmetric structure and nearly perfect surface quality and diameter uniformity, glass MNF can offer extraordinarily low waveguiding loss (e.g., 0.03 dB/m at 780-nm wavelength in silica MNFs[37]), nearly 100% power in evanescent waves, high waveguided power density (>20  W/μm2 for silica MNF at 1550-nm wavelength[54]), large length (e.g., >1  m), and a mechanical strength approaching the theoretical limit (e.g., >10  GPa for silica MNFs[61]), which are far beyond the reach of all other optical waveguides with similar mode sizes, ensuring its ability to continuously challenge the limits of light-based technology. Also, as an excellent platform merging fiber optics and nanotechnology, the optical MNF will continue to open up new frontiers of fiber optics and nanophotonics, as we have seen in the fields of MNF-based sensors and atom optics. Finally, we would like to end this review by looking into the future, regarding the challenges and opportunities of the MNF optics and technologies, as follows.

  • (1) Exploring the fundamental limits of MNF optics. Despite significant advances in the past two decades, there is plenty of space to go further, from fabrication, characterization, and functionalization of optical MNFs with higher precision to improving the performance of MNF-based light confining, waveguiding, sensing, lasing, and atom/molecule manipulation. One promising approach is to further improve and optimize the geometry of optical MNFs. For example, the surface roughness of an MNF may be reduced by suppression of surface capillary waves, and lower scattering loss may be achieved through high-pressure treatment[595]. Relying on the nearly perfect surface quality and diameter uniformity of the silica MNF, very recently Yang et al. predicted that a pair of strongly coupled silica MNFs can offer an optical field with a spatial confinement down to 0.15 nm and a peak-to-background ratio of about 20 dB[198], well beyond the reach of all other means. Such an atomic-sized optical confinement is promising to push the limits of MNF-based technologies ranging from optical nanoscopy, spectroscopy, and sensing to atom/molecule manipulation. Additionally, a high-quality smooth surface with fewer defects is critical to pursue the upper limit of waveguiding power in an optical MNF, which may be desired in high-power MNF optics.

  • (2) Expanding available MNF material systems. Since the ability of harnessing light of an MNF is intrinsically determined by its material (i.e., the polarization of the material), to adapt the MNF technology to a wider range of optical applications, it is necessary to expand the existing material systems. For example, with existing MNFs, it is very difficult to low-loss waveguide a vacuum ultraviolet (VUV, with a wavelength <200  nm) light due to material absorption, even with a silica MNF. Compared with the silica MNF, a recently demonstrated ice MNF[126], with much lower intrinsic material absorption at a wavelength shorter than 200 nm, offers the possibility for low-loss waveguiding in the VUV spectral range, and so far, it is possible to fabricate such a waveguide only in the form of an MNF. In the MIR region, chalcogenide-glass MNFs have been successfully demonstrated for optical waveguiding[147,235,236] and supercontinuum generation[302,304]. More fiber materials (e.g., with lower absorption) and techniques (e.g., for efficient in/out coupling) can be explored for lower-loss optical waveguiding and higher power operation with MNFs.

  • (3) MNF-based optical technology: from innovation to application. To date, MNF-based optics and technologies, including near-field coupling, atom optics, and optical sensors, have been employed in scientific research or prototype applications. However, compared with the mature fiber-optic technology, for real applications, there are many challenges regarding cost-effective fabrication, high-precision manipulation, and high-repeatability manufacturing of MNFs and related structures. In this regard, technological improvement and innovation are highly desired. For example, a high-yield parallel-fabrication of silica MNFs has been demonstrated[149]. Also, to precisely assemble the MNFs for practical applications, mature transferring, manipulating, and encapsulating systems with high stability and control accuracy are urgently needed.

Acknowledgments

This work was supported by the New Cornerstone Science Foundation (No. NCI202216); the National Natural Science Foundation of China (Nos. 62175213 and 92150302); the Natural Science Foundation of Zhejiang Province (No. LR21F050002); the Fundamental Research Funds for the Central Universities (No. 2023QZJH27); the National Key Research and Development Project of China (No. 2018YFB2200404).

References

1. 

C. V. Boys, “On the production, properties, and some suggested uses of the finest threads,” Proc. Phys. Soc. London, 9 8 https://doi.org/10.1088/1478-7814/9/1/303 PPSOAU 0370-1328 (1887). Google Scholar

2. 

A. Einstein, “Experimenteller nachweis der ampereschen molekularströme,” Naturwissenschaften, 3 237 https://doi.org/10.1007/BF01546392 NATWAY 0028-1042 (1915). Google Scholar

3. 

A. Einstein and W. J. De Haas, “Experimental proof of the existence of ampère’s molecular currents,” Proc. KNAW, 181 696 (1915). Google Scholar

4. 

D. Hondros and P. Debye, “Electromagnetic waves in dielectrical wires,” Ann. Phys., 337 465 https://doi.org/10.1002/andp.19103370802 (1910). Google Scholar

5. 

H. M. Barlow and A. L. Cullen, “Surface waves,” Proc. IEEE, 100 329 IEEPAD 0018-9219 (1953). Google Scholar

6. 

H. M. Barlow and A. E. Karbowiak, “An investigation of the characteristics of cylindrical surface waves,” Proc. IEEE, 100 321 IEEPAD 0018-9219 (1953). Google Scholar

7. 

E. Snitzer, “Optical wave-guide modes in small glass fibers. 1. theoretical,” J. Opt. Soc. Am., 49 1128 JOSAAH 0030-3941 (1959). Google Scholar

8. 

H. Osterberg et al., “Optical wave-guide modes in small glass fibers. 2. experimental,” J. Opt. Soc. Am., 49 1128 JOSAAH 0030-3941 (1959). Google Scholar

9. 

J. Hecht, City of Light: The Story of Fiber Optics, Oxford University Press, (1999). Google Scholar

10. 

A. C. S. Vanheel, “A new method of transporting optical images without aberrations,” Nature, 173 39 https://doi.org/10.1038/173039a0 (1954). Google Scholar

11. 

H. H. Hopkins and N. S. Kapany, “A flexible fibrescope, using static scanning,” Nature, 173 39 https://doi.org/10.1038/173039b0 (1954). Google Scholar

12. 

N. S. Kapany, “High-resolution fibre optics using sub-micron multiple fibres,” Nature, 184 881 https://doi.org/10.1038/184881a0 (1959). Google Scholar

13. 

K. C. Kao and G. A. Hockham, “Dielectric-fibre surface waveguides for optical frequencies,” Proc. Inst. Electr. Eng., 113 1151 https://doi.org/10.1049/piee.1966.0189 (1966). Google Scholar

14. 

J. D. Love and W. M. Henry, “Quantifying loss minimization in single-mode fiber tapers,” Electron. Lett., 22 912 https://doi.org/10.1049/el:19860622 ELLEAK 0013-5194 (1986). Google Scholar

15. 

W. K. Burns et al., “Loss mechanisms in single-mode fiber tapers,” J. Lightwave Technol., 4 608 https://doi.org/10.1109/JLT.1986.1074764 JLTEDG 0733-8724 (1986). Google Scholar

16. 

J. D. Love et al., “Tapered single-mode fibers and devices. Part 1: adiabaticity criteria,” IEEE Proc. J. Optoelectron., 138 343 https://doi.org/10.1049/ip-j.1991.0060 (1991). Google Scholar

17. 

R. J. Black et al., “Tapered single-mode fibers and devices. Part 2: experimental and theoretical quantification,” IEEE Proc. J. Optoelectron., 138 355 https://doi.org/10.1049/ip-j.1991.0061 (1991). Google Scholar

18. 

T. A. Birks and Y. W. Li, “The shape of fiber tapers,” J. Lightwave Technol., 10 432 https://doi.org/10.1109/50.134196 JLTEDG 0733-8724 (1992). Google Scholar

19. 

A. W. Snyder, “Coupled-mode theory for optical fibers,” J. Opt. Soc. Am., 62 1267 https://doi.org/10.1364/JOSA.62.001267 JOSAAH 0030-3941 (1972). Google Scholar

20. 

J. Bures and R. Ghosh, “Power density of the evanescent field in the vicinity of a tapered fiber,” J. Opt. Soc. Am. A, 16 1992 https://doi.org/10.1364/JOSAA.16.001992 JOAOD6 0740-3232 (1999). Google Scholar

21. 

D. T. Cassidy, D. C. Johnson and K. O. Hill, “Wavelength-dependent transmission of monomode optical fiber tapers,” Appl. Opt., 24 945 https://doi.org/10.1364/AO.24.000945 APOPAI 0003-6935 (1985). Google Scholar

22. 

R. Feced et al., “Acoustooptic attenuation filters based on tapered optical fibers,” IEEE J. Sel. Top. Quantum Electron., 5 1278 https://doi.org/10.1109/2944.806753 IJSQEN 1077-260X (1999). Google Scholar

23. 

F. Bilodeau et al., “Compact, low-loss, fused biconical taper couplers: overcoupled operation and antisymmetric supermode cutoff,” Opt. Lett., 12 634 https://doi.org/10.1364/OL.12.000634 OPLEDP 0146-9592 (1987). Google Scholar

24. 

T. E. Dimmick et al., “Carbon dioxide laser fabrication of fused-fiber couplers and tapers,” Appl. Opt., 38 6845 https://doi.org/10.1364/AO.38.006845 APOPAI 0003-6935 (1999). Google Scholar

25. 

H. S. Mackenzie and F. P. Payne, “Evanescent field amplification in a tapered single-mode optical fiber,” Electron. Lett., 26 130 https://doi.org/10.1049/el:19900089 ELLEAK 0013-5194 (1990). Google Scholar

26. 

M. G. Xu et al., “Temperature-independent strain sensor using a chirped Bragg grating in a tapered optical-fiber,” Electron. Lett., 31 823 https://doi.org/10.1049/el:19950542 ELLEAK 0013-5194 (1995). Google Scholar

27. 

C. Bariáin et al., “Optical fiber humidity sensor based on a tapered fiber coated with agarose gel,” Sens. Actuator B-Chem., 69 127 https://doi.org/10.1016/S0925-4005(00)00524-4 (2000). Google Scholar

28. 

T. A. Birks, W. J. Wadsworth and P. S. J. Russell, “Supercontinuum generation in tapered fibers,” Opt. Lett., 25 1415 https://doi.org/10.1364/OL.25.001415 OPLEDP 0146-9592 (2000). Google Scholar

29. 

L. M. Tong et al., “Subwavelength-diameter silica wires for low-loss optical wave guiding,” Nature, 426 816 https://doi.org/10.1038/nature02193 (2003). Google Scholar

30. 

L. M. Tong, J. Y. Lou and E. Mazur, “Single-mode guiding properties of subwavelength-diameter silica and silicon wire waveguides,” Opt. Express, 12 1025 https://doi.org/10.1364/OPEX.12.001025 OPEXFF 1094-4087 (2004). Google Scholar

31. 

S. G. Leon-Saval et al., “Supercontinuum generation in submicron fibre waveguides,” Opt. Express, 12 2864 https://doi.org/10.1364/OPEX.12.002864 OPEXFF 1094-4087 (2004). Google Scholar

32. 

X. Guo, Y. B. Ying and L. M. Tong, “Photonic nanowires: from subwavelength waveguides to optical sensors,” Accounts Chem. Res., 47 656 https://doi.org/10.1021/ar400232h (2014). Google Scholar

33. 

G. Brambilla, V. Finazzi and D. Richardson, “Ultra-low-loss optical fiber nanotapers,” Opt. Express, 12 2258 https://doi.org/10.1364/OPEX.12.002258 OPEXFF 1094-4087 (2004). Google Scholar

34. 

L. M. Tong et al., “Self-modulated taper drawing of silica nanowires,” Nanotechnology, 16 1445 https://doi.org/10.1088/0957-4484/16/9/004 NNOTER 0957-4484 (2005). Google Scholar

35. 

G. Brambilla, F. Xu and X. Feng, “Fabrication of optical fibre nanowires and their optical and mechanical characterisation,” Electron. Lett., 42 517 https://doi.org/10.1049/el:20060611 ELLEAK 0013-5194 (2006). Google Scholar

36. 

F. Orucevic, V. Lefèvre-Seguin and J. Hare, “Transmittance and near-field characterization of sub-wavelength tapered optical fibers,” Opt. Express, 15 13624 https://doi.org/10.1364/OE.15.013624 OPEXFF 1094-4087 (2007). Google Scholar

37. 

J. E. Hoffman et al., “Ultrahigh transmission optical nanofibers,” AIP Adv., 4 067124 https://doi.org/10.1063/1.4879799 AAIDBI 2158-3226 (2014). Google Scholar

38. 

Y. X. Xu, W. Fang and L. M. Tong, “Real-time control of micro/nanofiber waist diameter with ultrahigh accuracy and precision,” Opt. Express, 25 10434 https://doi.org/10.1364/OE.25.010434 OPEXFF 1094-4087 (2017). Google Scholar

39. 

Y. Kang et al., “Ultrahigh-precision diameter control of nanofiber using direct mode cutoff feedback,” IEEE Photon. Technol. Lett., 32 219 https://doi.org/10.1109/LPT.2020.2966804 IPTLEL 1041-1135 (2020). Google Scholar

40. 

N. Yao et al., “Ultra-long subwavelength micro/nanofibers with low loss,” IEEE Photon. Technol. Lett., 32 1069 https://doi.org/10.1109/LPT.2020.3011719 IPTLEL 1041-1135 (2020). Google Scholar

41. 

L. Shi et al., “Fabrication of submicron-diameter silica fibers using electric strip heater,” Opt. Express, 14 5055 https://doi.org/10.1364/OE.14.005055 OPEXFF 1094-4087 (2006). Google Scholar

42. 

L. Ding et al., “Ultralow loss single-mode silica tapers manufactured by a microheater,” Appl. Opt., 49 2441 https://doi.org/10.1364/AO.49.002441 APOPAI 0003-6935 (2010). Google Scholar

43. 

C. J. Ma et al., “Design and fabrication of tapered microfiber waveguide with good optical and mechanical performance,” J. Mod. Opt., 61 683 https://doi.org/10.1080/09500340.2014.909541 JMOPEW 0950-0340 (2014). Google Scholar

44. 

H. L. Sørensen, E. S. Polzik and J. Appel, “Heater self-calibration technique for shape prediction of fiber tapers,” J. Lightwave Technol., 32 1886 https://doi.org/10.1109/JLT.2014.2314319 JLTEDG 0733-8724 (2014). Google Scholar

45. 

Y. Yu et al., “Precise control of the optical microfiber tapering process based on monitoring of intermodal interference,” Appl. Opt., 53 8222 https://doi.org/10.1364/AO.53.008222 APOPAI 0003-6935 (2014). Google Scholar

46. 

F. Bayle and J. P. Meunier, “Efficient fabrication of fused-fiber biconical taper structures by a scanned CO2 laser beam technique,” Appl. Opt., 44 6402 https://doi.org/10.1364/AO.44.006402 APOPAI 0003-6935 (2005). Google Scholar

47. 

J. M. Ward et al., “Heat-and-pull rig for fiber taper fabrication,” Rev. Sci. Instrum., 85 111501 https://doi.org/10.1063/1.4901098 RSINAK 0034-6748 (2006). Google Scholar

48. 

L. Ç. Özcan et al., “Highly symmetric optical fiber tapers fabricated with a CO2 laser,” IEEE Photon. Technol. Lett., 19 656 https://doi.org/10.1109/LPT.2007.894963 IPTLEL 1041-1135 (2007). Google Scholar

49. 

J. M. Ward et al., “Contributed review: optical micro- and nanofiber pulling rig,” Rev. Sci. Instrum., 85 111501 https://doi.org/10.1063/1.4901098 RSINAK 0034-6748 (2014). Google Scholar

50. 

Y. Kang et al., “Fabrication methods and high-precision diameter control techniques of optical micro-/nanofibers,” Sci. Sin.-Phys. Mech. Astron., 50 084212 https://doi.org/10.1360/SSPMA-2020-0027 (2020). Google Scholar

51. 

D. X. Dai, J. Bauters and J. E. Bowers, “Passive technologies for future large-scale photonic integrated circuits on silicon: polarization handling, light non-reciprocity and loss reduction,” Light Sci. Appl., 1 e1 https://doi.org/10.1038/lsa.2012.1 (2012). Google Scholar

52. 

J. Chen et al., “Real-time measurement and control of nanofiber diameters using a femtowatt photodetector,” Opt. Express, 30 12008 https://doi.org/10.1364/OE.453599 OPEXFF 1094-4087 (2022). Google Scholar

53. 

S. K. Ruddell et al., “Ultra-low-loss nanofiber Fabry-Perot cavities optimized for cavity quantum electrodynamics,” Opt. Lett., 45 4875 https://doi.org/10.1364/OL.396725 OPLEDP 0146-9592 (2020). Google Scholar

54. 

J. B. Zhang et al., “High-power continuous-wave optical waveguiding in a silica micro/nanofibre,” Light Sci. Appl., 12 89 https://doi.org/10.1038/s41377-023-01109-2 (2023). Google Scholar

55. 

G. Brambilla et al., “Compound-glass optical nanowires,” Electron. Lett., 41 400 https://doi.org/10.1049/el:20058381 ELLEAK 0013-5194 (2005). Google Scholar

56. 

L. M. Tong et al., “Photonic nanowires directly drawn from bulk glasses,” Opt. Express, 14 82 https://doi.org/10.1364/OPEX.14.000082 OPEXFF 1094-4087 (2006). Google Scholar

57. 

E. C. Mägi et al., “Enhanced Kerr nonlinearity in sub-wavelength diameter As2Se3 chalcogenide fiber tapers,” Opt. Express, 15 10324 https://doi.org/10.1364/OE.15.010324 OPEXFF 1094-4087 (2007). Google Scholar

58. 

G. Brambilla et al., “Optical fiber nanowires and microwires: fabrication and applications,” Adv. Opt. Photonics, 1 107 https://doi.org/10.1364/AOP.1.000107 AOPAC7 1943-8206 (2009). Google Scholar

59. 

X. Q. Wu and L. M. Tong, “Optical microfibers and nanofibers,” Nanophotonics, 2 407 https://doi.org/10.1515/nanoph-2013-0033 (2013). Google Scholar

60. 

R. Ismaeel et al., “Optical microfiber passive components,” Laser Photon. Rev., 7 350 https://doi.org/10.1002/lpor.201200024 (2013). Google Scholar

61. 

G. Brambilla and D. N. Payne, “The ultimate strength of glass silica nanowires,” Nano Lett., 9 831 https://doi.org/10.1021/nl803581r NALEFD 1530-6984 (2009). Google Scholar

62. 

S. Holleis et al., “Experimental stress-strain analysis of tapered silica optical fibers with nanofiber waist,” Appl. Phys. Lett., 104 163109 https://doi.org/10.1063/1.4873339 APPLAB 0003-6951 (2014). Google Scholar

63. 

A. Godet et al., “Nonlinear elasticity of silica nanofiber,” APL Photonics, 4 080804 https://doi.org/10.1063/1.5103239 (2019). Google Scholar

64. 

L. Zhang, J. Y. Lou and L. M. Tong, “Micro/nanofiber optical sensors,” Photonic Sens., 1 31 https://doi.org/10.1007/s13320-010-0022-z (2010). Google Scholar

65. 

L. M. Tong, “Micro/nanofibre optical sensors: challenges and prospects,” Sensors, 18 903 https://doi.org/10.3390/s18030903 SNSRES 0746-9462 (2018). Google Scholar

66. 

L. Zhang, Y. Tang and L. M. Tong, “Micro-/nanofiber optics: merging photonics and material science on nanoscale for advanced sensing technology,” iScience, 23 100810 https://doi.org/10.1016/j.isci.2019.100810 (2020). Google Scholar

67. 

M. A. Foster et al., “Nonlinear optics in photonic nanowires,” Opt. Express, 16 1300 https://doi.org/10.1364/OE.16.001300 OPEXFF 1094-4087 (2008). Google Scholar

68. 

G. P. Agrawal, Nonlinear Fiber Optics, fifth editionElsevier, Academic Press, (2013). Google Scholar

69. 

F. Xu, Z. X. Wu and Y. Q. Lu, “Nonlinear optics in optical-fiber nanowires and their applications,” Prog. Quantum Electron., 55 35 https://doi.org/10.1016/j.pquantelec.2017.07.003 PQUEAH 0079-6727 (2017). Google Scholar

70. 

D. A. Akimov et al., “Generation of a spectrally asymmetric third harmonic with unamplified 30-fs Cr:forsterite laser pulses in a tapered fiber,” Appl. Phys. B, 76 515 https://doi.org/10.1007/s00340-002-1080-8 (2003). Google Scholar

71. 

V. Grubsky and A. Savchenko, “Glass micro-fibers for efficient third harmonic generation,” Opt. Express, 13 6798 https://doi.org/10.1364/OPEX.13.006798 OPEXFF 1094-4087 (2005). Google Scholar

72. 

M. A. Gouveia et al., “Second harmonic generation and enhancement in microfibers and loop resonators,” Appl. Phys. Lett., 102 201120 https://doi.org/10.1063/1.4807767 APPLAB 0003-6951 (2013). Google Scholar

73. 

J. C. Beugnot et al., “Brillouin light scattering from surface acoustic waves in a subwavelength-diameter optical fibre,” Nat. Commun., 5 5242 https://doi.org/10.1038/ncomms6242 NCAOBW 2041-1723 (2014). Google Scholar

74. 

A. Godet et al., “Brillouin spectroscopy of optical microfibers and nanofibers,” Optica, 4 1232 https://doi.org/10.1364/OPTICA.4.001232 (2017). Google Scholar

75. 

Y. H. Li, Y. Y. Zhao and L. J. Wang, “Demonstration of almost octave-spanning cascaded four-wave mixing in optical microfibers,” Opt. Lett., 37 3441 https://doi.org/10.1364/OL.37.003441 OPLEDP 0146-9592 (2012). Google Scholar

76. 

L. Y. Shan et al., “Design of nanofibres for efficient stimulated Raman scattering in the evanescent field,” J. Eur. Opt. Soc.-Rapid Publ., 8 13030 https://doi.org/10.2971/jeos.2013.13030 (2013). Google Scholar

77. 

L. Y. Shan et al., “Stimulated Raman scattering in the evanescent field of liquid immersed tapered nanofibers,” Appl. Phys. Lett., 102 201110 https://doi.org/10.1063/1.4807170 APPLAB 0003-6951 (2013). Google Scholar

78. 

C. M. B. Cordeiro et al., “Engineering the dispersion of tapered fibers for supercontinuum generation with a 1064 nm pump laser,” Opt. Lett., 30 1980 https://doi.org/10.1364/OL.30.001980 OPLEDP 0146-9592 (2005). Google Scholar

79. 

M. A. Foster et al., “Nonlinear pulse propagation and supercontinuum generation in photonic nanowires: experiment and simulation,” Appl. Phys. B, 81 363 https://doi.org/10.1007/s00340-005-1865-7 (2005). Google Scholar

80. 

A. Hartung, A. M. Heidt and H. Bartelt, “Nanoscale all-normal dispersion optical fibers for coherent supercontinuum generation at ultraviolet wavelengths,” Opt. Express, 20 13777 https://doi.org/10.1364/OE.20.013777 OPEXFF 1094-4087 (2012). Google Scholar

81. 

M. A. Foster et al., “Soliton-effect compression of supercontinuum to few-cycle durations in photonic nanowires,” Opt. Express, 13 6848 https://doi.org/10.1364/OPEX.13.006848 OPEXFF 1094-4087 (2005). Google Scholar

82. 

Y. H. Li et al., “Microfiber-enabled dissipative soliton fiber laser at 2 µm,” Opt. Lett., 43 6105 https://doi.org/10.1364/OL.43.006105 OPLEDP 0146-9592 (2018). Google Scholar

83. 

L. Z. Wang et al., “Femtosecond mode-locked fiber laser at 1 µm via optical microfiber dispersion management,” Sci. Rep., 8 4732 https://doi.org/10.1038/s41598-018-23027-9 SRCEC3 2045-2322 (2018). Google Scholar

84. 

Y. H. Li et al., “Optical microfiber-based ultrafast fiber lasers,” Appl. Phys. B, 125 192 https://doi.org/10.1007/s00340-019-7303-z (2019). Google Scholar

85. 

Y. C. Li et al., “Optical fiber technologies for nanomanipulation and biodetection: a review,” J. Lightwave Technol., 39 251 https://doi.org/10.1109/JLT.2020.3023456 JLTEDG 0733-8724 (2021). Google Scholar

86. 

T. Nieddu, V. Gokhroo and S. Nic Chormaic, “Optical nanofibres and neutral atoms,” J. Opt., 18 053001 https://doi.org/10.1088/2040-8978/18/5/053001 (2016). Google Scholar

87. 

P. Solano et al., “Optical nanofibers: a new platform for quantum optics,” Adv. Atom. Mol. Opt. Phys., 66 439 https://doi.org/10.1016/bs.aamop.2017.02.003 AAMPE9 1049-250X (2017). Google Scholar

88. 

K. P. Nayak et al., “Nanofiber quantum photonics,” J. Opt., 20 073001 https://doi.org/10.1088/2040-8986/aac35e (2018). Google Scholar

89. 

W. L. She, J. H. Yu and R. H. Feng, “Observation of a push force on the end face of a nanometer silica filament exerted by outgoing light,” Phys. Rev. Lett., 101 243601 https://doi.org/10.1103/PhysRevLett.101.243601 PRLTAO 0031-9007 (2008). Google Scholar

90. 

H. D. Zheng et al., “Accurate measurement of nanomechanical motion in a fiber-taper nano-optomechanical system,” Appl. Phys. Lett., 115 013104 https://doi.org/10.1063/1.5110272 APPLAB 0003-6951 (2019). Google Scholar

91. 

Y. Zhang et al., “A broadband and low-power light-control-light effect in a fiber-optic nano-optomechanical system,” Nanoscale, 12 9800 https://doi.org/10.1039/C9NR10953F NANOHL 2040-3364 (2020). Google Scholar

92. 

F. Yang et al., “Large evanescently-induced Brillouin scattering at the surrounding of a nanofibre,” Nat. Commun., 13 1432 https://doi.org/10.1038/s41467-022-29051-8 NCAOBW 2041-1723 (2022). Google Scholar

93. 

W. D. Xu et al., “Strong optomechanical interactions with long-lived fundamental acoustic waves,” Optica, 10 206 https://doi.org/10.1364/OPTICA.476764 (2023). Google Scholar

94. 

J. H. Yu, R. H. Feng and W. L. She, “Low-power all-optical switch based on the bend effect of a nm fiber taper driven by outgoing light,” Opt. Express, 17 4640 https://doi.org/10.1364/OE.17.004640 OPEXFF 1094-4087 (2009). Google Scholar

95. 

L. M. Tong, “Brief introduction to optical microfibers and nanofibers,” Front. Optoelectron. China, 3 54 https://doi.org/10.1007/s12200-009-0073-1 (2010). Google Scholar

96. 

G. Brambilla, “Optical fibre nanowires and microwires: a review,” J. Opt., 12 043001 https://doi.org/10.1088/2040-8978/12/4/043001 (2010). Google Scholar

97. 

L. M. Tong and M. Sumetsky, Subwavelength and Nanometer Diameter Optical Fibers, Zhejiang University Press, (2009). Google Scholar

98. 

N. S. Kapany, “Fiber optics. Part I. Optical properties of certain dielectric cylinders,” J. Opt. Soc. Am., 47 413 https://doi.org/10.1364/JOSA.47.000413 JOSAAH 0030-3941 (1957). Google Scholar

99. 

S. Bateson, “Critical study of the optical and mechanical properties of glass fibers,” J. Appl. Phys., 29 13 https://doi.org/10.1063/1.1722934 JAPIAU 0021-8979 (1958). Google Scholar

100. 

K. P. Jedrzejewski et al., “Tapered-beam expander for single-mode optical-fiber gap devices,” Electron. Lett., 22 105 https://doi.org/10.1049/el:19860073 ELLEAK 0013-5194 (1986). Google Scholar

101. 

J. C. Knight et al., “Phase-matched excitation of whispering-gallery-mode resonances by a fiber taper,” Opt. Lett., 22 1129 https://doi.org/10.1364/OL.22.001129 OPLEDP 0146-9592 (1997). Google Scholar

102. 

R. H. Doremus, “Viscosity of silica,” J. Appl. Phys., 92 7619 https://doi.org/10.1063/1.1515132 JAPIAU 0021-8979 (2002). Google Scholar

103. 

P. K. Gupta et al., “Nanoscale roughness of oxide glass surfaces,” J. Non-Cryst. Solids, 262 200 https://doi.org/10.1016/S0022-3093(99)00662-6 JNCSBJ 0022-3093 (2000). Google Scholar

104. 

T. Seydel et al., “Freezing of capillary waves at the glass transition,” Phys. Rev. B, 65 184207 https://doi.org/10.1103/PhysRevB.65.184207 (2002). Google Scholar

105. 

T. Sarlat et al., “Frozen capillary waves on glass surfaces: an AFM study,” Eur. Phys. J. B, 54 121 https://doi.org/10.1140/epjb/e2006-00420-6 EPJBFY 1434-6028 (2006). Google Scholar

106. 

H. J. Kbashi, “Fabrication of submicron-diameter and taper fibers using chemical etching,” J. Mater. Sci. Technol., 28 308 https://doi.org/10.1016/S1005-0302(12)60059-0 (2012). Google Scholar

107. 

D. P. Yu et al., “Amorphous silica nanowires: intensive blue light emitters,” Appl. Phys. Lett., 73 3076 https://doi.org/10.1063/1.122677 APPLAB 0003-6951 (1998). Google Scholar

108. 

N. Irawati et al., “Evanescent wave optical trapping and transport of polystyrene microspheres on microfibers,” Microw. Opt. Technol. Lett., 56 2630 https://doi.org/10.1002/mop.28659 MOTLEO 0895-2477 (2014). Google Scholar

109. 

Q. Yang et al., “Polymer micro or nanofibers for optical device applications,” J. Appl. Polym. Sci., 110 1080 https://doi.org/10.1002/app.28716 JAPNAB 1097-4628 (2008). Google Scholar

110. 

F. X. Gu et al., “Light-emitting polymer single nanofibers via waveguiding excitation,” ACS Nano, 4 5332 https://doi.org/10.1021/nn100775v ANCAC3 1936-0851 (2010). Google Scholar

111. 

C. Meng et al., “Quantum-dot-doped polymer nanofibers for optical sensing,” Adv. Mater., 23 3770 https://doi.org/10.1002/adma.201101392 ADVMEW 0935-9648 (2011). Google Scholar

112. 

F. X. Gu et al., “Polymer single-nanowire optical sensors,” Nano Lett., 8 2757 https://doi.org/10.1021/nl8012314 NALEFD 1530-6984 (2008). Google Scholar

113. 

S. Kujala et al., “Natural silk as a photonics component: a study on its light guiding and nonlinear optical properties,” Sci. Rep., 6 22358 https://doi.org/10.1038/srep22358 SRCEC3 2045-2322 (2016). Google Scholar

114. 

N. Huby et al., “Native spider silk as a biological optical fiber,” Appl. Phys. Lett., 102 123702 https://doi.org/10.1063/1.4798552 APPLAB 0003-6951 (2013). Google Scholar

115. 

J. T. Li et al., “Spider silk-inspired artificial fibers,” Adv. Sci., 9 2103965 https://doi.org/10.1002/advs.202103965 (2022). Google Scholar

116. 

X. G. Yang et al., “Light-emitting microfibers from lotus root for eco-friendly optical waveguides and biosensing,” Nano Lett., 24 566 https://doi.org/10.1021/acs.nanolett.3c03213 NALEFD 1530-6984 (2024). Google Scholar

117. 

H. B. Xin et al., “Escherichia coli-based biophotonic waveguides,” Nano Lett., 13 3408 https://doi.org/10.1021/nl401870d NALEFD 1530-6984 (2013). Google Scholar

118. 

T. Y. Cui et al., “From monomeric nanofibers to PbS nanoparticles/polymer composite nanofibers through the combined use of γ-irradiation and gas/solid reaction,” J. Am. Chem. Soc., 128 6298 https://doi.org/10.1021/ja060517w JACSAT 0002-7863 (2006). Google Scholar

119. 

Y. Dzenis, “Spinning continuous fibers for nanotechnology,” Science, 304 1917 https://doi.org/10.1126/science.1099074 SCIEAS 0036-8075 (2004). Google Scholar

120. 

X. F. Duan and C. M. Lieber, “General synthesis of compound semiconductor nanowires,” Adv. Mater., 12 298 https://doi.org/10.1002/(SICI)1521-4095(200002)12:4<298::AID-ADMA298>3.0.CO;2-Y ADVMEW 0935-9648 (2000). Google Scholar

121. 

P. V. Radovanovic et al., “General synthesis of manganese-doped II-VI and III-V semiconductor nanowires,” Nano Lett., 5 1407 https://doi.org/10.1021/nl050747t NALEFD 1530-6984 (2005). Google Scholar

122. 

J. Johansson et al., “Structural properties of <111>B -oriented III–V nanowires,” Nat. Mater., 5 574 https://doi.org/10.1038/nmat1677 NMAACR 1476-1122 (2006). Google Scholar

123. 

T. Mårtensson et al., “Epitaxial growth of indium arsenide nanowires on silicon using nucleation templates formed by self-assembled organic coatings,” Adv. Mater., 19 1801 https://doi.org/10.1002/adma.200700285 ADVMEW 0935-9648 (2007). Google Scholar

124. 

H. E. Labelle and A. I. Mlavsky, “Growth of sapphire filaments from the melt,” Nature, 216 574 https://doi.org/10.1038/216574b0 (1967). Google Scholar

125. 

R. K. Nubling and J. A. Harrington, “Optical properties of single-crystal sapphire fibers,” Appl. Opt., 36 5934 https://doi.org/10.1364/AO.36.005934 APOPAI 0003-6935 (1997). Google Scholar

126. 

P. Z. Xu et al., “Elastic ice microfibers,” Science, 373 187 https://doi.org/10.1126/science.abh3754 SCIEAS 0036-8075 (2021). Google Scholar

127. 

D. H. Jundt, M. M. Fejer and R. L. Byer, “Characterization of single-crystal sapphire fibers for optical power delivery systems,” Appl. Phys. Lett., 55 2170 https://doi.org/10.1063/1.102348 APPLAB 0003-6951 (1989). Google Scholar

128. 

M. Law, J. Goldberger and P. D. Yang, “Semiconductor nanowires and nanotubes,” Ann. Rev. Mater. Res., 34 83 https://doi.org/10.1146/annurev.matsci.34.040203.112300 ARMRCU 1531-7331 (2004). Google Scholar

129. 

N. Wang, Y. Cai and R. Q. Zhang, “Growth of nanowires,” Mater. Sci. Eng. R-Rep., 60 1 https://doi.org/10.1016/j.mser.2008.01.001 (2008). Google Scholar

130. 

S. Rackauskas and A. G. Nasibulin, “Nanowire growth without catalysts: applications and mechanisms at the atomic scale,” ACS Appl. Nano Mater., 3 7314 https://doi.org/10.1021/acsanm.0c01179 (2020). Google Scholar

131. 

B. W. Cui et al., “Low-dimensional and confined ice,” Ann. Rev. Mater. Res., 53 371 https://doi.org/10.1146/annurev-matsci-080921-101821 ARMRCU 1531-7331 (2023). Google Scholar

132. 

W. Lu and C. M. Lieber, “Semiconductor nanowires,” J. Phys. D-Appl. Phys., 39 R387 https://doi.org/10.1088/0022-3727/39/21/R01 (2006). Google Scholar

133. 

K. S. Shankar and A. K. Raychaudhuri, “Fabrication of nanowires of multicomponent oxides: review of recent advances,” Mater. Sci. Eng. C, 25 738 https://doi.org/10.1016/j.msec.2005.06.054 MSCEEE 0928-4931 (2005). Google Scholar

134. 

M. Hernández-Vélez, “Nanowires and 1D arrays fabrication: an overview,” Thin Solid Films, 495 51 https://doi.org/10.1016/j.tsf.2005.08.331 THSFAP 0040-6090 (2006). Google Scholar

135. 

P. Wang, Y. P. Wang and L. M. Tong, “Functionalized polymer nanofibers: a versatile platform for manipulating light at the nanoscale,” Light Sci. Appl., 2 e102 https://doi.org/10.1038/lsa.2013.58 (2013). Google Scholar

136. 

R. X. Yan, D. Gargas and P. D. Yang, “Nanowire photonics,” Nat. Photonics, 3 569 https://doi.org/10.1038/nphoton.2009.184 NPAHBY 1749-4885 (2009). Google Scholar

137. 

M. Sumetsky, “Optics of tunneling from adiabatic nanotapers,” Opt. Lett., 31 3420 https://doi.org/10.1364/OL.31.003420 OPLEDP 0146-9592 (2006). Google Scholar

138. 

S. Ravets et al., “Intermodal energy transfer in a tapered optical fiber: optimizing transmission,” J. Opt. Soc. Am. A, 30 2361 https://doi.org/10.1364/JOSAA.30.002361 JOAOD6 0740-3232 (2013). Google Scholar

139. 

J. D. Love and W. M. Henry, “Quantifying loss minimisation in single-mode fibre tapers,” Electron. Lett., 22 912 https://doi.org/10.1049/el:19860622 ELLEAK 0013-5194 (1986). Google Scholar

140. 

C. Vass, T. Smausz and B. Hopp, “Wet etching of fused silica: a multiplex study,” J. Phys. D-Appl. Phys., 37 2449 https://doi.org/10.1088/0022-3727/37/17/018 (2004). Google Scholar

141. 

P. Del’Haye et al., “Octave spanning tunable frequency comb from a microresonator,” Phys. Rev. Lett., 107 063901 https://doi.org/10.1103/PhysRevLett.107.063901 PRLTAO 0031-9007 (2011). Google Scholar

142. 

A. M. Clohessy et al., “Short low-loss nanowire tapers on singlemode fibres,” Electron. Lett., 41 954 https://doi.org/10.1049/el:20052367 ELLEAK 0013-5194 (2005). Google Scholar

143. 

A. Stiebeiner, R. Garcia-Fernandez and A. Rauschenbeutel, “Design and optimization of broadband tapered optical fibers with a nanofiber waist,” Opt. Express, 18 22677 https://doi.org/10.1364/OE.18.022677 OPEXFF 1094-4087 (2010). Google Scholar

144. 

F. Bilodeau et al., “Low-loss highly overcoupled fused couplers: fabrication and sensitivity to external pressure,” J. Lightwave Technol., 6 1476 https://doi.org/10.1109/50.7904 JLTEDG 0733-8724 (1988). Google Scholar

145. 

A. Felipe et al., “Stepwise fabrication of arbitrary fiber optic tapers,” Opt. Express, 20 19893 https://doi.org/10.1364/OE.20.019893 OPEXFF 1094-4087 (2012). Google Scholar

146. 

S. W. Harun et al., “Theoretical analysis and fabrication of tapered fiber,” Optik, 124 538 https://doi.org/10.1016/j.ijleo.2011.12.054 OTIKAJ 0030-4026 (2013). Google Scholar

147. 

D. W. Cai et al., “Chalcogenide glass microfibers for mid-infrared optics,” Photonics, 8 497 https://doi.org/10.3390/photonics8110497 (2021). Google Scholar

148. 

Y. Xie et al., “Mid-infrared chalcogenide microfiber knot resonators,” Photonics Res., 8 616 https://doi.org/10.1364/PRJ.386395 (2020). Google Scholar

149. 

H. B. Fang et al., “Parallel fabrication of silica optical microfibers and nanofibers,” Light Adv. Manuf., (2024). Google Scholar

150. 

C. E. Chryssou, “Theoretical analysis of tapering fused silica optical fibers using a carbon dioxide laser,” Opt. Eng., 38 1645 https://doi.org/10.1117/1.602271 (1999). Google Scholar

151. 

A. J. C. Grellier, N. K. Zayer and C. N. Pannell, “Heat transfer modelling in CO2 laser processing of optical fibres,” Opt. Commun., 152 324 https://doi.org/10.1016/S0030-4018(98)00164-3 OPCOB8 0030-4018 (1998). Google Scholar

152. 

M. Sumetsky, Y. Dulashko and A. Hale, “Fabrication and study of bent and coiled free silica nanowires: self-coupling microloop optical interferometer,” Opt. Express, 12 3521 https://doi.org/10.1364/OPEX.12.003521 OPEXFF 1094-4087 (2004). Google Scholar

153. 

Q. N. Jia et al., “Fibre tapering using plasmonic microheaters and deformation-induced pull,” Light Adv. Manuf., 4 5 https://doi.org/10.37188/lam.2023.005 (2023). Google Scholar

154. 

K. J. Huang, S. Y. Yang and L. M. Tong, “Modeling of evanescent coupling between two parallel optical nanowires,” Appl. Opt., 46 1429 https://doi.org/10.1364/AO.46.001429 APOPAI 0003-6935 (2007). Google Scholar

155. 

F. Warken and H. Giessen, “Fast profile measurement of micrometer-sized tapered fibers with better than 50-nm accuracy,” Opt. Lett., 29 1727 https://doi.org/10.1364/OL.29.001727 OPLEDP 0146-9592 (2004). Google Scholar

156. 

D. J. Little and D. M. Kane, “Subdiffraction-limited radius measurements of microcylinders using conventional bright-field optical microscopy,” Opt. Lett., 39 5196 https://doi.org/10.1364/OL.39.005196 OPLEDP 0146-9592 (2014). Google Scholar

157. 

A. Azzoune, P. Delaye and G. Pauliat, “Optical microscopy for measuring tapered fibers beyond the diffraction limit,” Opt. Express, 27 24403 https://doi.org/10.1364/OE.27.024403 OPEXFF 1094-4087 (2019). Google Scholar

158. 

M. Michihata et al., “In-process diameter measurement technique for micro-optical fiber with standing wave illumination,” Nanomanuf. Metrol., 4 28 https://doi.org/10.1007/s41871-020-00081-4 (2021). Google Scholar

159. 

M. Michihata et al., “Measurement of diameter of sub-micrometer fiber based on analysis of scattered light intensity distribution under standing wave illumination,” CIRP Annals, 71 421 https://doi.org/10.1016/j.cirp.2022.03.008 (2022). Google Scholar

160. 

M. Sumetsky et al., “Probing optical microfiber nonuniformities at nanoscale,” Opt. Lett., 31 2393 https://doi.org/10.1364/OL.31.002393 OPLEDP 0146-9592 (2006). Google Scholar

161. 

J. Keloth et al., “Diameter measurement of optical nanofibers using a composite photonic crystal cavity,” Opt. Lett., 40 4122 https://doi.org/10.1364/OL.40.004122 OPLEDP 0146-9592 (2015). Google Scholar

162. 

M. Zhu et al., “Diameter measurement of optical nanofiber based on high-order Bragg reflections using a ruled grating,” Opt. Lett., 43 559 https://doi.org/10.1364/OL.43.000559 OPLEDP 0146-9592 (2018). Google Scholar

163. 

L. S. Madsen et al., “Nondestructive profilometry of optical nanofibers,” Nano Lett., 16 7333 https://doi.org/10.1021/acs.nanolett.6b02460 NALEFD 1530-6984 (2016). Google Scholar

164. 

F. K. Fatemi et al., “Modal interference in optical nanofibers for sub-Angstrom radius sensitivity,” Optica, 4 157 https://doi.org/10.1364/OPTICA.4.000157 (2017). Google Scholar

165. 

P. F. Zhang et al., “Nondestructive measurement of nanofiber diameters using microfiber tip,” Opt. Express, 26 31500 https://doi.org/10.1364/OE.26.031500 OPEXFF 1094-4087 (2018). Google Scholar

166. 

J. E. Hoffman et al., “Rayleigh scattering in an optical nanofiber as a probe of higher-order mode propagation,” Optica, 2 416 https://doi.org/10.1364/OPTICA.2.000416 (2015). Google Scholar

167. 

Y. Haddad et al., “Microscopic imaging along tapered optical fibers by right-angle Rayleigh light scattering in linear and nonlinear regime,” Opt. Express, 29 39159 https://doi.org/10.1364/OE.438703 OPEXFF 1094-4087 (2021). Google Scholar

168. 

Y. H. Lai et al., “Fiber taper characterization by optical backscattering reflectometry,” Opt. Express, 25 22312 https://doi.org/10.1364/OE.25.022312 OPEXFF 1094-4087 (2017). Google Scholar

169. 

F. Jafari et al., “Profilometry of an optical microfiber based on modal evolution,” Opt. Lett., 45 6607 https://doi.org/10.1364/OL.411767 OPLEDP 0146-9592 (2020). Google Scholar

170. 

U. Wiedemann et al., “Measurement of submicrometre diameters of tapered optical fibres using harmonic generation,” Opt. Express, 18 7693 https://doi.org/10.1364/OE.18.007693 OPEXFF 1094-4087 (2010). Google Scholar

171. 

P. F. Jarschel et al., “Fiber taper diameter characterization using forward Brillouin scattering,” Opt. Lett., 43 995 https://doi.org/10.1364/OL.43.000995 OPLEDP 0146-9592 (2018). Google Scholar

172. 

S. A. Harfenist et al., “Direct drawing of suspended filamentary micro- and nanostructures from liquid polymers,” Nano Lett., 4 1931 https://doi.org/10.1021/nl048919u NALEFD 1530-6984 (2004). Google Scholar

173. 

P. Wang et al., “Polymer nanofibers embedded with aligned gold nanorods: a new platform for plasmonic studies and optical sensing,” Nano Lett., 12 3145 https://doi.org/10.1021/nl301055f NALEFD 1530-6984 (2012). Google Scholar

174. 

P. Wang et al., “Electron-beam-activated light-emitting polymer nanofibers,” Opt. Lett., 38 1040 https://doi.org/10.1364/OL.38.001040 OPLEDP 0146-9592 (2013). Google Scholar

175. 

N. Irawati et al., “Fabrication of polymer microfiber by direct drawing,” Microw. Opt. Technol. Lett., 57 820 https://doi.org/10.1002/mop.28967 MOTLEO 0895-2477 (2015). Google Scholar

176. 

Y. Zhao et al., “Continuous melt-drawing of highly aligned flexible and stretchable semiconducting microfibers for organic electronics,” Adv. Funct. Mater., 28 1705584 https://doi.org/10.1002/adfm.201705584 AFMDC6 1616-301X (2018). Google Scholar

177. 

X. B. Xing et al., “Ultracompact photonic coupling splitters twisted by PTT nanowires,” Nano Lett., 8 2839 https://doi.org/10.1021/nl8014507 NALEFD 1530-6984 (2008). Google Scholar

178. 

W. M. Mukhtar et al., “Analysis of biconical taper geometries to the transmission losses in optical microfibers,” Optoelectron. Adv. Mater.-Rapid Commun., 6 988 https://doi.org/https://oam-rc.inoe.ro/articles/analysis-of-biconical-taper-geometries-to-the-transmission-losses-in-optical-microfibers/ (2012). Google Scholar

179. 

P. C. Fan et al., “Higher-order diffraction of long-period microfiber gratings realized by arc discharge method,” Opt. Express, 24 25380 https://doi.org/10.1364/OE.24.025380 OPEXFF 1094-4087 (2016). Google Scholar

180. 

O. Yaghobi and H. R. Karimi-Alavijeh, “Single step process for optical microfiber in-line Mach-Zehnder interferometers fabrication,” IEEE Photon. Technol. Lett., 30 915 https://doi.org/10.1109/LPT.2018.2825296 IPTLEL 1041-1135 (2018). Google Scholar

181. 

C. De Marco et al., “Organic light-emitting nanofibers by solvent-resistant nanofluidics,” Adv. Mater., 20 4158 https://doi.org/10.1002/adma.200703033 ADVMEW 0935-9648 (2008). Google Scholar

182. 

R. S. Wagner and W. C. Ellis, “Vapor-liquid-solid mechanism of single crystal growth,” Appl. Phys. Lett., 4 89 https://doi.org/10.1063/1.1753975 APPLAB 0003-6951 (1964). Google Scholar

183. 

A. M. Morales and C. M. Lieber, “A laser ablation method for the synthesis of crystalline semiconductor nanowires,” Science, 279 208 https://doi.org/10.1126/science.279.5348.208 SCIEAS 0036-8075 (1998). Google Scholar

184. 

Y. Y. Wu, R. Fan and P. D. Yang, “Block-by-block growth of single-crystalline Si/SiGe superlattice nanowires,” Nano Lett., 2 83 https://doi.org/10.1021/nl0156888 NALEFD 1530-6984 (2002). Google Scholar

185. 

R. Grange et al., “Lithium niobate nanowires synthesis, optical properties, and manipulation,” Appl. Phys. Lett., 95 143105 https://doi.org/10.1063/1.3236777 APPLAB 0003-6951 (2009). Google Scholar

186. 

H. Deng et al., “Growth, patterning and alignment of organolead iodide perovskite nanowires for optoelectronic devices,” Nanoscale, 7 4163 https://doi.org/10.1039/C4NR06982J NANOHL 2040-3364 (2015). Google Scholar

187. 

J. G. Feng et al., “Single-crystalline layered metal-halide perovskite nanowires for ultrasensitive photodetectors,” Nat. Electron., 1 404 https://doi.org/10.1038/s41928-018-0101-5 NEREBX 0305-2257 (2018). Google Scholar

188. 

H. K. Yu et al., “Modeling bending losses of optical nanofibers or nanowires,” Appl. Opt., 48 4365 https://doi.org/10.1364/AO.48.004365 APOPAI 0003-6935 (2009). Google Scholar

189. 

Y. X. Yang et al., “Loss reduction in sharply bent optical nanofibers by coupling with Au nanoparticles,” Opt. Commun., 497 127167 https://doi.org/10.1016/j.optcom.2021.127167 OPCOB8 0030-4018 (2021). Google Scholar

190. 

F. Le Kien et al., “Coupling between guided modes of two parallel nanofibers,” New J. Phys., 22 123007 https://doi.org/10.1088/1367-2630/abc8af NJOPFM 1367-2630 (2020). Google Scholar

191. 

F. Le Kien et al., “Spatial distributions of the fields in guided normal modes of two coupled parallel optical nanofibers,” New J. Phys., 23 043006 https://doi.org/10.1088/1367-2630/abea44 NJOPFM 1367-2630 (2021). Google Scholar

192. 

F. Le Kien, S. Nic Chormaic and T. Busch, “Optical trap for an atom around the midpoint between two coupled identical parallel optical nanofibers,” Phys. Rev. A, 103 063106 https://doi.org/10.1103/PhysRevA.103.063106 (2021). Google Scholar

193. 

F. Le Kien, S. Nic Chormaic and T. Busch, “Optical force between two coupled identical parallel optical nanofibers,” Phys. Rev. A, 105 063517 https://doi.org/10.1103/PhysRevA.105.063517 (2022). Google Scholar

194. 

L. Q. Shao et al., “Twin-nanofiber structure for a highly efficient single-photon collection,” Opt. Express, 30 9147 https://doi.org/10.1364/OE.454616 OPEXFF 1094-4087 (2022). Google Scholar

195. 

S. S. Wang et al., “Modeling endface output patterns of optical micro/nanofibers,” Opt. Express, 16 8887 https://doi.org/10.1364/OE.16.008887 OPEXFF 1094-4087 (2008). Google Scholar

196. 

H. Wu et al., “Photonic nanolaser with extreme optical field confinement,” Phys. Rev. Lett., 129 013902 https://doi.org/10.1103/PhysRevLett.129.013902 PRLTAO 0031-9007 (2022). Google Scholar

197. 

L. Yang et al., “Generating a sub-nanometer-confined optical field in a nanoslit waveguiding mode,” Adv. Photonics, 5 046003 https://doi.org/10.1117/1.AP.5.4.046003 AOPAC7 1943-8206 (2023). Google Scholar

198. 

Y. X. Yang et al., “Generating a nanoscale blade-like optical field in a coupled nanofiber pair,” Photonics Res., 12 154 https://doi.org/10.1364/PRJ.506681 (2024). Google Scholar

199. 

J. Y. Lou, L. M. Tong and Z. Z. Ye, “Dispersion shifts in optical nanowires with thin dielectric coatings,” Opt. Express, 14 6993 https://doi.org/10.1364/OE.14.006993 OPEXFF 1094-4087 (2006). Google Scholar

200. 

C. J. Zhao et al., “Field and dispersion properties of subwavelength-diameter hollow optical fiber,” Opt. Express, 15 6629 https://doi.org/10.1364/OE.15.006629 OPEXFF 1094-4087 (2007). Google Scholar

201. 

F. Xu, “Miniature function-integrated devices based on optical microfibers,” J. Appl. Sci., 35 469 (2017). Google Scholar

202. 

A. W. Snyder and J. D. Love, Optical Waveguide Theory, Chapman and Hall, (1983). Google Scholar

203. 

F. Le Kien et al., “Higher-order modes of vacuum-clad ultrathin optical fibers,” Phys. Rev. A, 96 023835 https://doi.org/10.1103/PhysRevA.96.023835 (2017). Google Scholar

204. 

F. Le Kien et al., “Field intensity distributions and polarization orientations in a vacuum-clad subwavelength-diameter optical fiber,” Opt. Commun., 242 445 https://doi.org/10.1016/j.optcom.2004.08.044 OPCOB8 0030-4018 (2004). Google Scholar

205. 

M. Joos, A. Bramati and Q. Glorieux, “Complete polarization control for a nanofiber waveguide using the scattering properties,” Opt. Express, 27 18818 https://doi.org/10.1364/OE.27.018818 OPEXFF 1094-4087 (2019). Google Scholar

206. 

G. Tkachenko, F. Lei and S. Nic Chormaic, “Polarisation control for optical nanofibres by imaging through a single lens,” J. Opt., 21 125604 https://doi.org/10.1088/2040-8986/ab5204 (2019). Google Scholar

207. 

F. C. Lei et al., “Complete polarization control for a nanofiber waveguide using directional coupling,” Phys. Rev. Appl., 11 064041 https://doi.org/10.1103/PhysRevApplied.11.064041 PRAHB2 2331-7019 (2019). Google Scholar

208. 

Q. Y. Bao et al., “Circular-area-equivalence approach for determining propagation constants of a single-mode polygonal nanowire,” J. Opt. Soc. Am. B, 39 795 https://doi.org/10.1364/JOSAB.445738 JOBPDE 0740-3224 (2022). Google Scholar

209. 

M. Rusu et al., “Fiber taper for dispersion management in a mode-locked ytterbium fiber laser,” Opt. Lett., 31 2257 https://doi.org/10.1364/OL.31.002257 OPLEDP 0146-9592 (2006). Google Scholar

210. 

J. Zhou et al., “Broadband noise-like pulse generation at 1 μm via dispersion and nonlinearity management,” Opt. Lett., 46 1570 https://doi.org/10.1364/OL.420002 OPLEDP 0146-9592 (2021). Google Scholar

211. 

L. Cui et al., “Generation of correlated photon pairs in micro/nano-fibers,” Opt. Lett., 38 5063 https://doi.org/10.1364/OL.38.005063 OPLEDP 0146-9592 (2013). Google Scholar

212. 

P. Delaye et al., “Continuous-wave generation of photon pairs in silica nanofibers using single-longitudinal- and multilongitudinal-mode pumps,” Phys. Rev. A, 104 063715 https://doi.org/10.1103/PhysRevA.104.063715 (2021). Google Scholar

213. 

W. J. Li, Y. X. Gao and L. M. Tong, “Crosstalk in two intersecting optical microfibers,” IEEE Photon. Technol. Lett., 31 1514 https://doi.org/10.1109/LPT.2019.2936053 IPTLEL 1041-1135 (2019). Google Scholar

214. 

V. Bondarenko and Y. Zhao, “Needle beam: beyond-diffraction-limit concentration of field and transmitted power in dielectric waveguide,” Appl. Phys. Lett., 89 141103 https://doi.org/10.1063/1.2358310 APPLAB 0003-6951 (2006). Google Scholar

215. 

Y. Nakayama et al., “Tunable nanowire nonlinear optical probe,” Nature, 447 1098 https://doi.org/10.1038/nature05921 (2007). Google Scholar

216. 

A. V. Maslov and C. Z. Ning, “Reflection of guided modes in a semiconductor nanowire laser,” Appl. Phys. Lett., 83 1237 https://doi.org/10.1063/1.1599037 APPLAB 0003-6951 (2003). Google Scholar

217. 

L. K. van Vugt, S. Rühle and D. Vanmaekelbergh, “Phase-correlated nondirectional laser emission from the end facets of a ZnO nanowire,” Nano Lett., 6 2707 https://doi.org/10.1021/nl0616227 NALEFD 1530-6984 (2006). Google Scholar

218. 

Z. Ma et al., “Near-field characterization of optical micro/nanofibres,” Chin. Phys. Lett., 24 3006 https://doi.org/10.1088/0256-307X/24/10/081 CPLEEU 0256-307X (2007). Google Scholar

219. 

Y. Tamura et al., “The first 0.14-dB/km loss optical fiber and its impact on submarine transmission,” J. Lightwave Technol., 36 44 https://doi.org/10.1109/JLT.2018.2796647 JLTEDG 0733-8724 (2018). Google Scholar

220. 

A. Braslau et al., “Capillary waves on the surface of simple liquids measured by x-ray reflectivity,” Phys. Rev. A, 38 2457 https://doi.org/10.1103/PhysRevA.38.2457 (1988). Google Scholar

221. 

M. K. Sanyal et al., “X-ray-scattering study of capillary-wave fluctuations at a liquid surface,” Phys. Rev. Lett., 66 628 https://doi.org/10.1103/PhysRevLett.66.628 PRLTAO 0031-9007 (1991). Google Scholar

222. 

J. Jackle and K. Kawasaki, “Intrinsic roughness of glass surfaces,” J. Phys.-Condes. Matter, 7 4351 https://doi.org/10.1088/0953-8984/7/23/006 (1995). Google Scholar

223. 

G. Y. Zhai and L. M. Tong, “Roughness-induced radiation losses in optical micro or nanofibers,” Opt. Express, 15 13805 https://doi.org/10.1364/OE.15.013805 OPEXFF 1094-4087 (2007). Google Scholar

224. 

A. V. Kovalenko, V. N. Kurashov and A. V. Kisil, “Radiation losses in optical nanofibers with random rough surface,” Opt. Express, 16 5797 https://doi.org/10.1364/OE.16.005797 OPEXFF 1094-4087 (2008). Google Scholar

225. 

M. Sumetsky, “How thin can a microfiber be and still guide light?,” Opt. Lett., 31 870 https://doi.org/10.1364/OL.31.000870 OPLEDP 0146-9592 (2006). Google Scholar

226. 

M. Sumetsky et al., “Thinnest optical waveguide: experimental test,” Opt. Lett., 32 754 https://doi.org/10.1364/OL.32.000754 OPLEDP 0146-9592 (2007). Google Scholar

227. 

A. Hartung, S. Brueckner and H. Bartelt, “Limits of light guidance in optical nanofibers,” Opt. Express, 18 3754 https://doi.org/10.1364/OE.18.003754 OPEXFF 1094-4087 (2010). Google Scholar

228. 

A. Coillet et al., “Near-field characterization of glass microfibers on a low-index substrate,” Appl. Phys. B, 101 291 https://doi.org/10.1007/s00340-010-4056-0 (2010). Google Scholar

229. 

K. Y. Wang et al., “Quasi-guiding modes in microfibers on a high refractive index substrate,” ACS Photonics, 2 1278 https://doi.org/10.1021/acsphotonics.5b00176 (2015). Google Scholar

230. 

X. N. Zhang et al., “Energy attenuations in single microfiber and double-loop cavity supported by optical substrate,” Appl. Opt., 57 9351 https://doi.org/10.1364/AO.57.009351 APOPAI 0003-6935 (2018). Google Scholar

231. 

L. Skuja, “Optically active oxygen-deficiency-related centers in amorphous silicon dioxide,” J. Non-Cryst. Solids, 239 16 https://doi.org/10.1016/S0022-3093(98)00720-0 JNCSBJ 0022-3093 (1998). Google Scholar

232. 

L. Skuja et al., “Defects in oxide glasses,” Phys. Status Solidi, 2 15 https://doi.org/10.1002/pssc.200460102 PSSCGL 1862-6351 (2005). Google Scholar

233. 

F. Liao et al., “Enhancing monolayer photoluminescence on optical micro/nanofibers for low-threshold lasing,” Sci. Adv., 5 eaax7398 https://doi.org/10.1126/sciadv.aax7398 STAMCV 1468-6996 (2019). Google Scholar

234. 

S. Y. Linghu et al., “Thermal-mechanical-photo-activation effect on silica micro/nanofiber surfaces: origination, reparation and utilization,” Opt. Express, 30 22755 https://doi.org/10.1364/OE.460793 OPEXFF 1094-4087 (2022). Google Scholar

235. 

C. Baker and M. Rochette, “High nonlinearity and single-mode transmission in tapered multimode As2Se3-PMMA fibers,” IEEE Photonics J., 4 960 https://doi.org/10.1109/JPHOT.2012.2202103 (2012). Google Scholar

236. 

L. Z. Li et al., “Design, fabrication and characterization of PC, COP and PMMA-cladded As2Se3 microwires,” Opt. Mater. Express, 6 912 https://doi.org/10.1364/OME.6.000912 (2016). Google Scholar

237. 

D. Lee et al., “Fabrication method for ultra-long optical micro/nano-fibers,” Curr. Appl. Phys., 19 1334 https://doi.org/10.1016/j.cap.2019.08.018 1567-1739 (2019). Google Scholar

238. 

Z. Zou et al., “60-nm-thick basic photonic components and Bragg gratings on the silicon-on-insulator platform,” Opt. Express, 23 20784 https://doi.org/10.1364/OE.23.020784 OPEXFF 1094-4087 (2015). Google Scholar

239. 

I. Krasnokutska et al., “Ultra-low loss photonic circuits in lithium niobate on insulator,” Opt. Express, 26 897 https://doi.org/10.1364/OE.26.000897 OPEXFF 1094-4087 (2018). Google Scholar

240. 

X. C. Ji et al., “On-chip tunable photonic delay line,” APL Photonics, 4 090803 https://doi.org/10.1063/1.5111164 (2019). Google Scholar

241. 

G. Vienne, Y. H. Li and L. M. Tong, “Microfiber resonator in polymer matrix,” IEICE Trans. Electron., E90-C 415 https://doi.org/10.1093/ietele/e90-c.2.415 IELEEJ 0916-8524 (2007). Google Scholar

242. 

F. Xu and G. Brambilla, “Preservation of micro-optical fibers by embedding,” Jpn. J. Appl. Phys., 47 6675 https://doi.org/10.1143/JJAP.47.6675 (2008). Google Scholar

243. 

L. Zhang et al., “Simple and cost-effective fabrication of two-dimensional plastic nanochannels from silica nanowire templates,” Microfluid. Nanofluid., 5 727 https://doi.org/10.1007/s10404-008-0314-4 (2008). Google Scholar

244. 

L. Zhang et al., “Ultra-sensitive microfibre absorption detection in a microfluidic chip,” Lab Chip, 11 3720 https://doi.org/10.1039/c1lc20519f LCAHAM 1473-0197 (2011). Google Scholar

245. 

S.-M. Chuo and L. A. Wang, “Propagation loss, degradation and protective coating of long drawn microfibers,” Opt. Commun., 284 2825 https://doi.org/10.1016/j.optcom.2011.02.018 OPCOB8 0030-4018 (2011). Google Scholar

246. 

J. Pan et al., “A multifunctional skin-like wearable optical sensor based on an optical micro-/nanofibre,” Nanoscale, 12 17538 https://doi.org/10.1039/D0NR03446K NANOHL 2040-3364 (2020). Google Scholar

247. 

L. Zhang et al., “Ultrasensitive skin-like wearable optical sensors based on glass micro/nanofibers,” Opto-Electron. Adv., 3 19002201 https://doi.org/10.29026/oea.2020.190022 (2020). Google Scholar

248. 

N. Lou et al., “Embedded optical micro/nano-fibers for stable devices,” Opt. Lett., 35 571 https://doi.org/10.1364/OL.35.000571 OPLEDP 0146-9592 (2010). Google Scholar

249. 

Y. Chen, F. Xu and Y. Q. Lu, “Teflon-coated microfiber resonator with weak temperature dependence,” Opt. Express, 19 22923 https://doi.org/10.1364/OE.19.022923 OPEXFF 1094-4087 (2011). Google Scholar

250. 

L. M. Xiao et al., “Stable low-loss optical nanofibres embedded in hydrophobic aerogel,” Opt. Express, 19 764 https://doi.org/10.1364/OE.19.000764 OPEXFF 1094-4087 (2011). Google Scholar

251. 

A. Sulaiman et al., “Microfiber Mach-Zehnder interferometer embedded in low index polymer,” Opt. Laser Technol., 44 1186 https://doi.org/10.1016/j.optlastec.2011.09.032 OLTCAS 0030-3992 (2012). Google Scholar

252. 

P. F. Wang et al., “Packaged, high-Q, microsphere-resonator-based add-drop filter,” Opt. Lett., 39 5208 https://doi.org/10.1364/OL.39.005208 OPLEDP 0146-9592 (2014). Google Scholar

253. 

L. H. Liu et al., “Fabrication of highly stable microfiber structures via high-substituted hydroxypropyl cellulose coating for device and sensor applications,” Opt. Lett., 40 1492 https://doi.org/10.1364/OL.40.001492 OPLEDP 0146-9592 (2015). Google Scholar

254. 

L. Jin et al., “Low-loss microfiber splicing based on low-index polymer coating,” IEEE Photon. Technol. Lett., 28 1181 https://doi.org/10.1109/LPT.2016.2535265 IPTLEL 1041-1135 (2016). Google Scholar

255. 

M. Bouhadida, P. Delaye and S. Lebrun, “Long-term optical transmittance measurements of silica nanofibers,” Opt. Commun., 500 127336 https://doi.org/10.1016/j.optcom.2021.127336 OPCOB8 0030-4018 (2021). Google Scholar

256. 

Y. Xue et al., “Ultrasensitive temperature sensor based on an isopropanol-sealed optical microfiber taper,” Opt. Lett., 38 1209 https://doi.org/10.1364/OL.38.001209 OPLEDP 0146-9592 (2013). Google Scholar

257. 

W. Jin et al., “Robust microfiber photonic microcells for sensor and device applications,” Opt. Express, 22 28132 https://doi.org/10.1364/OE.22.028132 OPEXFF 1094-4087 (2014). Google Scholar

258. 

L. Zhao et al., “Highly sensitive temperature sensor based on an isopropanol-sealed optical microfiber coupler,” Appl. Phys. Lett., 113 111901 https://doi.org/10.1063/1.5049558 APPLAB 0003-6951 (2018). Google Scholar

259. 

Y. Jung et al., “Adiabatic higher-order mode microfibers based on a logarithmic index profile,” Opt. Express, 28 19126 https://doi.org/10.1364/OE.394098 OPEXFF 1094-4087 (2020). Google Scholar

260. 

K. P. Nayak et al., “Optical nanofiber as an efficient tool for manipulating and probing atomic fluorescence,” Opt. Express, 15 5431 https://doi.org/10.1364/OE.15.005431 OPEXFF 1094-4087 (2007). Google Scholar

261. 

M. M. Lai, J. D. Franson and T. B. Pittman, “Transmission degradation and preservation for tapered optical fibers in rubidium vapor,” Appl. Opt., 52 2595 https://doi.org/10.1364/AO.52.002595 APOPAI 0003-6935 (2013). Google Scholar

262. 

T. B. Pittman, D. E. Jones and J. D. Franson, “Ultralow-power nonlinear optics using tapered optical fibers in metastable xenon,” Phys. Rev. A, 88 053804 https://doi.org/10.1103/PhysRevA.88.053804 (2013). Google Scholar

263. 

H. P. Lamsal, J. D. Franson and T. B. Pittman, “Transmission characteristics of optical nanofibers in metastable xenon,” Appl. Opt., 58 6470 https://doi.org/10.1364/AO.58.006470 APOPAI 0003-6935 (2019). Google Scholar

264. 

R. W. Boyd, Nonlinear Optics, third editionAcademic Press, (2008). Google Scholar

265. 

M. Kolesik, E. M. Wright and J. V. Moloney, “Simulation of femtosecond pulse propagation in sub-micron diameter tapered fibers,” Appl. Phys. B, 79 293 https://doi.org/10.1007/s00340-004-1551-1 (2004). Google Scholar

266. 

H. Lu et al., “Optimization of supercontinuum generation in air-silica nanowires,” J. Opt. Soc. Am. B, 27 904 https://doi.org/10.1364/JOSAB.27.000904 JOBPDE 0740-3224 (2010). Google Scholar

267. 

J. Lægsgaard, “Modeling of nonlinear propagation in fiber tapers,” J. Opt. Soc. Am. B, 29 3183 https://doi.org/10.1364/JOSAB.29.003183 JOBPDE 0740-3224 (2012). Google Scholar

268. 

J. Laegsgaard, “Theory of surface second-harmonic generation in silica nanowires,” J. Opt. Soc. Am. B, 27 1317 https://doi.org/10.1364/JOSAB.27.001317 JOBPDE 0740-3224 (2010). Google Scholar

269. 

S. Richard, “Second-harmonic generation in tapered optical fibers,” J. Opt. Soc. Am. B, 27 1504 https://doi.org/10.1364/JOSAB.27.001504 JOBPDE 0740-3224 (2010). Google Scholar

270. 

A. Azzoune, P. Delaye and G. Pauliat, “Modeling photon pair generation by second-order surface nonlinearity in silica nanofibers,” J. Opt. Soc. Am. B, 38 1057 https://doi.org/10.1364/JOSAB.418661 JOBPDE 0740-3224 (2021). Google Scholar

271. 

W. Luo et al., “Efficient surface second-harmonic generation in slot micro/nano-fibers,” Opt. Express, 21 11554 https://doi.org/10.1364/OE.21.011554 OPEXFF 1094-4087 (2013). Google Scholar

272. 

G. X. Wu et al., “Quasi-phase-matching method based on coupling compensation for surface second-harmonic generation in optical fiber nanowire coupler,” ACS Photonics, 5 3916 https://doi.org/10.1021/acsphotonics.8b00786 (2018). Google Scholar

273. 

V. Grubsky and J. Feinberg, “Phase-matched third-harmonic UV generation using low-order modes in a glass micro-fiber,” Opt. Commun., 274 447 https://doi.org/10.1016/j.optcom.2007.02.023 OPCOB8 0030-4018 (2007). Google Scholar

274. 

A. Coillet and P. Grelu, “Third-harmonic generation in optical microfibers: from silica experiments to highly nonlinear glass prospects,” Opt. Commun., 285 3493 https://doi.org/10.1016/j.optcom.2012.04.020 OPCOB8 0030-4018 (2012). Google Scholar

275. 

T. Lee et al., “Broadband third harmonic generation in tapered silica fibres,” Opt. Express, 20 8503 https://doi.org/10.1364/OE.20.008503 OPEXFF 1094-4087 (2012). Google Scholar

276. 

M. I. M. Abdul Khudus et al., “Effect of intrinsic surface roughness on the efficiency of intermodal phase matching in silica optical nanofibers,” Opt. Lett., 40 1318 https://doi.org/10.1364/OL.40.001318 OPLEDP 0146-9592 (2015). Google Scholar

277. 

X. J. Jiang et al., “Fundamental-mode third harmonic generation in microfibers by pulse-induced quasi-phase matching,” Opt. Express, 25 22626 https://doi.org/10.1364/OE.25.022626 OPEXFF 1094-4087 (2017). Google Scholar

278. 

X. J. Jiang et al., “Optimized microfiber-based third-harmonic generation with adaptive control of phase mismatch,” Opt. Lett., 43 2728 https://doi.org/10.1364/OL.43.002728 OPLEDP 0146-9592 (2018). Google Scholar

279. 

X. J. Jiang et al., “Enhanced UV third-harmonic generation in microfibers by controlling nonlinear phase modulations,” Opt. Lett., 44 4191 https://doi.org/10.1364/OL.44.004191 OPLEDP 0146-9592 (2019). Google Scholar

280. 

Z. Hao et al., “Strain-controlled phase matching of optical harmonic generation in microfibers,” Phys. Rev. Appl., 19 L031002 https://doi.org/10.1103/PhysRevApplied.19.L031002 PRAHB2 2331-7019 (2023). Google Scholar

281. 

J. H. Kim et al., “Photon-pair source working in a silicon-based detector wavelength range using tapered micro/nanofibers,” Opt. Lett., 44 447 https://doi.org/10.1364/OL.44.000447 OPLEDP 0146-9592 (2019). Google Scholar

282. 

M. I. M. Abdul Khudus et al., “Phase matched parametric amplification via four-wave mixing in optical microfibers,” Opt. Lett., 41 761 https://doi.org/10.1364/OL.41.000761 OPLEDP 0146-9592 (2016). Google Scholar

283. 

R. Vacher and L. Boyer, “Brillouin scattering: a tool for the measurement of elastic and photoelastic constants,” Phys. Rev. B, 6 639 https://doi.org/10.1103/PhysRevB.6.639 (1972). Google Scholar

284. 

A. Kobyakov, M. Sauer and D. Chowdhury, “Stimulated Brillouin scattering in optical fibers,” Adv. Opt. Photonics, 2 1 https://doi.org/10.1364/AOP.2.000001 AOPAC7 1943-8206 (2009). Google Scholar

285. 

A. Godet et al., “Micronewton nanofiber force sensor using Brillouin scattering,” Opt. Express, 30 815 https://doi.org/10.1364/OE.443594 OPEXFF 1094-4087 (2022). Google Scholar

286. 

O. Florez et al., “Brillouin scattering self-cancellation,” Nat. Commun., 7 11759 https://doi.org/10.1038/ncomms11759 NCAOBW 2041-1723 (2016). Google Scholar

287. 

D. M. Chow et al., “Local activation of surface and hybrid acoustic waves in optical microwires,” Opt. Lett., 43 1487 https://doi.org/10.1364/OL.43.001487 OPLEDP 0146-9592 (2018). Google Scholar

288. 

M. Cao et al., “Inter-mode forward Brillouin scattering in nanofibers,” J. Lightwave Technol., 38 6911 https://doi.org/10.1109/JLT.2020.3017192 JLTEDG 0733-8724 (2020). Google Scholar

289. 

M. Cao et al., “Influence of optical mode polarization state on the Brillouin gain spectrum in optical microfibers,” J. Opt. Soc. Am. B, 39 1443 https://doi.org/10.1364/JOSAB.451021 JOBPDE 0740-3224 (2022). Google Scholar

290. 

M. Bouhadida et al., “Highly efficient and reproducible evanescent Raman converters based on a silica nanofiber immersed in a liquid,” Appl. Phys. B, 125 228 https://doi.org/10.1007/s00340-019-7340-7 (2019). Google Scholar

291. 

Y. Qi et al., “Nanofiber enhanced stimulated Raman spectroscopy for ultra-fast, ultra-sensitive hydrogen detection with ultra-wide dynamic range,” Optica, 6 570 https://doi.org/10.1364/OPTICA.6.000570 (2019). Google Scholar

292. 

R. R. Alfano and S. L. Shapiro, “Emission in the region 4000 to 7000 Å via four-photon coupling in glass,” Phys. Rev. Lett., 24 584 https://doi.org/10.1103/PhysRevLett.24.584 PRLTAO 0031-9007 (1970). Google Scholar

293. 

C. Lin and R. H. Stolen, “New nanosecond continuum for excited-state spectroscopy,” Appl. Phys. Lett., 28 216 https://doi.org/10.1063/1.88702 APPLAB 0003-6951 (1976). Google Scholar

294. 

M. A. Foster, K. D. Moll and A. L. Gaeta, “Optimal waveguide dimensions for nonlinear interactions,” Opt. Express, 12 2880 https://doi.org/10.1364/OPEX.12.002880 OPEXFF 1094-4087 (2004). Google Scholar

295. 

R. R. Gattass et al., “Supercontinuum generation in submicrometer diameter silica fibers,” Opt. Express, 14 9408 https://doi.org/10.1364/OE.14.009408 OPEXFF 1094-4087 (2006). Google Scholar

296. 

D. I. Yeom et al., “Low-threshold supercontinuum generation in highly nonlinear chalcogenide nanowires,” Opt. Lett., 33 660 https://doi.org/10.1364/OL.33.000660 OPLEDP 0146-9592 (2008). Google Scholar

297. 

D. D. Hudson et al., “Highly nonlinear chalcogenide glass micro/nanofiber devices: design, theory, and octave-spanning spectral generation,” Opt. Commun., 285 4660 https://doi.org/10.1016/j.optcom.2012.05.002 OPCOB8 0030-4018 (2012). Google Scholar

298. 

A. Al-kadry et al., “Broadband supercontinuum generation in As2Se3 chalcogenide wires by avoiding the two-photon absorption effects,” Opt. Lett., 38 1185 https://doi.org/10.1364/OL.38.001185 OPLEDP 0146-9592 (2013). Google Scholar

299. 

C. W. Rudy et al., “Octave-spanning supercontinuum generation in in situ tapered As2Se3 fiber pumped by a thulium-doped fiber laser,” Opt. Lett., 38 2865 https://doi.org/10.1364/OL.38.002865 OPLEDP 0146-9592 (2013). Google Scholar

300. 

A. Marandi et al., “Mid-infrared supercontinuum generation in tapered chalcogenide fiber for producing octave-spanning frequency comb around 3 µm,” Opt. Express, 20 24218 https://doi.org/10.1364/OE.20.024218 OPEXFF 1094-4087 (2012). Google Scholar

301. 

A. Al-Kadry et al., “Two octaves mid-infrared supercontinuum generation in As2Se3 microwires,” Opt. Express, 22 31131 https://doi.org/10.1364/OE.22.031131 OPEXFF 1094-4087 (2014). Google Scholar

302. 

Y. N. Sun et al., “Fabrication and characterization of multimaterial chalcogenide glass fiber tapers with high numerical apertures,” Opt. Express, 23 23472 https://doi.org/10.1364/OE.23.023472 OPEXFF 1094-4087 (2015). Google Scholar

303. 

D. D. Hudson et al., “Toward all-fiber supercontinuum spanning the mid-infrared,” Optica, 4 1163 https://doi.org/10.1364/OPTICA.4.001163 (2017). Google Scholar

304. 

Y. Y. Wang et al., “14–72 μm broadband supercontinuum generation in an As-S chalcogenide tapered fiber pumped in the normal dispersion regime,” Opt. Lett., 42 3458 https://doi.org/10.1364/OL.42.003458 OPLEDP 0146-9592 (2017). Google Scholar

305. 

S. Gao and X. Y. Bao, “Chalcogenide taper and its nonlinear effects and sensing applications,” iScience, 23 100802 https://doi.org/10.1016/j.isci.2019.100802 (2020). Google Scholar

306. 

Y. Y. Wang and S. X. Dai, “Mid-infrared supercontinuum generation in chalcogenide glass fibers: a brief review,” PhotoniX, 2 9 https://doi.org/10.1186/s43074-021-00031-3 (2021). Google Scholar

307. 

C. M. Lieber, “One-dimensional nanostructures: chemistry, physics &amp; applications,” Solid State Commun., 107 607 https://doi.org/10.1016/S0038-1098(98)00209-9 SSCOA4 0038-1098 (1998). Google Scholar

308. 

M. Naraghi et al., “Novel method for mechanical characterization of polymeric nanofibers,” Rev. Sci. Instrum., 78 085108 https://doi.org/10.1063/1.2771092 RSINAK 0034-6748 (2007). Google Scholar

309. 

M. Dao et al., “Toward a quantitative understanding of mechanical behavior of nanocrystalline metals,” Acta Mater., 55 4041 https://doi.org/10.1016/j.actamat.2007.01.038 ACMAFD 1359-6454 (2007). Google Scholar

310. 

T. Zhu and J. Li, “Ultra-strength materials,” Prog. Mater. Sci., 55 710 https://doi.org/10.1016/j.pmatsci.2010.04.001 PRMSAQ 0079-6425 (2010). Google Scholar

311. 

L. P. Dávila, V. J. Leppert and E. M. Bringa, “The mechanical behavior and nanostructure of silica nanowires via simulations,” Scr. Mater., 60 843 https://doi.org/10.1016/j.scriptamat.2008.12.057 SCMAF7 1359-6462 (2009). Google Scholar

312. 

F. L. Yuan and L. P. Huang, “Size-dependent elasticity of amorphous silica nanowire: a molecular dynamics study,” Appl. Phys. Lett., 103 201905 https://doi.org/10.1063/1.4830038 APPLAB 0003-6951 (2013). Google Scholar

313. 

S. Romeis et al., “Correlation of enhanced strength and internal structure for heat-treated submicron stober silica particles,” Part. Part. Syst. Charact., 31 664 https://doi.org/10.1002/ppsc.201300306 PPCHEZ 0934-0866 (2014). Google Scholar

314. 

X. Q. Chen et al., “Mechanical resonance of quartz microfibers and boundary condition effects,” J. Appl. Phys., 95 4823 https://doi.org/10.1063/1.1697635 JAPIAU 0021-8979 (2004). Google Scholar

315. 

E. C. Silva et al., “Size effects on the stiffness of silica nanowires,” Small, 2 239 https://doi.org/10.1002/smll.200500311 SMALBC 1613-6810 (2006). Google Scholar

316. 

H. Ni, X. D. Li and H. S. Gao, “Elastic modulus of amorphous SiO2 nanowires,” Appl. Phys. Lett., 88 043108 https://doi.org/10.1063/1.2165275 APPLAB 0003-6951 (2006). Google Scholar

317. 

Z. W. Ma et al., “Tensile strength and failure behavior of bare single mode fibers,” Opt. Fiber Technol., 52 101966 https://doi.org/10.1016/j.yofte.2019.101966 1068-5200 (2019). Google Scholar

318. 

M. Guerette and L. P. Huang, “Nonlinear elasticity of silica fibers studied by in-situ Brillouin light scattering in two-point bend test,” Am. Ceram. Soc. Bull., 94 40 ACSBA7 0002-7812 (2015). Google Scholar

319. 

M. Guerette et al., “Nonlinear elasticity of silica glass,” J. Am. Ceram. Soc., 99 841 https://doi.org/10.1111/jace.14043 JACTAW 0002-7820 (2015). Google Scholar

320. 

T. Fett et al., “Effect of damage by hydroxyl generation on strength of silica fibers,” J. Am. Ceram. Soc., 101 2724 https://doi.org/10.1111/jace.15508 JACTAW 0002-7820 (2018). Google Scholar

321. 

L. M. Tong et al., “Assembly of silica nanowires on silica aerogels for microphotonic devices,” Nano Lett., 5 259 https://doi.org/10.1021/nl0481977 NALEFD 1530-6984 (2005). Google Scholar

322. 

K. Zheng et al., “Electron-beam-assisted superplastic shaping of nanoscale amorphous silica,” Nat. Commun., 1 24 https://doi.org/10.1038/ncomms1021 NCAOBW 2041-1723 (2010). Google Scholar

323. 

J. H. Luo et al., “Size-dependent brittle-to-ductile transition in silica glass nanofibers,” Nano Lett., 16 105 https://doi.org/10.1021/acs.nanolett.5b03070 NALEFD 1530-6984 (2016). Google Scholar

324. 

X. Guo et al., “Direct coupling of plasmonic and photonic nanowires for hybrid nanophotonic components and circuits,” Nano Lett., 9 4515 https://doi.org/10.1021/nl902860d NALEFD 1530-6984 (2009). Google Scholar

325. 

X. Y. Li et al., “All-fiber hybrid photon-plasmon circuits: integrating nanowire plasmonics with fiber optics,” Opt. Express, 21 15698 https://doi.org/10.1364/OE.21.015698 OPEXFF 1094-4087 (2013). Google Scholar

326. 

H. K. Yu et al., “Single nanowire optical correlator,” Nano Lett., 14 3487 https://doi.org/10.1021/nl5010477 NALEFD 1530-6984 (2014). Google Scholar

327. 

C. G. Xin et al., “Single CdTe nanowire optical correlator for femtojoule pulses,” Nano Lett., 16 4807 https://doi.org/10.1021/acs.nanolett.6b00893 NALEFD 1530-6984 (2016). Google Scholar

328. 

Z. X. Shi et al., “Miniature optical correlator in a single-nanowire Sagnac loop,” ACS Photonics, 7 3264 https://doi.org/10.1021/acsphotonics.0c01417 (2020). Google Scholar

329. 

J. B. Zhang et al., “Single microwire optical autocorrelator at 2-μm wavelength,” IEEE Photon. Technol. Lett., 34 207 https://doi.org/10.1109/LPT.2022.3146878 IPTLEL 1041-1135 (2022). Google Scholar

330. 

Z. X. Shi et al., “Single-nanowire thermo-optic modulator based on a Varshni shift,” ACS Photonics, 7 2571 https://doi.org/10.1021/acsphotonics.0c00917 (2020). Google Scholar

331. 

Y. X. Xu et al., “Microfiber coupled superconducting nanowire single-photon detectors,” Opt. Commun., 405 48 https://doi.org/10.1016/j.optcom.2017.07.087 OPCOB8 0030-4018 (2017). Google Scholar

332. 

L. X. You et al., “Microfiber-coupled superconducting nanowire single-photon detector for near-infrared wavelengths,” Opt. Express, 25 31221 https://doi.org/10.1364/OE.25.031221 OPEXFF 1094-4087 (2017). Google Scholar

333. 

X. T. Hou et al., “Ultra-broadband microfiber-coupled superconducting single-photon detector,” Opt. Express, 27 25241 https://doi.org/10.1364/OE.27.025241 OPEXFF 1094-4087 (2019). Google Scholar

334. 

Z. Y. Zhang et al., “Subwavelength-diameter silica wire for light in-coupling to silicon-based waveguide,” Chin. Opt. Lett., 5 577 https://doi.org/https://opg.optica.org/col/abstract.cfm?URI=col-5-10-577 CJOEE3 1671-7694 (2007). Google Scholar

335. 

X. W. Shen et al., “Highly efficient fiber-to-chip evanescent coupling based on subwavelength-diameter optical fibers,” Chin. Opt. Lett., 9 050604 https://doi.org/10.3788/COL201109.050604 CJOEE3 1671-7694 (2011). Google Scholar

336. 

B. G. Chen et al., “Flexible integration of free-standing nanowires into silicon photonics,” Nat. Commun., 8 20 https://doi.org/10.1038/s41467-017-00038-0 NCAOBW 2041-1723 (2017). Google Scholar

337. 

S. Khan et al., “Low-loss, high-bandwidth fiber-to-chip coupling using capped adiabatic tapered fibers,” APL Photonics, 5 056101 https://doi.org/10.1063/1.5145105 (2020). Google Scholar

338. 

Y. Y. Jin et al., “Efficient fiber-to-chip interface via an intermediated CdS nanowire,” Laser Photon. Rev., 17 2200919 https://doi.org/10.1002/lpor.202200919 (2023). Google Scholar

339. 

G. Son et al., “High-efficiency broadband light coupling between optical fibers and photonic integrated circuits,” Nanophotonics, 7 1845 https://doi.org/10.1515/nanoph-2018-0075 (2018). Google Scholar

340. 

W. L. Wu et al., “Efficient mode conversion from a standard single-mode fiber to a subwavelength-diameter microfiber,” Nanomaterials, 13 3003 https://doi.org/10.3390/nano13233003 (2023). Google Scholar

341. 

F. N. Xia et al., “Two-dimensional material nanophotonics,” Nat. Photonics, 8 899 https://doi.org/10.1038/nphoton.2014.271 NPAHBY 1749-4885 (2014). Google Scholar

342. 

B. Guo et al., “2D layered materials: synthesis, nonlinear optical properties, and device applications,” Laser Photon. Rev., 13 1800327 https://doi.org/10.1002/lpor.201800327 (2019). Google Scholar

343. 

J. Z. Wang et al., “Evanescent-light deposition of graphene onto tapered fibers for passive Q-switch and mode-locker,” IEEE Photonics J., 4 1295 https://doi.org/10.1109/JPHOT.2012.2208736 (2012). Google Scholar

344. 

X. Q. Wu et al., “Effective transfer of micron-size graphene to microfibers for photonic applications,” Carbon, 96 1114 https://doi.org/10.1016/j.carbon.2015.10.069 CRBNAH 0008-6223 (2016). Google Scholar

345. 

W. Li et al., “Ultrafast all-optical graphene modulator,” Nano Lett., 14 955 https://doi.org/10.1021/nl404356t NALEFD 1530-6984 (2014). Google Scholar

346. 

J. H. Chen et al., “An all-optical modulator based on a stereo graphene–microfiber structure,” Light Sci. Appl., 4 e360 https://doi.org/10.1038/lsa.2015.133 (2015). Google Scholar

347. 

S. L. Yu et al., “All-optical graphene modulator based on optical Kerr phase shift,” Optica, 3 541 https://doi.org/10.1364/OPTICA.3.000541 (2016). Google Scholar

348. 

M. Liu et al., “Recent progress on applications of 2D material-decorated microfiber photonic devices in pulse shaping and all-optical signal processing,” Nanophotonics, 9 2641 https://doi.org/10.1515/nanoph-2019-0564 (2020). Google Scholar

349. 

B. Q. Jiang et al., “High-efficiency second-order nonlinear processes in an optical microfibre assisted by few-layer GaSe,” Light Sci. Appl., 9 63 https://doi.org/10.1038/s41377-020-0304-1 (2020). Google Scholar

350. 

X. M. Liu et al., “Graphene-clad microfibre saturable absorber for ultrafast fibre lasers,” Sci. Rep., 6 26024 https://doi.org/10.1038/srep26024 SRCEC3 2045-2322 (2016). Google Scholar

351. 

S. H. K. Yap et al., “Two-dimensional MoS2 nanosheet-functionalized optical microfiber for room-temperature volatile organic compound detection,” ACS Appl. Nano Mater., 4 13440 https://doi.org/10.1021/acsanm.1c02937 (2021). Google Scholar

352. 

A. W. Schell et al., “Coupling quantum emitters in 2D materials with tapered fibers,” ACS Photonics, 4 761 https://doi.org/10.1021/acsphotonics.7b00025 (2017). Google Scholar

353. 

J. H. Chen et al., “Tunable and enhanced light emission in hybrid WS2-optical-fiber-nanowire structures,” Light Sci. Appl., 8 8 https://doi.org/10.1038/s41377-018-0115-9 (2019). Google Scholar

354. 

J. L. Xiao et al., “Optical fibre taper-enabled waveguide photoactuators,” Nat. Commun., 13 363 https://doi.org/10.1038/s41467-022-28021-4 NCAOBW 2041-1723 (2022). Google Scholar

355. 

S. M. Spillane et al., “Ideality in a fiber-taper-coupled microresonator system for application to cavity quantum electrodynamics,” Phys. Rev. Lett., 91 043902 https://doi.org/10.1103/PhysRevLett.91.043902 PRLTAO 0031-9007 (2003). Google Scholar

356. 

S. I. Shopova et al., “Plasmonic enhancement of a whispering-gallery-mode biosensor for single nanoparticle detection,” Appl. Phys. Lett., 98 243104 https://doi.org/10.1063/1.3599584 APPLAB 0003-6951 (2011). Google Scholar

357. 

Z. Shen et al., “Experimental realization of optomechanically induced non-reciprocity,” Nat. Photonics, 10 657 https://doi.org/10.1038/nphoton.2016.161 NPAHBY 1749-4885 (2016). Google Scholar

358. 

X. Y. Zhang et al., “Symmetry-breaking-induced nonlinear optics at a microcavity surface,” Nat. Photonics, 13 21 https://doi.org/10.1038/s41566-018-0297-y NPAHBY 1749-4885 (2018). Google Scholar

359. 

B. Jiang et al., “Room-temperature continuous-wave upconversion white microlaser using a rare-earth-doped microcavity,” ACS Photonics, 9 2956 https://doi.org/10.1021/acsphotonics.2c00379 (2022). Google Scholar

360. 

Y. C. Dong, X. Y. Jin and K. Y. Wang, “Packaged and robust microcavity device based on a microcylinder-taper coupling system,” Appl. Opt., 54 4016 https://doi.org/10.1364/AO.54.004016 APOPAI 0003-6935 (2015). Google Scholar

361. 

X. Y. Lu et al., “Heralding single photons from a high-Q silicon microdisk,” Optica, 3 1331 https://doi.org/10.1364/OPTICA.3.001331 (2016). Google Scholar

362. 

S. B. Gorajoobi, G. S. Murugan and M. N. Zervas, “Design of rare-earth-doped microbottle lasers,” Opt. Express, 26 26339 https://doi.org/10.1364/OE.26.026339 OPEXFF 1094-4087 (2018). Google Scholar

363. 

D. Farnesi et al., “Long period grating-based fiber coupler to whispering gallery mode resonators,” Opt. Lett., 39 6525 https://doi.org/10.1364/OL.39.006525 OPLEDP 0146-9592 (2014). Google Scholar

364. 

D. Farnesi et al., “Quasi-distributed and wavelength selective addressing of optical micro-resonators based on long period fiber gratings,” Opt. Express, 23 21175 https://doi.org/10.1364/OE.23.021175 OPEXFF 1094-4087 (2015). Google Scholar

365. 

Y. G. Ma et al., “Direct measurement of propagation losses in silver nanowires,” Opt. Lett., 35 1160 https://doi.org/10.1364/OL.35.001160 OPLEDP 0146-9592 (2010). Google Scholar

366. 

M. Sumetsky et al., “Optical microfiber loop resonator,” Appl. Phys. Lett., 86 161108 https://doi.org/10.1063/1.1906317 APPLAB 0003-6951 (2005). Google Scholar

367. 

M. Sumetsky et al., “The microfiber loop resonator: theory, experiment, and application,” J. Lightwave Technol., 24 242 https://doi.org/10.1109/JLT.2005.861127 JLTEDG 0733-8724 (2006). Google Scholar

368. 

X. Guo et al., “Demonstration of critical coupling in microfiber loops wrapped around a copper rod,” Appl. Phys. Lett., 91 073512 https://doi.org/10.1063/1.2771526 APPLAB 0003-6951 (2007). Google Scholar

369. 

S. S. Wang et al., “All-fiber Fabry-Perot resonators based on microfiber Sagnac loop mirrors,” Opt. Lett., 34 253 https://doi.org/10.1364/OL.34.000253 OPLEDP 0146-9592 (2009). Google Scholar

370. 

X. S. Jiang et al., “Demonstration of optical microfiber knot resonators,” Appl. Phys. Lett., 88 223501 https://doi.org/10.1063/1.2207986 APPLAB 0003-6951 (2006). Google Scholar

371. 

L. M. Xiao and T. A. Birks, “High finesse microfiber knot resonators made from double-ended tapered fibers,” Opt. Lett., 36 1098 https://doi.org/10.1364/OL.36.001098 OPLEDP 0146-9592 (2011). Google Scholar

372. 

J. M. De Freitas, T. A. Birks and M. Rollings, “Optical micro-knot resonator hydrophone,” Opt. Express, 23 5850 https://doi.org/10.1364/OE.23.005850 OPEXFF 1094-4087 (2015). Google Scholar

373. 

J. H. Li et al., “Versatile hybrid plasmonic microfiber knot resonator,” Opt. Lett., 42 3395 https://doi.org/10.1364/OL.42.003395 OPLEDP 0146-9592 (2017). Google Scholar

374. 

Z. X. Ding et al., “All-fiber ultrafast laser generating gigahertz-rate pulses based on a hybrid plasmonic microfiber resonator,” Adv. Photonics, 2 026002 https://doi.org/10.1117/1.AP.2.2.026002 AOPAC7 1943-8206 (2020). Google Scholar

375. 

P. Pal and W. H. Knox, “Low loss fusion splicing of micron scale silica fibers,” Opt. Express, 16 11568 https://doi.org/10.1364/OE.16.011568 OPEXFF 1094-4087 (2008). Google Scholar

376. 

P. Pal and W. H. Knox, “Fabrication and characterization of fused microfiber resonators,” IEEE Photon. Technol. Lett., 21 766 https://doi.org/10.1109/LPT.2009.2017506 IPTLEL 1041-1135 (2009). Google Scholar

377. 

P. Wang et al., “Fusion spliced microfiber closed-loop resonators,” IEEE Photon. Technol. Lett., 22 1075 https://doi.org/10.1109/LPT.2010.2049646 IPTLEL 1041-1135 (2010). Google Scholar

378. 

W. Li et al., “Fusion splicing soft glass microfibers for photonic devices,” IEEE Photon. Technol. Lett., 23 831 https://doi.org/10.1109/LPT.2011.2140369 IPTLEL 1041-1135 (2011). Google Scholar

379. 

M. Sumetsky, “Optical fiber microcoil resonators,” Opt. Express, 12 2303 https://doi.org/10.1364/OPEX.12.002303 OPEXFF 1094-4087 (2004). Google Scholar

380. 

M. Sumetsky, Y. Dulashko and M. Fishteyn, “Demonstration of a multi-turn microfiber coil resonator,” in Optical Fiber Communication Conference and Exposition and The National Fiber Optic Engineers Conference, (2007). Google Scholar

381. 

F. Xu and G. Brambilla, “Manufacture of 3-D microfiber coil resonators,” IEEE Photon. Technol. Lett., 19 1481 https://doi.org/10.1109/LPT.2007.903762 IPTLEL 1041-1135 (2007). Google Scholar

382. 

F. Xu, P. Horak and G. Brambilla, “Optimized design of microcoil resonators,” J. Lightwave Technol., 25 1561 https://doi.org/10.1109/JLT.2007.895546 JLTEDG 0733-8724 (2007). Google Scholar

383. 

F. Xu, P. Horak and G. Brambilla, “Conical and biconical ultra-high-Q optical-fiber nanowire microcoil resonator,” Appl. Opt., 46 570 https://doi.org/10.1364/AO.46.000570 APOPAI 0003-6935 (2007). Google Scholar

384. 

Y. Jung et al., “Embedded optical microfiber coil resonator with enhanced high-Q,” IEEE Photon. Technol. Lett., 22 1638 https://doi.org/10.1109/LPT.2010.2076332 IPTLEL 1041-1135 (2010). Google Scholar

385. 

Y. Wu et al., “Miniature interferometric humidity sensors based on silica/polymer microfiber knot resonators,” Sens. Actuator B-Chem., 155 258 https://doi.org/10.1016/j.snb.2010.12.030 (2011). Google Scholar

386. 

A. D. D. Le and Y. G. Han, “Relative humidity sensor based on a few-mode microfiber knot resonator by mitigating the group index difference of a few-mode microfiber,” J. Lightwave Technol., 36 904 https://doi.org/10.1109/JLT.2017.2756639 JLTEDG 0733-8724 (2018). Google Scholar

387. 

Y. G. Han, “Relative humidity sensors based on microfiber knot resonators-a review,” Sensors, 19 5196 https://doi.org/10.3390/s19235196 SNSRES 0746-9462 (2019). Google Scholar

388. 

F. Xu and G. Brambilla, “Embedding optical microfiber coil resonators in Teflon,” Opt. Lett., 32 2164 https://doi.org/10.1364/OL.32.002164 OPLEDP 0146-9592 (2007). Google Scholar

389. 

F. Xu, P. Horak and G. Brambilla, “Optical microfiber coil resonator refractometric sensor,” Opt. Express, 15 7888 https://doi.org/10.1364/OE.15.007888 OPEXFF 1094-4087 (2007). Google Scholar

390. 

X. S. Jiang et al., “All-fiber add-drop filters based on microfiber knot resonators,” Opt. Lett., 32 1710 https://doi.org/10.1364/OL.32.001710 OPLEDP 0146-9592 (2007). Google Scholar

391. 

X. S. Jiang et al., “Demonstration of microfiber knot laser,” Appl. Phys. Lett., 89 143513 https://doi.org/10.1063/1.2359439 APPLAB 0003-6951 (2006). Google Scholar

392. 

X. S. Jiang et al., “Microfiber knot dye laser based on the evanescent-wave-coupled gain,” Appl. Phys. Lett., 90 233501 https://doi.org/10.1063/1.2746935 APPLAB 0003-6951 (2007). Google Scholar

393. 

Q. Yang et al., “Hybrid structure laser based on semiconductor nanowires and a silica microfiber knot cavity,” Appl. Phys. Lett., 94 101108 https://doi.org/10.1063/1.3093821 APPLAB 0003-6951 (2009). Google Scholar

394. 

Z. L. Xu et al., “Light velocity control in monolithic microfiber bridged ring resonator,” Optica, 4 945 https://doi.org/10.1364/OPTICA.4.000945 (2017). Google Scholar

395. 

Y. H. Li and L. M. Tong, “Mach-Zehnder interferometers assembled with optical microfibers or nanofibers,” Opt. Lett., 33 303 https://doi.org/10.1364/OL.33.000303 OPLEDP 0146-9592 (2008). Google Scholar

396. 

J. H. Wo et al., “Refractive index sensor using microfiber-based Mach-Zehnder interferometer,” Opt. Lett., 37 67 https://doi.org/10.1364/OL.37.000067 OPLEDP 0146-9592 (2012). Google Scholar

397. 

Y. H. Chen et al., “Hybrid Mach-Zehnder interferometer and knot resonator based on silica microfibers,” Opt. Commun., 283 2953 https://doi.org/10.1016/j.optcom.2010.03.052 OPCOB8 0030-4018 (2010). Google Scholar

398. 

X. Fang, C. R. Liao and D. N. Wang, “Femtosecond laser fabricated fiber Bragg grating in microfiber for refractive index sensing,” Opt. Lett., 35 1007 https://doi.org/10.1364/OL.35.001007 OPLEDP 0146-9592 (2010). Google Scholar

399. 

D. W. Cai et al., “Mid-infrared microfiber Bragg gratings,” Opt. Lett., 45 6114 https://doi.org/10.1364/OL.403893 OPLEDP 0146-9592 (2020). Google Scholar

400. 

J. L. Kou et al., “Demonstration of a compact temperature sensor based on first-order Bragg grating in a tapered fiber probe,” Opt. Express, 19 18452 https://doi.org/10.1364/OE.19.018452 OPEXFF 1094-4087 (2011). Google Scholar

401. 

Y. X. Liu et al., “Compact microfiber Bragg gratings with high-index contrast,” Opt. Lett., 36 3115 https://doi.org/10.1364/OL.36.003115 OPLEDP 0146-9592 (2011). Google Scholar

402. 

P. Romagnoli et al., “Fabrication of optical nanofibre-based cavities using focussed ion-beam milling: a review,” Appl. Phys. B, 126 111 https://doi.org/10.1007/s00340-020-07456-x (2020). Google Scholar

403. 

W. Liang et al., “Highly sensitive fiber Bragg grating refractive index sensors,” Appl. Phys. Lett., 86 151122 https://doi.org/10.1063/1.1904716 APPLAB 0003-6951 (2005). Google Scholar

404. 

R. Gao, J. Ye and X. Xin, “Directional acoustic signal measurement based on the asymmetrical temperature distribution of the parallel microfiber array,” Opt. Express, 27 34113 https://doi.org/10.1364/OE.27.034113 OPEXFF 1094-4087 (2019). Google Scholar

405. 

S. M. Lee, S. S. Saini and M. Y. Jeong, “Simultaneous measurement of refractive index, temperature, and strain using etched-core fiber Bragg grating sensors,” IEEE Photon. Technol. Lett., 22 1431 https://doi.org/10.1109/LPT.2010.2057416 IPTLEL 1041-1135 (2010). Google Scholar

406. 

H. F. Xuan, W. Jin and M. Zhang, “CO2 laser induced long period gratings in optical microfibers,” Opt. Express, 17 21882 https://doi.org/10.1364/OE.17.021882 OPEXFF 1094-4087 (2009). Google Scholar

407. 

Z. Y. Xu, Y. H. Li and L. J. Wang, “Long-period grating inscription on polymer functionalized optical microfibers and its applications in optical sensing,” Photonics Res., 4 45 https://doi.org/10.1364/PRJ.4.000045 (2016). Google Scholar

408. 

Y. Ran et al., “Type IIa Bragg gratings formed in microfibers,” Opt. Lett., 40 3802 https://doi.org/10.1364/OL.40.003802 OPLEDP 0146-9592 (2015). Google Scholar

409. 

P. Xiao et al., “Spectral tuning of the diameter-dependent-chirped Bragg gratings written in microfibers,” Opt. Express, 24 29750 https://doi.org/10.1364/OE.24.029749 OPEXFF 1094-4087 (2016). Google Scholar

410. 

F. Xu et al., “A microfiber Bragg grating based on a microstructured rod: a proposal,” IEEE Photon. Technol. Lett., 22 218 https://doi.org/10.1109/LPT.2009.2037515 IPTLEL 1041-1135 (2010). Google Scholar

411. 

D. Monzon-Hernandez et al., “Optical microfibers decorated with PdAu nanoparticles for fast hydrogen sensing,” Sens. Actuator B-Chem., 151 219 https://doi.org/10.1016/j.snb.2010.09.018 (2010). Google Scholar

412. 

J. Li et al., “Refractive index sensor based on silica microfiber doped with Ag microparticles,” Opt. Laser Technol., 94 40 https://doi.org/10.1016/j.optlastec.2017.03.024 OLTCAS 0030-3992 (2017). Google Scholar

413. 

P. Wang et al., “Single-band 2-nm-line-width plasmon resonance in a strongly coupled Au nanorod,” Nano Lett., 15 7581 https://doi.org/10.1021/acs.nanolett.5b03330 NALEFD 1530-6984 (2015). Google Scholar

414. 

N. Zhou et al., “Au nanorod-coupled microfiber optical humidity sensors,” Opt. Express, 27 8180 https://doi.org/10.1364/OE.27.008180 OPEXFF 1094-4087 (2019). Google Scholar

415. 

F. X. Gu et al., “Free-space coupling of nanoantennas and whispering-gallery microcavities with narrowed linewidth and enhanced sensitivity,” Laser Photon. Rev., 9 682 https://doi.org/10.1002/lpor.201500137 (2015). Google Scholar

416. 

Y. Y. Jin et al., “Strong coupling of a plasmonic nanoparticle to a semiconductor nanowire,” Nanophotonics, 10 2875 https://doi.org/10.1515/nanoph-2021-0214 (2021). Google Scholar

417. 

N. Zhou et al., “Strong mode coupling-enabled hybrid photon-plasmon laser with a microfiber-coupled nanorod,” Sci. Adv., 8 eabn2026 https://doi.org/10.1126/sciadv.abn2026 STAMCV 1468-6996 (2022). Google Scholar

418. 

Q. Ai et al., “Ultranarrow second-harmonic resonances in hybrid plasmon-fiber cavities,” Nano Lett., 18 5576 https://doi.org/10.1021/acs.nanolett.8b02005 NALEFD 1530-6984 (2018). Google Scholar

419. 

Q. Ai et al., “Giant second harmonic generation enhancement in a high-Q doubly resonant hybrid plasmon–fiber cavity system,” ACS Nano, 15 19409 https://doi.org/10.1021/acsnano.1c05970 ANCAC3 1936-0851 (2021). Google Scholar

420. 

Q. Ai et al., “Multiphoton photoluminescence in hybrid plasmon–fiber cavities with Au and Au@Pd nanobipyramids: two-photon versus four-photon processes and rapid quenching,” ACS Photonics, 8 2088 https://doi.org/10.1021/acsphotonics.1c00470 (2021). Google Scholar

421. 

F. Chiavaioli et al., “Femtomolar detection by nanocoated fiber label-free biosensors,” ACS Sens., 3 936 https://doi.org/10.1021/acssensors.7b00918 (2018). Google Scholar

422. 

Y. K. Liu et al., “Plasmonic fiber-optic photothermal anemometers with carbon nanotube coatings,” J. Lightwave Technol., 37 3373 https://doi.org/10.1109/JLT.2019.2916572 JLTEDG 0733-8724 (2019). Google Scholar

423. 

A. Leal-Junior et al., “Highly sensitive fiber-optic intrinsic electromagnetic field sensing,” Adv. Photon. Res., 2 2000078 https://doi.org/10.1002/adpr.202000078 (2020). Google Scholar

424. 

F. Chiavaioli and D. Janner, “Fiber optic sensing with lossy mode resonances: applications and perspectives,” J. Lightwave Technol., 39 3855 https://doi.org/10.1109/JLT.2021.3052137 JLTEDG 0733-8724 (2021). Google Scholar

425. 

M. Lobry et al., “Plasmonic fiber grating biosensors demodulated through spectral envelopes intersection,” J. Lightwave Technol., 39 7288 https://doi.org/10.1109/JLT.2021.3112854 JLTEDG 0733-8724 (2021). Google Scholar

426. 

J. Scheuer and M. Sumetsky, “Optical-fiber microcoil waveguides and resonators and their applications for interferometry and sensing,” Laser Photon. Rev., 5 465 https://doi.org/10.1002/lpor.201000022 (2011). Google Scholar

427. 

J. L. Kou et al., “Microfiber-based Bragg gratings for sensing applications: a review,” Sensors, 12 8861 https://doi.org/10.3390/s120708861 SNSRES 0746-9462 (2012). Google Scholar

428. 

J. Y. Lou, Y. P. Wang and L. M. Tong, “Microfiber optical sensors: a review,” Sensors, 14 5823 https://doi.org/10.3390/s140405823 SNSRES 0746-9462 (2014). Google Scholar

429. 

P. F. Wang et al., “Optical microfibre based photonic components and their applications in label-free biosensing,” Biosensors-Basel, 5 471 https://doi.org/10.3390/bios5030471 (2015). Google Scholar

430. 

G. Y. Chen, D. G. Lancaster and T. M. Monro, “Optical microfiber technology for current, temperature, acceleration, acoustic, humidity and ultraviolet light sensing,” Sensors, 18 72 https://doi.org/10.3390/s18010072 SNSRES 0746-9462 (2017). Google Scholar

431. 

L. T. Gai, J. Li and Y. Zhao, “Preparation and application of microfiber resonant ring sensors: a review,” Opt. Laser Technol., 89 126 https://doi.org/10.1016/j.optlastec.2016.10.002 OLTCAS 0030-3992 (2017). Google Scholar

432. 

Y. Peng et al., “Research advances in microfiber humidity sensors,” Small, 14 1800524 https://doi.org/10.1002/smll.201800524 SMALBC 1613-6810 (2018). Google Scholar

433. 

Y. Wu et al., “Optical graphene gas sensors based on microfibers: a review,” Sensors, 18 941 https://doi.org/10.3390/s18040941 SNSRES 0746-9462 (2018). Google Scholar

434. 

J. H. Chen, D. R. Li and F. Xu, “Optical microfiber sensors: sensing mechanisms, and recent advances,” J. Lightwave Technol., 37 2577 https://doi.org/10.1109/JLT.2018.2877434 JLTEDG 0733-8724 (2019). Google Scholar

435. 

Y. P. Li et al., “Recent advances in microfiber sensors for highly sensitive biochemical detection,” J. Phys. D-Appl. Phys., 52 493002 https://doi.org/10.1088/1361-6463/ab3d4e (2019). Google Scholar

436. 

Z. L. Ran et al., “Fiber-optic microstructure sensors: a review,” Photonic Sens., 11 227 https://doi.org/10.1007/s13320-021-0632-7 (2021). Google Scholar

437. 

Y. N. Zhang et al., “Microfiber knot resonators: structure, spectral properties, and sensing applications,” Laser Photon. Rev., 18 2300765 https://doi.org/10.1002/lpor.202300765 (2024). Google Scholar

438. 

F. Warken et al., “Ultra-sensitive surface absorption spectroscopy using sub-wavelength diameter optical fibers,” Opt. Express, 15 11952 https://doi.org/10.1364/OE.15.011952 OPEXFF 1094-4087 (2007). Google Scholar

439. 

X. C. Yu et al., “Single nanoparticle detection and sizing using a nanofiber pair in an aqueous environment,” Adv. Mater., 26 7462 https://doi.org/10.1002/adma.201402085 ADVMEW 0935-9648 (2014). Google Scholar

440. 

X. C. Yu et al., “Optically sizing single atmospheric particulates with a 10-nm resolution using a strong evanescent field,” Light Sci. Appl., 7 18003 https://doi.org/10.1038/lsa.2018.3 (2018). Google Scholar

441. 

K. W. Li et al., “Gold nanoparticle amplified optical microfiber evanescent wave absorption biosensor for cancer biomarker detection in serum,” Talanta, 120 419 https://doi.org/10.1016/j.talanta.2013.11.085 TLNTA2 0039-9140 (2014). Google Scholar

442. 

Y. Y. Huang et al., “A fiber-optic sensor for neurotransmitters with ultralow concentration: near-infrared plasmonic electromagnetic field enhancement using raspberry-like meso-SiO2 nanospheres,” Nanoscale, 9 14929 https://doi.org/10.1039/C7NR05032A NANOHL 2040-3364 (2017). Google Scholar

443. 

P. Xiao et al., “Efficiently writing Bragg grating in high-birefringence elliptical microfiber for label-free immunosensing with temperature compensation,” Adv. Fiber Mater., 3 321 https://doi.org/10.1007/s42765-021-00087-7 (2021). Google Scholar

444. 

K. Q. Kieu and M. Mansuripur, “Biconical fiber taper sensors,” IEEE Photon. Technol. Lett., 18 2239 https://doi.org/10.1109/LPT.2006.884742 IPTLEL 1041-1135 (2006). Google Scholar

445. 

W. B. Ji et al., “Ultrahigh sensitivity refractive index sensor based on optical microfiber,” IEEE Photon. Technol. Lett., 24 1872 https://doi.org/10.1109/LPT.2012.2217738 IPTLEL 1041-1135 (2012). Google Scholar

446. 

P. F. Wang et al., “High-sensitivity, evanescent field refractometric sensor based on a tapered, multimode fiber interference,” Opt. Lett., 36 2233 https://doi.org/10.1364/OL.36.002233 OPLEDP 0146-9592 (2011). Google Scholar

447. 

G. Salceda-Delgado et al., “Optical microfiber mode interferometer for temperature-independent refractometric sensing,” Opt. Lett., 37 1974 https://doi.org/10.1364/OL.37.001974 OPLEDP 0146-9592 (2012). Google Scholar

448. 

Q. Z. Sun et al., “Multimode microfiber interferometer for dual-parameters sensing assisted by fresnel reflection,” Opt. Express, 23 12777 https://doi.org/10.1364/OE.23.012777 OPEXFF 1094-4087 (2015). Google Scholar

449. 

M. Z. Muhammad et al., “Non-adiabatic silica microfiber for strain and temperature sensors,” Sens. Actuator A-Phys., 192 130 https://doi.org/10.1016/j.sna.2012.12.036 (2013). Google Scholar

450. 

W. Li et al., “High-sensitivity microfiber strain and force sensors,” Opt. Commun., 314 28 https://doi.org/10.1016/j.optcom.2013.09.072 OPCOB8 0030-4018 (2014). Google Scholar

451. 

Y. Z. Zheng et al., “Optical fiber magnetic field sensor based on magnetic fluid and microfiber mode interferometer,” Opt. Commun., 336 5 https://doi.org/10.1016/j.optcom.2014.09.026 OPCOB8 0030-4018 (2015). Google Scholar

452. 

L. F. Luo et al., “Highly sensitive magnetic field sensor based on microfiber coupler with magnetic fluid,” Appl. Phys. Lett., 106 193507 https://doi.org/10.1063/1.4921267 APPLAB 0003-6951 (2015). Google Scholar

453. 

K. W. Li et al., “Ultrasensitive optical microfiber coupler based sensors operating near the turning point of effective group index difference,” Appl. Phys. Lett., 109 13 https://doi.org/10.1063/1.4961981 APPLAB 0003-6951 (2016). Google Scholar

454. 

Y. Chen et al., “A miniature reflective micro-force sensor based on a microfiber coupler,” Opt. Express, 22 2443 https://doi.org/10.1364/OE.22.002443 OPEXFF 1094-4087 (2014). Google Scholar

455. 

S. L. Pu et al., “Ultrasensitive refractive-index sensors based on tapered fiber coupler with Sagnac loop,” IEEE Photon. Technol. Lett., 28 1073 https://doi.org/10.1109/LPT.2016.2529181 IPTLEL 1041-1135 (2016). Google Scholar

456. 

J. Y. Lou, L. M. Tong and Z. Z. Ye, “Modeling of silica nanowires for optical sensing,” Opt. Express, 13 2135 https://doi.org/10.1364/OPEX.13.002135 OPEXFF 1094-4087 (2005). Google Scholar

457. 

A. A. Jasim et al., “Microfibre Mach-Zehnder interferometer and its application as a current sensor,” IET Optoelectron., 6 298 https://doi.org/10.1049/iet-opt.2012.0025 (2012). Google Scholar

458. 

H. P. Luo et al., “Refractive index sensitivity characteristics near the dispersion turning point of the multimode microfiber-based Mach-Zehnder interferometer,” Opt. Lett., 40 5042 https://doi.org/10.1364/OL.40.005042 OPLEDP 0146-9592 (2015). Google Scholar

459. 

C. L. Hou et al., “Novel high sensitivity accelerometer based on a microfiber loop resonator,” Opt. Eng., 49 014402 https://doi.org/10.1117/1.3294883 (2010). Google Scholar

460. 

M. Belal et al., “Optical fiber microwire current sensor,” Opt. Lett., 35 3045 https://doi.org/10.1364/OL.35.003045 OPLEDP 0146-9592 (2010). Google Scholar

461. 

G. Y. Chen et al., “Resonantly enhanced Faraday rotation in an microcoil current sensor,” IEEE Photon. Technol. Lett., 24 860 https://doi.org/10.1109/LPT.2012.2189381 IPTLEL 1041-1135 (2012). Google Scholar

462. 

X. D. Xie et al., “A high-sensitivity current sensor utilizing CrNi wire and microfiber coils,” Sensors, 14 8423 https://doi.org/10.3390/s140508423 SNSRES 0746-9462 (2014). Google Scholar

463. 

G. Y. Chen, G. Brambilla and T. P. Newson, “Compact acoustic sensor based on air-backed mandrel coiled with optical microfiber,” Opt. Lett., 37 4720 https://doi.org/10.1364/OL.37.004720 OPLEDP 0146-9592 (2012). Google Scholar

464. 

R. Lorenzi, Y. M. Jung and G. Brambilla, “In-line absorption sensor based on coiled optical microfiber,” Appl. Phys. Lett., 98 173504 https://doi.org/10.1063/1.3582924 APPLAB 0003-6951 (2011). Google Scholar

465. 

H. Y. Mei et al., “Coiled optical nanofiber for optofluidic absorbance detection,” ACS Sens., 4 2267 https://doi.org/10.1021/acssensors.9b00913 (2019). Google Scholar

466. 

W. Yu et al., “Highly sensitive and fast response strain sensor based on evanescently coupled micro/nanofibers,” Opto-Electron. Adv., 5 210101 https://doi.org/10.29026/oea.2022.210101 (2022). Google Scholar

467. 

C. R. Liao, D. N. Wang and Y. Wang, “Microfiber in-line Mach-Zehnder interferometer for strain sensing,” Opt. Lett., 38 757 https://doi.org/10.1364/OL.38.000757 OPLEDP 0146-9592 (2013). Google Scholar

468. 

I. Martincek and D. Kacik, “A PDMS microfiber Mach-Zehnder interferometer and determination of nanometer displacements,” Opt. Fiber Technol., 40 13 https://doi.org/10.1016/j.yofte.2017.10.006 1068-5200 (2018). Google Scholar

469. 

K. J. Liu et al., “Highly sensitive vibration sensor based on the dispersion turning point microfiber Mach-Zehnder interferometer,” Opt. Express, 29 32983 https://doi.org/10.1364/OE.439959 OPEXFF 1094-4087 (2021). Google Scholar

470. 

J. H. Li, J. H. Chen and F. Xu, “Sensitive and wearable optical microfiber sensor for human health monitoring,” Adv. Mater. Technol., 3 1800296 https://doi.org/10.1002/admt.201800296 (2018). Google Scholar

471. 

S. S. Wang et al., “Modeling seawater salinity and temperature sensing based on directional coupler assembled by polyimide-coated micro/nanofibers,” Appl. Opt., 54 10283 https://doi.org/10.1364/AO.54.010283 APOPAI 0003-6935 (2015). Google Scholar

472. 

Y. X. Jiang et al., “Highly sensitive temperature sensor using packaged optical microfiber coupler filled with liquids,” Opt. Express, 26 356 https://doi.org/10.1364/OE.26.000356 OPEXFF 1094-4087 (2018). Google Scholar

473. 

Y. G. Han, “Investigation of temperature sensitivity of a polymer-overlaid microfiber Mach-Zehnder interferometer,” Sensors, 17 2403 https://doi.org/10.3390/s17102403 SNSRES 0746-9462 (2017). Google Scholar

474. 

A. A. Jasim et al., “Inline microfiber Mach-Zehnder interferometer for high temperature sensing,” IEEE Sens. J., 13 626 https://doi.org/10.1109/JSEN.2012.2224106 ISJEAZ 1530-437X (2013). Google Scholar

475. 

Y. Wu et al., “Miniature fiber-optic temperature sensors based on silica/polymer microfiber knot resonators,” Opt. Express, 17 18142 https://doi.org/10.1364/OE.17.018142 OPEXFF 1094-4087 (2009). Google Scholar

476. 

X. Zeng et al., “A temperature sensor based on optical microfiber knot resonator,” Opt. Commun., 282 3817 https://doi.org/10.1016/j.optcom.2009.05.079 OPCOB8 0030-4018 (2009). Google Scholar

477. 

J. Li et al., “A high sensitivity temperature sensor based on packaged microfibre knot resonator,” Sens. Actuator A-Phys., 263 369 https://doi.org/10.1016/j.sna.2017.06.031 (2017). Google Scholar

478. 

Y. T. Bai et al., “Simultaneous measurement of relative humidity and temperature using a microfiber coupler coated with molybdenum disulfide nanosheets,” Opt. Mater. Express, 9 2846 https://doi.org/10.1364/OME.9.002846 (2019). Google Scholar

479. 

Y. Z. Tan et al., “Microfiber Mach-Zehnder interferometer based on long period grating for sensing applications,” Opt. Express, 21 154 https://doi.org/10.1364/OE.21.000154 OPEXFF 1094-4087 (2013). Google Scholar

480. 

H. P. Luo et al., “Microfiber-based inline Mach-Zehnder interferometer for dual-parameter measurement,” IEEE Photonics J., 7 7100908 https://doi.org/10.1109/JPHOT.2015.2395133 (2015). Google Scholar

481. 

Z. Chen et al., “Optically tunable microfiber-knot resonator,” Opt. Express, 19 14217 https://doi.org/10.1364/OE.19.014217 OPEXFF 1094-4087 (2011). Google Scholar

482. 

S. S. Pal et al., “Etched multimode microfiber knot-type loop interferometer refractive index sensor,” Rev. Sci. Instrum., 82 095107 https://doi.org/10.1063/1.3633955 RSINAK 0034-6748 (2011). Google Scholar

483. 

L. P. Sun et al., “Miniature highly-birefringent microfiber loop with extremely-high refractive index sensitivity,” Opt. Express, 20 10180 https://doi.org/10.1364/OE.20.010180 OPEXFF 1094-4087 (2012). Google Scholar

484. 

S. C. Yan et al., “Differential twin receiving fiber-optic magnetic field and electric current sensor utilizing a microfiber coupler,” Opt. Express, 23 9407 https://doi.org/10.1364/OE.23.009407 OPEXFF 1094-4087 (2015). Google Scholar

485. 

X. L. Li and H. Ding, “All-fiber magnetic-field sensor based on microfiber knot resonator and magnetic fluid,” Opt. Lett., 37 5187 https://doi.org/10.1364/OL.37.005187 OPLEDP 0146-9592 (2012). Google Scholar

486. 

X. L. Li and H. Ding, “Temperature insensitive magnetic field sensor based on ferrofluid clad microfiber resonator,” IEEE Photon. Technol. Lett., 26 2426 https://doi.org/10.1109/LPT.2014.2358454 IPTLEL 1041-1135 (2014). Google Scholar

487. 

W. C. Zhou et al., “Ultrasensitive label-free optical microfiber coupler biosensor for detection of cardiac troponin I based on interference turning point effect,” Biosens. Bioelectron., 106 99 https://doi.org/10.1016/j.bios.2018.01.061 BBIOE4 0956-5663 (2018). Google Scholar

488. 

Y. Zhang et al., “Refractive index sensing based on higher-order mode reflection of a microfiber Bragg grating,” Opt. Express, 18 26345 https://doi.org/10.1364/OE.18.026345 OPEXFF 1094-4087 (2010). Google Scholar

489. 

Y. Ran et al., “193nm excimer laser inscribed Bragg gratings in microfibers for refractive index sensing,” Opt. Express, 19 18577 https://doi.org/10.1364/OE.19.018577 OPEXFF 1094-4087 (2011). Google Scholar

490. 

Y. X. Zhang et al., “Magnetic field and temperature dual-parameter sensor based on nonadiabatic tapered microfiber cascaded with FBG,” IEEE Access, 10 15478 https://doi.org/10.1109/ACCESS.2022.3148211 (2022). Google Scholar

491. 

W. Luo et al., “Ultra-highly sensitive surface-corrugated microfiber Bragg grating force sensor,” Appl. Phys. Lett., 101 133502 https://doi.org/10.1063/1.4754838 APPLAB 0003-6951 (2012). Google Scholar

492. 

Y. Wu et al., “Graphene-coated microfiber Bragg grating for high-sensitivity gas sensing,” Opt. Lett., 39 1235 https://doi.org/10.1364/OL.39.001235 OPLEDP 0146-9592 (2014). Google Scholar

493. 

D. D. Sun et al., “In-situ DNA hybridization detection with a reflective microfiber grating biosensor,” Biosens. Bioelectron., 61 541 https://doi.org/10.1016/j.bios.2014.05.065 BBIOE4 0956-5663 (2014). Google Scholar

494. 

T. Liu et al., “A label-free cardiac biomarker immunosensor based on phase-shifted microfiber Bragg grating,” Biosens. Bioelectron., 100 155 https://doi.org/10.1016/j.bios.2017.08.061 BBIOE4 0956-5663 (2018). Google Scholar

495. 

D. R. Li et al., “Label-free fiber nanograting sensor for real-time in situ early monitoring of cellular apoptosis,” Adv. Photonics, 4 016001 AOPAC7 1943-8206 (2022). Google Scholar

496. 

R. Gao, M. Y. Zhang and Z. M. Qi, “Miniature all-fibre microflown directional acoustic sensor based on crossed self-heated micro-Co2+-doped optical fibre Bragg gratings,” Appl. Phys. Lett., 113 134102 https://doi.org/10.1063/1.5043519 APPLAB 0003-6951 (2018). Google Scholar

497. 

E. L. Song et al., “Near-infrared microfiber Bragg grating for sensitive measurement of tension and bending,” Opt. Express, 31 15674 https://doi.org/10.1364/OE.487533 OPEXFF 1094-4087 (2023). Google Scholar

498. 

J. Villatoro and D. Monzon-Hernandez, “Fast detection of hydrogen with nano fiber tapers coated with ultra thin palladium layers,” Opt. Express, 13 5087 https://doi.org/10.1364/OPEX.13.005087 OPEXFF 1094-4087 (2005). Google Scholar

499. 

L. Zhang et al., “Fast detection of humidity with a subwavelength-diameter fiber taper coated with gelatin film,” Opt. Express, 16 13349 https://doi.org/10.1364/OE.16.013349 OPEXFF 1094-4087 (2008). Google Scholar

500. 

F. Schedin et al., “Detection of individual gas molecules adsorbed on graphene,” Nat. Mater., 6 652 https://doi.org/10.1038/nmat1967 NMAACR 1476-1122 (2007). Google Scholar

501. 

S. C. Yan et al., “Optical electrical current sensor utilizing a graphene-microfiber-integrated coil resonator,” Appl. Phys. Lett., 107 053502 https://doi.org/10.1063/1.4928247 APPLAB 0003-6951 (2015). Google Scholar

502. 

Q. Z. Sun et al., “Graphene-assisted microfiber for optical-power-based temperature sensor,” IEEE Photon. Technol. Lett., 28 383 https://doi.org/10.1109/LPT.2015.2495107 IPTLEL 1041-1135 (2016). Google Scholar

503. 

S. Wang et al., “Highly sensitive temperature sensor based on gain competition mechanism using graphene coated microfiber,” IEEE Photonics J., 10 6802008 https://doi.org/10.1109/JPHOT.2018.2827073 (2018). Google Scholar

504. 

C. B. Yu et al., “Graphene oxide deposited microfiber knot resonator for gas sensing,” Opt. Mater. Express, 6 727 https://doi.org/10.1364/OME.6.000727 (2016). Google Scholar

505. 

Y. Y. Huang et al., “Ultrasensitive and in situ DNA detection in various pH environments based on a microfiber with a graphene oxide linking layer,” Rsc Adv., 7 14322 https://doi.org/10.1039/C7RA90036H (2017). Google Scholar

506. 

Y. Huang et al., “Ultrafast response optical microfiber interferometric VOC sensor based on evanescent field interaction with ZIF-8/graphene oxide nanocoating,” Adv. Opt. Mater., 10 2101561 https://doi.org/10.1002/adom.202101561 2195-1071 (2022). Google Scholar

507. 

Y. Yin et al., “Ultra-high-resolution detection of Pb2+ ions using a black phosphorus functionalized microfiber coil resonator,” Photonics Res., 7 622 https://doi.org/10.1364/PRJ.7.000622 (2019). Google Scholar

508. 

G. W. Chen et al., “Highly sensitive all-optical control of light in WS2 coated microfiber knot resonator,” Opt. Express, 26 27650 https://doi.org/10.1364/OE.26.027650 OPEXFF 1094-4087 (2018). Google Scholar

509. 

J. G. Chen et al., “A microfiber sensor for the trace copper ions detection based on ternary sensitive film,” Adv. Mater. Interfaces, 9 2200491 https://doi.org/10.1002/admi.202200491 (2022). Google Scholar

510. 

H. T. Li et al., “Single-molecule detection of biomarker and localized cellular photothermal therapy using an optical microfiber with nanointerface,” Sci. Adv., 5 eaax4659 https://doi.org/10.1126/sciadv.aax4659 STAMCV 1468-6996 (2019). Google Scholar

511. 

N. M. Y. Zhang et al., “One-step synthesis of cyclodextrin-capped gold nanoparticles for ultra-sensitive and highly-integrated plasmonic biosensors,” Sens. Actuator B-Chem., 286 429 https://doi.org/10.1016/j.snb.2019.01.166 (2019). Google Scholar

512. 

A. X. Xiao et al., “Ultrasensitive detection and cellular photothermal therapy via a self-photothermal modulation biosensor,” Adv. Opt. Mater., 11 2202711 https://doi.org/10.1002/adom.202202711 2195-1071 (2023). Google Scholar

513. 

S. P. Wang et al., “Optical-nanofiber-enabled gesture-recognition wristband for human-machine interaction with the assistance of machine learning,” Adv. Intell. Syst., 5 2200412 https://doi.org/10.1002/aisy.202200412 (2023). Google Scholar

514. 

H. T. Zhu et al., “Spatiotemporal hemodynamic monitoring via configurable skin-like microfiber Bragg grating group,” Opto-Electron. Adv., 6 230018 https://doi.org/10.29026/oea.2023.230018 (2023). Google Scholar

515. 

S. Q. Ma et al., “Optical micro/nano fibers enabled smart textiles for human-machine interface,” Adv. Fiber Mater., 4 1108 https://doi.org/10.1007/s42765-022-00163-6 (2022). Google Scholar

516. 

H. T. Liu et al., “Optical microfibers for sensing proximity and contact in human-machine interfaces,” ACS Appl. Mater. Interfaces, 14 14447 https://doi.org/10.1021/acsami.1c23716 AAMICK 1944-8244 (2022). Google Scholar

517. 

L. Y. Li et al., “Wearable alignment-free microfiber-based sensor chip for precise vital signs monitoring and cardiovascular assessment,” Adv. Fiber Mater., 4 475 https://doi.org/10.1007/s42765-021-00121-8 (2022). Google Scholar

518. 

Y. Tang et al., “Optical micro/nanofiber-enabled compact tactile sensor for hardness discrimination,” ACS Appl. Mater. Interfaces, 13 4560 https://doi.org/10.1021/acsami.0c20392 AAMICK 1944-8244 (2021). Google Scholar

519. 

Z. Zhang et al., “Optical micro/nanofibre embedded soft film enables multifunctional flow sensing in microfluidic chips,” Lab Chip, 20 2572 https://doi.org/10.1039/D0LC00178C LCAHAM 1473-0197 (2020). Google Scholar

520. 

Z. Zhang et al., “A multifunctional airflow sensor enabled by optical micro/nanofiber,” Adv. Fiber Mater., 3 359 https://doi.org/10.1007/s42765-021-00097-5 (2021). Google Scholar

521. 

H. T. Zhu et al., “Self-assembled wavy optical microfiber for stretchable wearable sensor,” Adv. Opt. Mater., 9 2002206 https://doi.org/10.1002/adom.202002206 2195-1071 (2021). Google Scholar

522. 

C. P. Jiang et al., “Finger-skin-inspired flexible optical sensor for force sensing and slip detection in robotic grasping,” Adv. Mater. Technol., 6 2100285 https://doi.org/10.1002/admt.202100285 (2021). Google Scholar

523. 

R. Y. Zhao et al., “Research on acoustic sensing device based on microfiber knot resonator,” J. Micromech. Microeng., 32 085003 https://doi.org/10.1088/1361-6439/ac7842 JMMIEZ 0960-1317 (2022). Google Scholar

524. 

Z. L. Xu et al., “Highly sensitive refractive index sensor based on cascaded microfiber knots with Vernier effect,” Opt. Express, 23 6662 https://doi.org/10.1364/OE.23.006662 OPEXFF 1094-4087 (2015). Google Scholar

525. 

Y. T. Yi et al., “High-performance ultrafast humidity sensor based on microknot resonator-assisted Mach-Zehnder for monitoring human breath,” ACS Sens., 5 3404 https://doi.org/10.1021/acssensors.0c00863 (2020). Google Scholar

526. 

Z. Zhang et al., “A new route for fabricating polymer optical microcavities,” Nanoscale, 11 5203 https://doi.org/10.1039/C8NR10007A NANOHL 2040-3364 (2019). Google Scholar

527. 

P. P. Niu et al., “High-sensitive and disposable myocardial infarction biomarker immunosensor with optofluidic microtubule lasing,” Nanophotonics, 11 3351 https://doi.org/10.1515/nanoph-2022-0208 (2022). Google Scholar

528. 

W. Luo, F. Xu and Y. Q. Lu, “Reconfigurable optical-force-drive chirp and delay line in micro- or nanofiber Bragg grating,” Phys. Rev. A, 91 053831 https://doi.org/10.1103/PhysRevA.91.053831 (2015). Google Scholar

529. 

R. Gauthier, “Computation of the optical trapping force using an FDTD based technique,” Opt. Express, 13 3707 https://doi.org/10.1364/OPEX.13.003707 OPEXFF 1094-4087 (2005). Google Scholar

530. 

Z. H. Liu et al., “Tapered fiber optical tweezers for microscopic particle trapping: fabrication and application,” Opt. Express, 14 12510 https://doi.org/10.1364/OE.14.012510 OPEXFF 1094-4087 (2006). Google Scholar

531. 

L. Novotny and B. Hecht, Principles of nano-optics, second editionCambridge university press, (2012). Google Scholar

532. 

G. Brambilla et al., “Optical manipulation of microspheres along a subwavelength optical wire,” Opt. Lett., 32 3041 https://doi.org/10.1364/OL.32.003041 OPLEDP 0146-9592 (2007). Google Scholar

533. 

L. L. Xu, Y. Li and B. J. Li, “Size-dependent trapping and delivery of submicro-spheres using a submicrofibre,” New J. Phys., 14 033020 https://doi.org/10.1088/1367-2630/14/3/033020 NJOPFM 1367-2630 (2012). Google Scholar

534. 

H. X. Lei et al., “Bidirectional optical transportation and controllable positioning of nanoparticles using an optical nanofiber,” Nanoscale, 4 6707 https://doi.org/10.1039/c2nr31993d NANOHL 2040-3364 (2012). Google Scholar

535. 

Y. Zhang and B. J. Li, “Particle sorting using a subwavelength optical fiber,” Laser Photon. Rev., 7 289 https://doi.org/10.1002/lpor.201200087 (2013). Google Scholar

536. 

H. B. Xin, C. Cheng and B. J. Li, “Trapping and delivery of Escherichia coli in a microfluidic channel using an optical nanofiber,” Nanoscale, 5 6720 https://doi.org/10.1039/c3nr02088f NANOHL 2040-3364 (2013). Google Scholar

537. 

A. Maimaiti et al., “Higher order microfibre modes for dielectric particle trapping and propulsion,” Sci. Rep., 5 9077 https://doi.org/10.1038/srep09077 SRCEC3 2045-2322 (2015). Google Scholar

538. 

A. Maimaiti et al., “Nonlinear force dependence on optically bound micro-particle arrays in the evanescent fields of fundamental and higher order microfibre modes,” Sci. Rep., 6 30131 https://doi.org/10.1038/srep30131 SRCEC3 2045-2322 (2016). Google Scholar

539. 

F. Le Kien and A. Rauschenbeutel, “Negative azimuthal force of nanofiber-guided light on a particle,” Phys. Rev. A, 88 063845 https://doi.org/10.1103/PhysRevA.88.063845 (2013). Google Scholar

540. 

G. Tkachenko et al., “Light-induced rotation of dielectric microparticles around an optical nanofiber,” Optica, 7 59 https://doi.org/10.1364/OPTICA.374441 (2020). Google Scholar

541. 

J. S. Lu et al., “Light-induced pulling and pushing by the synergic effect of optical force and photophoretic force,” Phys. Rev. Lett., 118 043601 https://doi.org/10.1103/PhysRevLett.118.043601 PRLTAO 0031-9007 (2017). Google Scholar

542. 

W. W. Tang et al., “Micro-scale opto-thermo-mechanical actuation in the dry adhesive regime,” Light Sci. Appl., 10 193 https://doi.org/10.1038/s41377-021-00622-6 (2021). Google Scholar

543. 

W. Lyu et al., “Light-induced in-plane rotation of microobjects on microfibers,” Laser Photon. Rev., 16 2100561 https://doi.org/10.1002/lpor.202100561 (2022). Google Scholar

544. 

W. Lyu et al., “Nanomotion of micro-objects driven by light-induced elastic waves on solid interfaces,” Phys. Rev. Appl., 19 024049 https://doi.org/10.1103/PhysRevApplied.19.024049 PRAHB2 2331-7019 (2023). Google Scholar

545. 

S. Y. Linghu et al., “Plasmon-driven nanowire actuators for on-chip manipulation,” Nat. Commun., 12 385 https://doi.org/10.1038/s41467-020-20683-2 NCAOBW 2041-1723 (2021). Google Scholar

546. 

H. Fujiwara et al., “Optical selection and sorting of nanoparticles according to quantum mechanical properties,” Sci. Adv., 7 eabd9551 https://doi.org/10.1126/sciadv.abd9551 STAMCV 1468-6996 (2021). Google Scholar

547. 

Y. H. Li et al., “Modeling rare-earth doped microfiber ring lasers,” Opt. Express, 14 7073 https://doi.org/10.1364/OE.14.007073 OPEXFF 1094-4087 (2006). Google Scholar

548. 

W. Fan et al., “Tunable dual-wavelength narrow-linewidth microfiber laser,” Appl. Phys. Express, 6 072701 https://doi.org/10.7567/APEX.6.072701 APEPC4 1882-0778 (2013). Google Scholar

549. 

Z. S. Zhang et al., “Single-frequency microfiber single-knot laser,” Appl. Phys. Express, 6 042702 https://doi.org/10.7567/APEX.6.042702 APEPC4 1882-0778 (2013). Google Scholar

550. 

W. Fan et al., “A wavelength tunable single frequency microfiber laser,” Laser Physics Letters, 11 015104 https://doi.org/10.1088/1612-2011/11/1/015104 1612-2011 (2014). Google Scholar

551. 

V. D. Ta et al., “Whispering gallery mode microlasers and refractive index sensing based on single polymer fiber,” Laser Photon. Rev., 7 133 https://doi.org/10.1002/lpor.201200074 (2013). Google Scholar

552. 

S. C. Yang, Y. Wang and H. D. Sun, “Advances and prospects for whispering gallery mode microcavities,” Adv. Opt. Mater., 3 1136 https://doi.org/10.1002/adom.201500232 2195-1071 (2015). Google Scholar

553. 

F. X. Gu et al., “Single whispering-gallery mode lasing in polymer bottle microresonators via spatial pump engineering,” Light Sci. Appl., 6 e17061 https://doi.org/10.1038/lsa.2017.61 (2017). Google Scholar

554. 

F. M. Xie et al., “Single-mode lasing via loss engineering in fiber-taper-coupled polymer bottle microresonators,” Photonics Res., 5 B29 https://doi.org/10.1364/PRJ.5.000B29 (2017). Google Scholar

555. 

P. L. Yang et al., “65-fs Yb-doped all-fiber laser using tapered fiber for nonlinearity and dispersion management,” Opt. Lett., 43 1730 https://doi.org/10.1364/OL.43.001730 OPLEDP 0146-9592 (2018). Google Scholar

556. 

A. Al-Kadry et al., “Mode-locked fiber laser based on chalcogenide microwires,” Opt. Lett., 40 4309 https://doi.org/10.1364/OL.40.004309 OPLEDP 0146-9592 (2015). Google Scholar

557. 

V. I. Balykin et al., “Atom trapping and guiding with a subwavelength-diameter optical fiber,” Phys. Rev. A, 70 011401 https://doi.org/10.1103/PhysRevA.70.011401 (2004). Google Scholar

558. 

F. Le Kien, V. I. Balykin and K. Hakuta, “Atom trap and waveguide using a two-color evanescent light field around a subwavelength-diameter optical fiber,” Phys. Rev. A, 70 063403 https://doi.org/10.1103/PhysRevA.70.063403 (2004). Google Scholar

559. 

E. Vetsch et al., “Optical interface created by laser-cooled atoms trapped in the evanescent field surrounding an optical nanofiber,” Phys. Rev. Lett., 104 203603 https://doi.org/10.1103/PhysRevLett.104.203603 PRLTAO 0031-9007 (2010). Google Scholar

560. 

J. D. Thompson et al., “Coupling a single trapped atom to a nanoscale optical cavity,” Science, 340 1202 https://doi.org/10.1126/science.1237125 SCIEAS 0036-8075 (2013). Google Scholar

561. 

A. Goban et al., “Superradiance for atoms trapped along a photonic crystal waveguide,” Phys. Rev. Lett., 115 063601 https://doi.org/10.1103/PhysRevLett.115.063601 PRLTAO 0031-9007 (2015). Google Scholar

562. 

J. Fu et al., “Atom waveguide and 1D optical lattice using a two-color evanescent light field around an optical micro/nano-fiber,” Chin. Opt. Lett., 6 112 https://doi.org/10.3788/COL20080602.0112 CJOEE3 1671-7694 (2008). Google Scholar

563. 

S. T. Dawkins et al., “Dispersive optical interface based on nanofiber-trapped atoms,” Phys. Rev. Lett., 107 243601 https://doi.org/10.1103/PhysRevLett.107.243601 PRLTAO 0031-9007 (2011). Google Scholar

564. 

A. Goban et al., “Demonstration of a state-insensitive, compensated nanofiber trap,” Phys. Rev. Lett., 109 033603 https://doi.org/10.1103/PhysRevLett.109.033603 PRLTAO 0031-9007 (2012). Google Scholar

565. 

D. Reitz et al., “Coherence properties of nanofiber-trapped cesium atoms,” Phys. Rev. Lett., 110 243603 https://doi.org/10.1103/PhysRevLett.110.243603 PRLTAO 0031-9007 (2013). Google Scholar

566. 

J. B. Béguin et al., “Generation and detection of a sub-poissonian atom number distribution in a one-dimensional optical lattice,” Phys. Rev. Lett., 113 263603 https://doi.org/10.1103/PhysRevLett.113.263603 PRLTAO 0031-9007 (2014). Google Scholar

567. 

J. A. Grover et al., “Photon-correlation measurements of atomic-cloud temperature using an optical nanofiber,” Phys. Rev. A, 92 013850 https://doi.org/10.1103/PhysRevA.92.013850 (2015). Google Scholar

568. 

K. P. Nayak, J. Wang and J. Keloth, “Real-time observation of single atoms trapped and interfaced to a nanofiber cavity,” Phys. Rev. Lett., 123 213602 https://doi.org/10.1103/PhysRevLett.123.213602 PRLTAO 0031-9007 (2019). Google Scholar

569. 

Y. Meng et al., “Imaging and localizing individual atoms interfaced with a nanophotonic waveguide,” Phys. Rev. Lett., 125 053603 https://doi.org/10.1103/PhysRevLett.125.053603 PRLTAO 0031-9007 (2020). Google Scholar

570. 

R. Pennetta et al., “Collective radiative dynamics of an ensemble of cold atoms coupled to an optical waveguide,” Phys. Rev. Lett., 128 073601 https://doi.org/10.1103/PhysRevLett.128.073601 PRLTAO 0031-9007 (2022). Google Scholar

571. 

S. Pucher et al., “Atomic spin-controlled non-reciprocal Raman amplification of fibre-guided light,” Nat. Photonics, 16 380 https://doi.org/10.1038/s41566-022-00987-z NPAHBY 1749-4885 (2022). Google Scholar

572. 

A. K. Patnaik, J. Q. Liang and K. Hakuta, “Slow light propagation in a thin optical fiber via electromagnetically induced transparency,” Phys. Rev. A, 66 063808 https://doi.org/10.1103/PhysRevA.66.063808 (2002). Google Scholar

573. 

B. Gouraud et al., “Demonstration of a memory for tightly guided light in an optical nanofiber,” Phys. Rev. Lett., 114 180503 https://doi.org/10.1103/PhysRevLett.114.180503 PRLTAO 0031-9007 (2015). Google Scholar

574. 

C. Sayrin et al., “Storage of fiber-guided light in a nanofiber-trapped ensemble of cold atoms,” Optica, 2 353 https://doi.org/10.1364/OPTICA.2.000353 (2015). Google Scholar

575. 

S. Kato and T. Aoki, “Strong coupling between a trapped single atom and an all-fiber cavity,” Phys. Rev. Lett., 115 093603 https://doi.org/10.1103/PhysRevLett.115.093603 PRLTAO 0031-9007 (2015). Google Scholar

576. 

S. K. Ruddell et al., “Collective strong coupling of cold atoms to an all-fiber ring cavity,” Optica, 4 576 https://doi.org/10.1364/OPTICA.4.000576 (2017). Google Scholar

577. 

R. Yalla et al., “Cavity quantum electrodynamics on a nanofiber using a composite photonic crystal cavity,” Phys. Rev. Lett., 113 143601 https://doi.org/10.1103/PhysRevLett.113.143601 PRLTAO 0031-9007 (2014). Google Scholar

578. 

H. Takashima et al., “Detailed numerical analysis of photon emission from a single light emitter coupled with a nanofiber Bragg cavity,” Opt. Express, 24 15050 https://doi.org/10.1364/OE.24.015050 OPEXFF 1094-4087 (2016). Google Scholar

579. 

W. F. Li, J. J. Du and S. Nic Chormaic, “Tailoring a nanofiber for enhanced photon emission and coupling efficiency from single quantum emitters,” Opt. Lett., 43 1674 https://doi.org/10.1364/OL.43.001674 OPLEDP 0146-9592 (2018). Google Scholar

580. 

D. Ðonlagić, “In-line higher order mode filters based on long highly uniform fiber tapers,” J. Lightwave Technol., 24 3532 https://doi.org/10.1109/JLT.2006.878497 JLTEDG 0733-8724 (2006). Google Scholar

581. 

Y. Chen et al., “Compact optical short-pass filters based on microfibers,” Opt. Lett., 33 2565 https://doi.org/10.1364/OL.33.002565 OPLEDP 0146-9592 (2008). Google Scholar

582. 

Y. Jung, G. Brambilla and D. J. Richardson, “Broadband single-mode operation of standard optical fibers by using a sub-wavelength optical wire filter,” Opt. Express, 16 14661 https://doi.org/10.1364/OE.16.014661 OPEXFF 1094-4087 (2008). Google Scholar

583. 

Y. Jung, G. Brambilla and D. J. Richardson, “Optical microfiber coupler for broadband single-mode operation,” Opt. Express, 17 5273 https://doi.org/10.1364/OE.17.005273 OPEXFF 1094-4087 (2009). Google Scholar

584. 

C. R. Liao et al., “Twisted optical microfibers for refractive index sensing,” IEEE Photon. Technol. Lett., 23 848 https://doi.org/10.1109/LPT.2011.2138126 IPTLEL 1041-1135 (2011). Google Scholar

585. 

M. Ding, P. F. Wang and G. Brambilla, “A microfiber coupler tip thermometer,” Opt. Express, 20 5402 https://doi.org/10.1364/OE.20.005402 OPEXFF 1094-4087 (2012). Google Scholar

586. 

Y. Yu et al., “High sensitivity all optical fiber conductivity-temperature-depth (CTD) sensing based on an optical microfiber coupler (OMC),” J. Lightwave Technol., 37 2739 https://doi.org/10.1109/JLT.2018.2878475 JLTEDG 0733-8724 (2019). Google Scholar

587. 

S. L. Yu et al., “Graphene decorated microfiber for ultrafast optical modulation,” Opt. Express, 23 10764 https://doi.org/10.1364/OE.23.010764 OPEXFF 1094-4087 (2015). Google Scholar

588. 

S. L. Yu et al., “2D materials for optical modulation: challenges and opportunities,” Adv. Mater., 29 1606128 https://doi.org/10.1002/adma.201606128 ADVMEW 0935-9648 (2017). Google Scholar

589. 

X. T. Gan et al., “Graphene-assisted all-fiber phase shifter and switching,” Optica, 2 468 https://doi.org/10.1364/OPTICA.2.000468 (2015). Google Scholar

590. 

Y. Z. Wang et al., “All-optical phosphorene phase modulator with enhanced stability under ambient conditions,” Laser Photon. Rev., 12 1800016 https://doi.org/10.1002/lpor.201800016 (2018). Google Scholar

591. 

X. Y. Wang et al., “Two-dimensional material integrated micro-nano fiber, the new opportunity in all-optical signal processing,” Nanophotonics, 12 2073 https://doi.org/10.1515/nanoph-2023-0223 (2023). Google Scholar

592. 

Q. Q. Cen et al., “Microtaper leaky-mode spectrometer with picometer resolution,” eLight, 3 9 https://doi.org/10.1186/s43593-023-00041-7 (2023). Google Scholar

593. 

X. Wang et al., “Subwavelength focusing by a micro/nanofiber array,” J. Opt. Soc. Am. A, 26 1827 https://doi.org/10.1364/JOSAA.26.001827 JOAOD6 0740-3232 (2009). Google Scholar

594. 

X. Hao et al., “Far-field super-resolution imaging using near-field illumination by micro-fiber,” Appl. Phys. Lett., 102 013104 https://doi.org/10.1063/1.4773572 APPLAB 0003-6951 (2013). Google Scholar

595. 

Y. J. Yang et al., “Topological pruning enables ultra-low Rayleigh scattering in pressure-quenched silica glass,” npj Comput. Mater., 6 139 https://doi.org/10.1038/s41524-020-00408-1 (2020). Google Scholar
CC BY: © The Authors. Published by CLP and SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Jianbin Zhang, Hubiao Fang, Pan Wang, Wei Fang, Lei Zhang, Xin Guo, and Limin Tong "Optical microfiber or nanofiber: a miniature fiber-optic platform for nanophotonics," Photonics Insights 3(1), R02 (1 March 2024). https://doi.org/10.3788/PI.2024.R02
Received: 5 December 2023; Accepted: 8 February 2024; Published: 1 March 2024
Advertisement
Advertisement
KEYWORDS
Silica

Near field optics

Waveguides

Glasses

Evanescence

Fiber optics

Chemical species

Back to Top