3
Fiber Mid‐IR Lasers

3.1 Introduction

Fiber lasers can produce, in a compact device, outputs with remarkable beam quality and high output power and can operate with long life, high power efficiency, and low cost of ownership. In many cases, fiber lasers are replacing solid‐state lasers in research, defense, and industry. Extending the emission wavelength into the mid‐infrared is one of the avenues of fiber laser development, which will clearly benefit numerous existing and future applications, e.g. in the field of medicine, remote sensing, welding of polymers, and for pumping longer‐wavelength mid‐IR or terahertz optical parametric oscillators.

Fiber lasers are typically pumped by laser diodes with free‐space or fiber‐coupled output (Figure 3.1) and consist of a special optical fiber whose core has been infused with active ions that create the optical gain. For mid‐IR fiber lasers, the dopants in the gain medium are rare‐earth ions such as thulium (Tm3+), erbium (Er3+), holmium (Ho3+), or dysprosium (Dy3+) [1].

Although the gain media for fiber lasers are similar to those of solid‐state bulk lasers, the waveguiding effect and the small effective mode area usually lead to substantially different properties. For example, fiber lasers often operate with much higher laser gain and higher resonator losses. Also, the spatial beam quality in fiber lasers is typically close to diffraction‐limited. Similar to solid‐state crystalline lasers, the lasing wavelength is determined by the energy levels of the rare‐earth cation doped into the core of the fiber.

Figure 3.2 shows the laser transitions responsible for infrared emission of rare‐earth‐cation‐doped fiber lasers, while the emission cross sections of rare‐earth laser transitions used in mid‐IR fiber lasers as a function of emission wavelength are plotted in Figure 3.3.

Image described by caption.

Figure 3.1 Setup of a simple fiber laser. Pump light is launched from the left‐hand side through a dichroic mirror into the core of the doped fiber. The generated laser light is extracted on the right‐hand side.

Energy-level diagrams for Tm3+, Er3+, Ho3+, and Dy3+ (left–right) illustrating laser transitions of rare-earth cations that produce mid-IR emission wavelengths.

Figure 3.2 Laser transitions of rare‐earth cations that produce mid‐IR emission wavelengths. ESA, excited state absorption.

Source: reproduced from figure 3 of [1], with permission of Springer Nature.

3.2 Continuous‐wave Mid‐IR Fiber Lasers

3.2.1 Tm‐based Fiber Lasers

Fiber lasers employing the Tm3+ cation [25] with emission at around 2 μm are the most powerful, efficient, and developed of the mid‐IR fiber lasers. These devices are pumped with robust diode lasers emitting at ~790 nm, and their output can be tuned from 1.86 to 2.09 μm [6]. With high concentrations of Tm3+ dopant ions, exceeding 2.5 weight%, cross‐relaxation between neighboring Tm3+ cations can nearly double the slope efficiency [7]. Similar to the solid‐state Tm‐based lasers reviewed in Chapter 2, the cross‐relaxation between Tm3+ cations creates two excited Tm3+ cations in the upper laser level for every pump photon excitation (Figure 3.4). This “two‐for‐one” excitation is resonant in silicate glass and provides an efficient way to generate 2 μm light using a commercial laser diode pump.

Image described by caption and surrounding text.

Figure 3.3 The emission cross section of common rare‐earth laser transitions used in fiber lasers emitting in the mid‐IR, as a function of emission wavelength. The energy levels that participate in each of the laser transitions are shown as labels. The oxide glass here refers to aluminosilicate and the fluoride glass refers to ZBLAN (ZrF4–BaF2–LaF3–AlF3–NaF composition).

Source: reproduced from figure 4 of [1], with permission of Springer Nature.

Kilowatt‐class thulium‐doped silica fiber laser with the central wavelength of 2.045 μm was reported in [5, 8]. The layout of such a laser is depicted in Figure 3.5 and is based on a 50‐W seed laser at 2.045 μm and a two‐stage power amplifier configuration. High‐power operation of the Tm‐fiber laser has been made possible due to (i) development of high‐brightness, fiber‐coupled 790‐nm pump lasers with nearly 40% electrical‐to‐optical conversion, (ii) fiber‐based pump‐coupling optics that provide a highly efficient combining of light from several multimode sources into one fiber, and (iii) “two‐for‐one” photon recycling due to the cross‐relaxation process. The power levels achieved in this work are: >500 W at 62% slope efficiency with respect to diode pump optical power with 6 pump lasers and 1.05 kW at 53% slope efficiency with 12 pump lasers (Figure 3.6). With the 20‐μm core diameter of the fiber amplifier, the laser showed single‐spatial‐mode operation. These results represent the highest continuous wave (CW) power level ever generated in this wavelength range [8].

Image described by caption.

Figure 3.4 Energy‐transfer processes between neighboring cations relevant to the operation of 2‐μm Tm3+ and 3‐μm Er3+ fiber lasers. (a) Cross‐relaxation between Tm3+ cations creates two excited Tm3+ cations in the upper laser level for every pump photon excitation. (b) Energy transfer between excited Er3+ cations forces the de‐excitation of one Er3+ cation and the further excitation of the other Er3+ cation taking part in the process.

Layout of two-stage power amplifier of the 1.05‐kW Tm‐doped silica fiber laser operating at λ = 2.045 μm. Master oscillator, pumps, pump couplers, cladding stripper, etc. are labeled.

Figure 3.5 Two‐stage power amplifier of the 1.05‐kW Tm‐doped silica fiber laser operating at λ = 2.045 μm. ΜΟ, master oscillator.

Source: reproduced from [8], with permission of Q‐peak, Inc.

The Tm‐fiber laser technology is now commercialized. For example, single‐spatial‐mode Tm‐fiber lasers with the wavelength selectable anywhere between 1900 and 2000 nm and with output power in the range of 1–200 W are now standard commercial products offered by IPG Photonics.1

Graph of signal power versus pump power displaying an ascending line with diamond markers having a slope of 53.2%.

Figure 3.6 Output power versus diode pump optical power of the 1.05‐kW single‐spatial‐mode Tm‐doped silica fiber laser at 2.045 μm.

Source: reproduced from [8], with permission of Q‐peak, Inc.

3.2.2 Ho‐based Fiber Lasers

The 5I7 → 5I8 transition of Ho3+ has a peak emission wavelength of 2.1 μm, which overlaps with the important atmospheric transmission window. This Ho transition can be pumped by 1150‐nm laser diodes, but it is also well suited for resonant pumping by the output of Tm3+‐doped silicate glass fiber lasers, where the quantum defect can be as small as 7% [9].

Alternatively, a very practical solution is a codoped Tm3+, Ho3+‐fiber laser, where thulium provides efficient excitation using standard 790‐nm laser diodes, and where the resonant energy transfer process efficiently populates the Ho3+ ion. With such a fiber, and using a bidirectional 793‐nm diode pumping, 83 W of CW output power has recently been achieved at 2.105 μm, with up to 42% slope efficiency [10].

The long‐wavelength 5I6 → 5I7 (λ ≈ 2.9 μm) laser transition of Ho3+ is self‐terminating, because the lifetime of the upper laser level is shorter than that of the lower laser level. A simplified energy‐level diagram for the cascade laser is shown in Figure 3.7. As a solution to this problem, a cascaded lasing of both the 3 μm (5I6 → 5I7) and 2.1 μm (5I7 → 5I8) transitions of Ho3+ has recently been demonstrated [11]. The laser used Ho3+‐doped ZBLAN fiber (ZBLAN is the most common fluoride glass used in mid‐IR fiber lasers with the following composition: ZrF4–BaF2–LaF3–AlF3–NaF) and two commercial 1150‐nm diode lasers (total launched pump power of 7.6 W) to pump directly the upper laser level 5I6. Cascading in this case helps to quickly unload the long‐living 5I7 energy level, which serves as a lower laser level for the 3‐μm transition. The laser produced 0.77 W at 3.002 μm and 0.24 W at 2.1 μm, making this system the first watt‐level fiber laser operating at ~3 μm in the mid‐IR.

Image described by caption and surrounding text.

Figure 3.7 Energy‐level diagram for a Ho3+‐doped cascaded ZBLAN fiber laser showing the pump, laser, and energy‐transfer upconversion processes. GSA, ground state absorption; ESA, excited state absorption.

Source: reproduced from figure 1 of [11], with permission of OSA, The Optical Society.

De‐excitation of the lower laser level can also be achieved by codoping of Ho3+‐fibers with praseodymium (Pr3+) ions. An efficient high‐power operation of a 2.94‐μm Ho3+, Pr3+‐doped fluoride glass fiber laser was demonstrated in [12]. The core of the laser was made of a Ho3+, Pr3+‐doped ZBLAN and the diode pumping was performed through fiber cladding. The laser produced a maximum output power of 2.5 W at 32% slope efficiency (that is 82% of the Stokes limit) with respect to the diode laser pump at 1150 nm. It is noteworthy that the emission wavelength of 2.94 μm measured at maximum pump power is particularly well suited for medical applications because it overlaps with the fundamental OH absorption peak of water in human tissue.

Finally, the laser action at the long‐wavelength 5I5 → 5I6 transition (λ = 3.9 μm) with a cascade laser scheme in a Ho3+‐doped fluoride fiber system was demonstrated under cryogenic (77 K) conditions [13]. With the fiber length of 340 cm, the pump wavelength of 885 nm, and a launched pump power of 900 mW, the researchers achieved 11 mW of CW output power at 3.9 μm. Simultaneously, λ = 1.2 μm (5I6 → 5I8 transition) radiation was emitted with 70‐mW output power. This result represents the longest wavelength emitted from a fiber laser.

3.2.3 Er‐based Fiber Lasers

The laser transitions between manifolds 4I11/2 and 4I13/2 of rare‐earth cation Er3+ (see Figure 3.2) are suitable for tuning across the mid‐IR range of 2.7–3.05 μm. There is a good overlap of the Er3+ absorption peak, which corresponds to the excitation of the upper laser level 4I11/2 with the wavelength range of highly developed diode lasers emitting at ~980 nm. This makes Er‐doped fiber lasers suitable candidates for efficient mid‐IR sources. Operation in the mid‐IR has necessitated the use of fluoride glass fiber (e.g. ZBLAN), owing to its low phonon energy. As an additional bonus, the energy transfer upconversion process in Er3+ ions (see Figure 3.4b) effectively depopulates the lower laser level and recycles the excitation to produce slope efficiencies as high as 35.6%, a value that exceeds the Stokes efficiency limit [14].

The first 1‐W class 3‐μm laser was demonstrated nearly two decades ago by Jackson et al. [15]. It was based on diode‐pumped Er3+, Pr3+:ZBLAN fiber and emitted at 2.71 μm with the output power of 1.7 W and slope efficiency of 17.3%. Since then, Er‐based fiber lasers have shown markedly improved performance. Zhu and Jain demonstrated the first 10‐W‐level fiber laser of the 3‐μm range [16]. In their experiment, the authors used a 4‐m‐long heavily erbium‐doped Er:ZBLAN double‐clad fiber, pumped by a 975‐nm laser‐diode array. The output power of 9 W was obtained at 2.785 μm at the launched pump power of 42.8 W. The output, however, fluctuated showing a pulsing behavior, and at higher pump rates the authors observed optical damage of the fiber end facets [16]. This unstable behavior was attributed to the poor thermal conductivity of Er:ZBLAN, its low melting temperature, and the large amount of heat that was generated by Er3+ dopants in the process of optical excitation.

Tokita et al. used liquid cooling to demonstrate 24 W of power from an Er:ZBLAN fiber laser operating at λ = 2.8 μm with 975 nm diode pumping [17]. The whole fiber was soaked in a fluorocarbon liquid coolant, which circulated to maintain a constant temperature of 20 °C. The 4.2‐m‐long fiber had an active core diameter of 25 μm, and the maximum power of 24 W was obtained at 166 W of the diode pump [17]. Later, the same team developed a 10‐W wavelength‐tunable Er:ZBLAN fiber laser operating without direct liquid cooling [18]. Instead, the fiber was cooled conductively when placed between aluminum plates that were maintained at 20 °C by water cooling. Also, to remove the heat from the output fiber end, the latter was polished into a spherical shape and put in optical contact with a 2‐mm‐thick sapphire plate (see Figure 3.8). The pump power was 93 W at 975 nm and an external diffraction grating in the Littrow configuration was used to tune the wavelength from 2.71 to 2.88 μm.

Image described by caption.

Figure 3.8 Schematic diagram of the grating‐tunable Er‐doped ZBLAN fiber laser pumped by a 975‐nm laser diode. The inset shows how the spherical polishing of the fiber end ensures optical contact between the fiber core and the sapphire plate. A diffraction grating in the Littrow configuration was used to tune the wavelength.

Source: reproduced from figure 1 of [18], with permission of OSA, The Optical Society.

Faucher et al. [19] obtained the output power of 20.6 W from an Er3+‐doped fluoride glass single‐mode all‐fiber laser source at λ = 2.825 μm. The fiber was passively cooled by using an aluminum spool; it was 4.6‐m long and had a single‐mode core diameter of 16 μm. An undoped fiber with highly reflective (99.4%) Bragg grating written in it was fusion‐spliced to the doped fiber and served as an end mirror. The pump source consisted of three 976‐nm, 30‐W fiber‐coupled modules, and the slope efficiency of the laser was up to 35.4% with respect to the absorbed pump power, which is higher than the Stokes efficiency of 34.3%.

Recently, Aydin et al. reported a passively cooled cascade Er3+‐doped fluoride fiber laser (λ = 2.825 μm) with an output power of ~13 W and a record slope efficiency of 50% (absolute optical efficiency of 37%) with respect to absorbed pump power at 976 nm [20]. The presence of an excited‐state absorption band centered at 1.675 μm, between the 4I13/2 and 4I9/2 levels (see Figure 3.2), which partially overlaps the cascaded (4I13/2 to 4I15/2) secondary emission at 1.614 μm, is believed to be responsible for the recycling of the excitation back to the upper laser level of the mid‐IR transition. One of the key features of this approach was to resonate the emission at 1.614 μm, in addition to resonating the 2.8‐μm emission.

Fortin et al. demonstrated a λ = 2.94 μm erbium‐doped fluoride glass fiber laser delivering a record output power of 30.5 W in CW mode [21]. The laser operated on the long‐wavelength transition – from the lowest Stark level of the 4I11/2 to the highest Stark level of the 4I13/2 manifold of the Er3+ ion. The passively cooled all‐fiber laser cavity was formed by fiber Bragg gratings, and the laser was pumped using seven laser diodes operating around 980 nm with combined power of 188 W. The overall laser efficiency with respect to the launched pump power at 980 nm was 16%. The laser produced a narrow (0.15 nm) linewidth and had a single‐mode beam quality with a M2 < 1.2 [21]. Since the 2.94 μm wavelength is resonant with the liquid water OH absorption peak, this laser has potential to find many applications in the biomedical field.

Image described by caption.

Figure 3.9 Schematic of the 3.44 μm Er‐fiber laser that relies on a dual pumping scheme at 974 and 1976 nm. The inset shows a simplified energy diagram for erbium ions in fluoride glasses. A residual pump stripper (RPS) made of high‐index acrylate polymer was applied to remove the remaining 791 nm pump. HR‐DM, high reflection dichroic mirror; LR‐FBG, low reflection fiber Bragg grating.

Source: reproduced from figure 1 of [22], with permission of OSA, The Optical Society.

On the longer wavelength side, a 3.44 μm all‐fiber laser emitting up to 1.5 W at room temperature was demonstrated in [22]. The laser operated on the 4F9/2 → 4I9/2 transition of erbium‐doped fluoride glasses and relied on a dual pumping scheme: at 974 and 1976 nm. In this case, a 974‐nm pump was used to excite Er3+ ions to the long‐lived level 4I11/2, thus creating a virtual ground state, while a 1976‐nm pump was used to populate the 3.5‐μm transition upper level 4F9/2 (Figure 3.9). By combining a dichroic mirror deposited on the input fiber tip and a fiber Bragg grating as an output coupler, a stable laser emission was produced with a bandwidth of less than 0.6 nm. The laser had an efficiency of 19% with respect to the launched pump power at 1976 nm and no saturation was observed, provided there is sufficient co‐pumping at 974 nm.

With an improved (so‐called “monolithic”) design of the dual‐wavelength‐pumped erbium‐doped fluorozirconate (Er3+:ZrF4) laser cavity, a multi‐watt laser operation at 3.55 μm was demonstrated [23]. The fiber laser cavity was bounded by two fiber Bragg gratings, which allowed high‐power performance without damage of the ZrF4 fiber tip. The maximum output power and total optical efficiency were, respectively, 5.6 W and 26.4%. The laser power was the highest ever achieved from a fiber laser at this wavelength.

Based on the idea of dual‐wavelength pumping, a broadly tunable long‐wavelength erbium‐doped Er:ZBLAN fiber laser was reported in [24]. The laser cavity consisted of a 2.8‐m‐long Er3+‐doped ZBLAN fiber gain medium with a 16‐μm‐diameter core and a diffraction grating. With the dual pumping at 977 and 1973 nm, the emission was centered at 3.5 μm, and the laser was tunable across 450 nm. A maximum power of 1.45 W at 3.47 μm was achieved with a 4‐W pump at 977 nm and a 5.5‐W pump at 1973 nm (overall optical efficiency 15%). The longest wavelength achieved was 3.78 μm, which is the longest wavelength emission from any fiber laser operating at room temperature [24].

3.2.4 Dy‐based Fiber Lasers

Majewski and Jackson presented an efficient and power‐scalable pump scheme for a 3‐μm laser, based on dysprosium‐doped fiber [25]. The authors used the free‐running 2.8‐μm emission from an Er3+‐doped fluoride fiber laser to directly excite, with a very small quantum defect of 8%, the upper laser level of the 6H13/26H15/2 transition of the Dy3+ ion. The Dy3+ laser output at 3.04 μm was produced with a slope efficiency of 51%. When the length of the Dy3+‐doped fluoride fiber was extended from 92 to 140 cm, the authors observed a maximum emission wavelength of 3.26 μm, with a slope efficiency of 32% with respect to the launched pump power.

3.2.5 Raman Fiber Lasers

Fundamentally, the Raman process relies on photon–phonon scattering, where an optical phonon (lattice vibration) mediates the process of frequency downconversion. The amount of red shift of the Stokes wave, with respect to the pump, depends on the frequency of the optical phonon for a given glass medium. In optical fibers, a variety of redshifted laser wavelengths extending beyond 2 μm were produced via stimulated Raman scattering. Typically, a Raman fiber laser consists of a piece of fiber enclosed between two highly reflective Bragg gratings written directly into the same fiber.

Using a germanate‐based Raman gain cavity, Dianov et al. produced the output at 2.193 μm, as the fourth Stokes component, with an average power of 210 mW. A 1608‐nm Er/Yb‐doped fiber laser with an average power of 5 W was utilized as a pump [26]. Cumberland et al. reported Raman 2.105‐μm emission wavelength from a high‐concentration GeO2 fiber Raman laser, with a maximum output of 4.6 W. The pump was a 22‐W thulium‐doped fiber laser at 1.938 μm [27].

In order to access longer mid‐IR wavelengths, materials with extended spectral transmission windows, e.g. glasses based on fluoride, tellurite, or chalcogenide, must be considered. Fortin et al. reported a Raman laser based on a fluoride glass optical fiber [28]. The laser was pumped by a Tm3+:silica CW fiber laser operating at 1940 nm with an average power of 9.6 W. A maximum output power of 580 mW was measured at 2185 nm (frequency shift of 579 cm−1), corresponding to 6% conversion efficiency with respect to the launched pump power. Later, the same team demonstrated a multi‐watt fluoride glass Raman fiber laser emitting at 2.23 μm. The laser was based on a nested (Raman laser inside pump laser) cavity with the pump being a Tm‐doped silica fiber laser at 1.98 μm. The maximum Stokes output power recorded at 2.23 μm was as high as 3.7 W [29].

Chalcogenide glasses, arsenic sulfide (As2S3) and selenide (As2Se3) – have long been identified as ideal candidates for mid‐IR Raman media because of (i) a high Raman gain coefficient (more than 50 times larger than that of a fluoride glass) and (ii) a transparency window that extends to beyond 6 and 9 μm respectively. Jackson et al. reported an As2Se3 chalcogenide Raman laser operating simultaneously on multiple Stokes bands. A 2.051‐μm Tm3+‐doped silica fiber laser with a 2‐W average power was employed as the pump source. The first Stokes emission was observed at 2.062 μm with an output power of 0.64 W. Two other Raman active vibrational modes of the fiber produced first Stokes outputs at 2.102 μm (0.2 W) and 2.166 μm (16 mW), respectively [30].

Bernier et al. reported the first Raman fiber laser emitting above 3 μm [31, 32]. The pump was an Er3+ fluoride glass fiber laser emitting at 3.005 μm (that is at the long‐wavelength edge of the Er3+ emission band) operating in the quasi‐CW mode with a pulse duration of 5 ms and a repetition rate of 20 Hz. Using a Fabry–Pérot cavity formed by a single‐mode As2S3 chalcogenide glass fiber and fiber Bragg gratings (Figure 3.10), the authors demonstrated a Raman fiber laser action at λ = 3.34 μm. The laser output average power was 50 mW (0.6 W peak) at the launched pump average power 250 mW [31].

A long‐wavelength quasi‐CW fiber Raman source, operating at room temperature, was reported in [33]. The 3.77‐μm output wavelength was generated by two cascades of Raman shifting; the Raman laser was based on a single‐mode As2S3 fiber and nested Fabry–Pérot cavities formed by two pairs of fiber Bragg gratings. Similar to [32], the pump was an Er3+:fluoride glass fiber laser at 3.005 μm operating in the quasi‐CW mode (5‐ms pulses, 20 Hz, duty cycle 10%). The first and second Stokes orders were at 3.345 and 3.766 μm, respectively. The maximum average output power of 9 mW was measured at 3.766 μm for an averaged launched pump power of 371 mW, corresponding to a peak power of 112 mW [33].

Image described by caption.

Figure 3.10 Schematic of the Raman As2S3 fiber laser emitting at 3.34 μm. The pump was an Er3+ ZBLAN fiber laser at 3.005 μm operating in the quasi‐CW mode (5 ms, 20 Hz). FBG, fiber Bragg grating.

Source: reproduced from figure 9 of [32], with permission of SPIE.

3.3 Q‐switched Mid‐IR Fiber Lasers

Mid‐IR fiber lasers operating in the Q‐switched mode with microsecond to nanosecond pulse duration are desirable for such applications as light detection and ranging (LIDAR), pumping longer‐wavelength mid‐IR optical parametric oscillations (OPOs), supercontinuum generation, polymer processing, and for laser surgery where efficient and painless tissue removal can be achieved due to fast energy deposition.

In the 2‐μm region, both Tm‐ and Ho‐fiber lasers are suitable for pulsed‐mode operation. Eichhorn and Jackson demonstrated a Q‐switched Tm3+‐doped fiber laser providing multi‐watt average power at λ = 1.98 μm [34]. The silica double‐clad fiber was 2.3 m long with a core diameter of 20 μm and was diode‐pumped at 792 nm. The laser was Q‐switched with an acousto‐optic modulator (AOM) and generated up to 30 W of average power at 110 kHz repetition rate, with 270 μJ pulse energy, pulse duration 41 ns, and peak power 6.6 kW. The maximum power, according to the authors, was limited by the buildup of amplified spontaneous emission (ASE). With a diffraction grating in the resonator, the laser was tuned from 1.93 to 2.05 μm with a pulse width of 41–50 ns at 125 kHz repetition rate [34]. Kadwani et al. reported an acousto‐optically Q‐switched oscillator based on Tm‐doped photonic crystal fiber (PCF). The pump was a 100‐W diode laser at 793 nm, and the active fiber had a large mode‐field area >1000 μm2. The laser maintained single‐mode beam quality in a polarized beam and provided 435‐μJ, 49‐ns pulses at 10 kHz repetition rate at 2 μm wavelength, corresponding to an average power of 4.4 W and peak power as high as 8.9 kW [35].

A pulsed‐mode operation with an acousto‐optic Q‐switching of a Tm3+, Ho3+‐codoped double‐clad silica fiber laser was reported by Eichhorn and Jackson [36]. Here, thulium provides efficient excitation using standard 790‐nm laser diodes, while a resonant energy transfer process efficiently populates the Ho3+ ions. The fiber was symmetrically diode‐pumped from two sides at 792 nm, with up to 30 W per each side. With the use of an intracavity diffraction grating and by taking advantage of both Tm3+ and Ho3+ emission bands, the laser was tuned in the range 1.95–2.13 μm. In the Ho3+ band, the shortest pulses were reached at λ = 2.07 μm at 20 kHz repetition frequency with pulse duration of 58 ns and pulse energy of 250 μJ, corresponding to an average power of 5 W. In the Tm3+ emission band, at λ = 2.02 μm, the maximum average power was 15 W at 100 kHz repetition frequency (pulse energy 150 μJ) [36]. A high average power nanosecond system based on thulium‐doped fiber amplifier with 793‐nm diode pump was achieved in a two‐stage oscillator–amplifier setup [37]. The pulses with variable duration between 4 and 72 ns were generated by mode‐locking a long‐cavity low repetition rate oscillator with the aid of a nonlinear optical loop mirror. At 1.07 MHz repetition frequency and 72 ns pulse duration, 100.4 W of average power was produced at λ ≈ 2 μm.

A number of groups achieved Q‐switched fiber laser operation in the 3‐μm band. Hu et al. demonstrated a Q‐switched laser based on codoped Ho3+, Pr3+ fiber operating at 2.87 μm. (Codoping Ho with Pr ions allows de‐excitation of the 5I7 state in holmium, providing high slope efficiencies.) With an acousto‐optical Q‐switching (Figure 3.11), the authors achieved 78‐ns‐long pulses at 120 kHz, with an average power of 0.72 W and slope efficiency of 20% with respect to the launched pump power at 1150 nm [38].

Tokita et al. demonstrated an impressive result in terms of 3‐μm fiber laser peak and average power [39]. A diode‐pumped 2.8 μm laser oscillator, based on 35‐μm‐core Er‐doped ZBLAN fiber, was Q‐switched with an AOM. At a repetition rate of 120 kHz and a maximum 975‐nm diode pump power of 75 W, the pulse energy reached 100 μJ at a pulse duration of 90 ns, corresponding to the average output power of 12.4 W (0.9 kW peak).

Compared to actively Q‐switched fiber lasers, passively Q‐switched lasers have the advantages of low cost and compactness. Several passively Q‐switched fiber lasers around 3 μm with ~μs pulse duration have been reported recently, using different saturable absorbers: graphene [40, 41], Fe2+:ZnSe [41], semiconductor saturable absorber mirror (SESAM) based on indium arsenide (InAs) [42], and using a topological insulator Bi2Te3 [43]. (Topological insulators represent a novel class of so‐called Dirac materials. In the bulk state, they have a narrow band gap while at the surface they are in the gapless metallic state and saturable absorption is observed in a broadband spectral range via the Pauli blocking effect.) The average output power in [42, 43] exceeded 300 mW at repetition frequencies around 50–100 kHz (see Table 3.1).

Schematic of a 2.87‐μm acousto‐optically Q‐switched Ho3+Pr3+ fiber laser. Pump diodes, PBS, dichroic mirror, AOM, etc. are labeled.

Figure 3.11 Schematic of a 2.87‐μm acousto‐optically Q‐switched Ho3+Pr3+ fiber laser.

Source: reproduced from figure 1 of [38], with permission of OSA, The Optical Society.

3.4 Mode‐locked Mid‐IR Fiber Lasers

Fiber‐based mode‐locked lasers are desirable sources of ultrashort pulses because of their compactness and environmental reliability, as has already been demonstrated by their near‐IR counterparts. Comparatively large gain bandwidths of fiber lasers, due to the amorphous nature of the glass host, typically allow achieving sub‐100‐fs pulse durations. Mode‐locked lasers in the mid‐IR can further fuel the ultrafast laser science and offer a wide variety of promising applications, such as generation of supercontinuum and optical frequency combs, and precision laser surgery, to name a few.

Ultrafast fiber lasers are usually based on passively mode‐locked oscillators. While erbium at λ = 1.55 μm and ytterbium at λ = 1.05 μm silica fiber lasers still dominate in the commercial realm, enormous progress took place recently in the development of thulium (Tm) and holmium (Ho) doped fiber lasers emitting near 2.0 μm and Er‐fluoride fiber lasers emitting near 3.0 μm [52].

Tm‐doped silica‐fiber‐based 2‐μm lasers are the most mature among the mid‐IR mode‐locked fiber lasers. In the early stage of their development, passively mode‐locked femtosecond Tm‐doped fiber oscillators were demonstrated in the mid‐1990s. For example, Nelson et al. [53] used additive‐pulse mode‐locking based on nonlinear polarization evolution to obtain tunable 1.8–1.9 μm pulses in the soliton regime, whereas Sharp et al. [54] reported a mode‐locked thulium‐doped silica fiber laser using an InGaAs saturable absorber. Later, watt‐level femtosecond Tm‐fiber lasers were developed. For example, by amplification of a Raman‐shifted output of a 1.56‐μm Er‐doped fiber laser in a Tm‐doped fiber having a 25‐μm‐diameter core, Imeshev and Fermann obtained 108‐fs pulses at 2 μm wavelength with an average power as high as 3.1 W, pulse energy of 31 nJ, and peak power of 230 kW [44]. Subsequently, stable 2‐μm Tm‐fiber‐based oscillators were developed using SESAM combined with nonlinear polarization evolution for mode‐locking mechanism. After amplification in a dispersion‐compensated Tm‐fiber amplifier, high‐power (>2.5 W) sub‐70‐fs pulses at ~2 μm were generated [45, 55]. These developments have also led to creating fully stabilized optically referenced frequency combs discussed in Chapter 6.

Table 3.1 Summary of mid‐IR fiber lasers.

Dopant/fiber Wavelength (μm) Laser characteristics Ref.
Continuous wave, CW
Tm3+‐silica 2.045 Ave. power 1 kW, pump at 790 nm [8]
Tm3+, Ho3+‐silica 2.105 Ave. power 83 W, pump at 793 nm [10]
Ho3+‐ZBLAN 3
2.1
Cascaded laser, ave. power 0.77 W at 3 μm, and 0.24 W at 2.1 μm, pump at 1150 nm [11]
Ho3+, Pr3+‐ZBLAN 2.94 Ave. power 2.5 W, pump at 1150 nm [12]
Er3+‐ZBLAN 2.8 Ave. power 24 W, pump at 975 nm, liquid fiber cooling [17]
Er3+‐fluoride glass 2.825 Ave. power 20.6 W, pump at 980 nm (90 W) [19]
Er3+‐fluoride glass 2.825 Ave. power 12.5 W, pump at 976 nm (33.5 W), cascade 1.6 μm, slope optical efficiency 50% [20]
Er3+‐fluoride glass 2.94 Ave. power 30.5 W, pump at 980 nm (188 W) [21]
Er3+‐fluoride glass 3.44 Ave. power 1.5 W, dual pump: at 974 and 1976 nm [22]
Er3+‐fluoride glass 3.55 Ave. power 5.6 W, dual pump: at 974 and 1976 nm, total optical efficiency 26.4% [23]
Er3+‐ZBLAN 3.33–3.78 Grating‐tunable, ave. power 1.45 W @ 3.47 μm dual pump: at 977 and 1973 nm [24]
Dy3+‐ZBLAN 3.04 Ave. power 80 mW, pump at 2.8 μm (300 mW) [25]
Raman lasers
Fluoride glass 2.23 Ave. power 3.7 W, nested cavity, pump at 1.98 μm [29]
As2S3 glass 3.34 Quasi‐CW mode (5 ms, 20 Hz), ave. power 50 mW, pump at λ = 3.005 μm (250 mW) [31]
As2S3 glass 3.77 Quasi‐CW mode (5 ms, 20 Hz), ave. power 9 mW, pump at λ = 3.005 μm (371 mW) [33]
Q‐switched
Tm3+‐silica 1.98 Ave. power 30 W (peak 6.6 kW), AOM Q‐sw, 41 ns, 270 μJ, 110 kHz [34]
Tm3+‐silica 1.95 Ave. power 4.4 W (peak 8.9 kW), AOM Q‐sw, 49 ns, 435 μJ, 10 kHz [35]
Tm3+, Ho3+‐silica 2.07 Ave. power 5 W, AOM Q‐sw, 58 ns, 250 μJ, 20 kHz [34]
Tm3+‐silica ~2 Ave. power 100 W, nonlinear optical loop mirror, oscillator–amplifier, 72 ns, 94 μJ, 1 MHz [37]
Ho3+, Pr3+‐ZBLAN 2.87 Ave. power 0.72 W, AOM Q‐sw, 78 ns, 6 μJ, 120 kHz [38]
Er3+‐ZBLAN 2.8 Ave. power 12.4 W, AOM Q‐sw, 90 ns, 103 μJ, 120 kHz [39]
Ho3+‐ZBLAN 2.97 Ave. power 317 mW, SESAM, 1.68 μs, 6.65 μJ, 48 kHz [42]
Ho3+‐ZBLAN 2.98 Ave. power 327 mW, Bi2Te3 topological insulator, 1.37 μs, 4 μJ, 82 kHz [43]
Mode‐locked
Tm3+‐silica 1.98 Ave. power 3.1 W, Raman‐shifted Er‐laser and Tm‐amplifier, 108 fs, 100 MHz [44]
Tm3+‐silica 1.94–1.97 Ave. power 2.5 W, SESAM mode‐locked oscillator, 100 fs, 418 MHz [45]
Er3+‐ZBLAN 2.8 Ave. power 440 mW, SESAM and a fiber Bragg grating, 60 ps, 52 MHz [46]
Ho3+, Pr3+‐ZBLAN 2.9 Ave. power 70 mW, InAs saturable absorber, 6 ps, 25 MHz [47]
Er3+‐ZBLAN 2.8 Ave. power 206 mW, nonlinear polarization rotation, 497 fs, 57 MHz, peak 6.4 kW [48]
Er3+‐ZBLAN 2.8 Ave. power 40 mW, nonlinear polarization rotation, 207 fs, 55 MHz, peak 3.5 kW [49]
Er3+‐ZBLAN 2.8 Ave. power 676 mW, nonlinear polarization rotation, 270 fs, 97 MHz [50]
Er3+‐ZBLAN 2.8–3.6 Ave. power ~2 W at 3.4 μm, Raman self‐frequency shifted solitons, 160 fs, pump at 2.8 μm [51]

A mode‐locked soliton fiber laser based on Ho3+‐doped glass (λ = 2.08 μm) was demonstrated with graphene as a saturable absorber [56]. The laser was pumped by a CW Tm‐fiber laser (0.5 W at 1950 nm) and produced 811‐fs duration pulses at 34 MHz repetition rate and 44 mW average power.

Based on a variety of saturable absorbers, passively mode‐locked lasers in the 3‐μm range were also reported, with pulse durations ~20 ps and average power ~100 mW. These include an Er3+‐doped ZBLAN fiber laser at 2.8 μm, passively mode‐locked by a Fe2+:ZnSe crystal inside the laser cavity [57] and Ho3+, Pr3+ codoped ZBLAN fiber laser at 2.87 μm, which used an InAs‐based SESAM as passive mode‐locker [58]. Yet, the above lasers exhibited “pulsing” – both Q‐switching and mode‐locking behavior simultaneously.

Haboucha et al. reported a truly continuous mode‐locking from an Er3+‐ZBLAN glass fiber laser operating at 2.8 μm by using SESAM in conjunction with a fiber Bragg grating inside a linear cavity [46]. The Bragg grating allowed for the generation of a stable, self‐starting pulse train with a repetition rate of 52 MHz, pulse duration of 60 ps, and an average power of 440 mW. In parallel, Hu et al. demonstrated a stable mode‐locked fiber ring laser near 2.9 μm based on a Ho3+, Pr3+‐codoped ZBLAN fiber using an InAs saturable absorber. This configuration enabled the generation of 6‐ps pulses at 25 MHz repetition rate with an average power of 70 mW [47].

Most recently, two groups have independently demonstrated the possibility of generating femtosecond pulses from an Er3+‐doped fluoride (ZBLAN) fiber at 2.8 μm based on nonlinear polarization rotation [48, 49]. This technique relies upon the high Kerr nonlinearity of the fluoride fiber (n2 = 2.1 × 10−20 m2/W) to create an intensity‐dependent rotation of the polarization state inside the fiber, which, when combined with polarizing optics (quarter wave plate, half wave plate, and an optical isolator), makes an effective saturable absorber for mode‐locking. Hu and colleagues were able to produce 497‐fs pulses at 2.8 μm with 206 mW average power at 57 MHz repetition rate (Figure 3.12) [48]. Similarly, Duval et al. [49] reported a passively mode‐locked Er‐ZBLAN fiber ring laser based on nonlinear polarization evolution. The laser generated 207 fs pulses at λ = 2.8 μm with a repetition rate of 55 MHz at an average power of 40 mW. Since ZBLAN fiber has anomalous dispersion at 2.8 μm, both of these lasers were reported to operate in the soliton regime. By varying the Er3+:fluoride fiber length and the output coupling, Duval et al. were able to scale up the 2.8‐μm mode‐locked laser power. They demonstrated stable 270‐fs pulses with the repetition rate of 96.6 MHz and the average power of 675 mW (pulse energy 7 nJ, peak power 23 kW) [50].

The same team reported on a tunable (2.8–3.6 μm) and high‐power (>1 W) ultrafast fiber laser system based on erbium‐doped zirconium fluoride glass [51]. Tuning of the central wavelength of the pulses was achieved through a process known as soliton Raman self‐frequency shift, which is facilitated by the anomalous dispersion of fluoride glass fibers in this spectral range. An ultrafast fiber oscillator (the pump) operating at λ = 2.8 μm and a repetition rate of 58 MHz was similar to the one described in [50]. It was followed by a second erbium‐doped fluoride glass fiber, a portion of which (L = 1.25 m) served as an amplifier, while the remaining unpumped segment (L = 8 or 22 m) acted as a passive fiber to shift the amplified soliton to longer wavelengths. As the amplifier pump power was increased, the spectrum of a clean and isolated soliton continuously shifted toward longer wavelengths in the passive portion of the fiber, reaching a maximum central wavelength of 3.4 μm in the 8‐m fiber and 3.59 μm in the 22‐m fiber. Overall, high‐energy Raman soliton pulses tunable from 2.8 to 3.6 μm were generated. For example, at λ = 3.4 μm, 160‐fs pulses were produced with the pulse energy of 37 nJ (200 kW peak), corresponding to an average output power >2 W [51].

Experimental setup for generating mode‐locked pulses from an Er3+ ZBLAN fiber laser at 2.8 μm. RF analyzer, oscilloscope, spectrometer, PBS, isolator, erbium-doped ZBLAN fiber, etc. are labeled.

Figure 3.12 Experimental setup for generating mode‐locked pulses from an Er3+ ZBLAN fiber laser at 2.8 μm. A ring cavity is formed with a free‐space section containing polarization optics to provide the nonlinear polarization evolution mode‐locking. The pulses were analyzed in the time domain, using a mid‐IR frequency‐resolved optical gating (FROG) setup.

Source: reproduced from figure 1 of [48], with permission of OSA, The Optical Society.

The main results for mid‐IR fiber lasers are presented in Table 3.1.

3.5 Summary

So far, the most successful fiber lasers in the mid‐IR are based on the Tm3+ (3H4 → 3H6) laser transition, the Er3+ (4I11/2 → 4I13/2) transition, and the Ho3+ (5I6 → 5I7) transition. They cover, respectively, 1.9–2.1, 2.7–3.0, and 2.8–3.05 μm spectral ranges.

Other wavelengths can be reached via dual‐wavelength pumping (Er3+, 3.44–3.55 μm), Raman effect (chalcogenide glass, 3.34 μm), or via cryogenic cooling (Ho3+, 3.9 μm).

In terms of the output power, Tm‐doped fibers enable getting 1‐kW CW power in a single‐spatial mode at 2.045 μm with optical‐to‐optical conversion efficiency higher than that dictated by the quantum defect, thanks to the “two‐for‐one” cross‐relaxation process. At longer wavelengths, achieving high‐power operation is more challenging. The main cause of power decline is the increase of quantum defect since the most efficient high‐power laser diodes typically emit in the near‐IR. Thus, there is more energy dissipation, and created heat becomes an ever‐increasing fraction of the absorbed pump energy as the wavelength increases. Nevertheless, based on the 3‐μm transition in erbium, benefiting from “two‐for‐one” cross‐relaxation process, more than 30 W of average power has been obtained at 2.94 μm with passively cooled Er3+‐doped fluoride glass fiber.

Mid‐IR fiber lasers operating in the CW, Q‐switched, and mode‐locked modes open up a plethora of new applications in biomedicine, frequency combs generation, and in nonlinear optics and spectroscopy. For further reading, excellent reviews on mid‐IR fiber lasers can be found in [1, 32].

References

  1. 1 Jackson, S.D. (2012). Towards high‐power mid‐infrared emission from a fibre laser. Nat. Photonics 6: 423.
  2. 2 Hanna, D.C., Percival, R.M., Smart, R.G., and Tropper, A.C. (1990). Efficient and tunable operation of a Tm‐doped fiber laser. Opt. Commun. 75: 283.
  3. 3 Jackson, S.D. and King, T.A. (1998). High‐power diode‐cladding‐pumped Tm‐doped silica fiber laser. Opt. Lett. 23: 1462.
  4. 4 Hayward, R.A., Clarkson, W.A., Turner, P.W., Nilsson, J., Grudinin, A.B., and Hanna, D.C. (2000). Efficient cladding‐pumped Tm‐doped silica fiber laser with high power singlemode output at 2 μm. Electron. Lett. 36: 711.
  5. 5 Moulton, P.F., Rines, G.A., Slobodtchikov, E.V., Wall, K.F., Frith, G., Samson, B., and Carter, A.L.G. (2009). Tm‐doped fiber lasers: fundamentals and power scaling. IEEE J. Sel. Top. Quant. Electron. 15: 85.
  6. 6 Clarkson, W.A., Barnes, N.P., Turner, P.W., Nilsson, J., and Hanna, D.C. (2002). High‐power cladding‐pumped Tm‐doped silica fiber laser with wavelength tuning from 1860 to 2090 nm. Opt. Lett. 27: 1989.
  7. 7 Jackson, S.D. (2004). Cross relaxation and energy transfer upconversion processes relevant to the functioning of 2 μm Tm3+‐doped silica fibre lasers. Opt. Commun. 230: 197.
  8. 8 Ehrenreich, T., Leveille, R., Majid, I., Tankala, K., Rines, G., and Moulton, P. (2010). 1 kW, all‐glass Tm:fiber laser. Proc. SPIE 7580: 7580–7112.
  9. 9 Jackson, S.D. (2006). Midinfrared holmium fiber lasers. IEEE J. Quant. Electron. 42: 187.
  10. 10 Jackson, S.D., Sabella, A., Hemming, A., Bennetts, S., and Lancaster, D.G. (2007). High‐power 83 W holmium‐doped silica fiber laser operating with high beam quality. Opt. Lett. 32: 241.
  11. 11 Li, J., Hudson, D.D., and Jackson, S.D. (2011). High‐power diode‐pumped fiber laser operating at 3 μm. Opt. Lett. 36: 3642.
  12. 12 Jackson, S.D. (2009). High‐power and highly efficient diode‐cladding pumped holmium‐doped fluoride fiber laser operating at 2.94 μm. Opt. Lett. 34: 2327.
  13. 13 Schneide, J., Carbonnier, C., and Unrau, U.B. (1997). Characterization of a Ho3+‐doped fluoride fiber laser with a 3.9 μm emission wavelength. Appl. Opt. 36: 8595.
  14. 14 Faucher, D., Bernier, M., Caron, N., and Vallée, R. (2009). Erbium‐doped all‐fiber laser at 2.94 μm. Opt. Lett. 34: 3313.
  15. 15 Jackson, S.D., King, T.A., and Pollnau, M. (1999). Diode‐pumped 1.7‐W erbium 3‐μm fiber laser. Opt. Lett. 24: 1133.
  16. 16 Zhu, X. and Jain, R. (2007). 10‐W‐level diode‐pumped compact 2.78 μm ZBLAN fiber laser. Opt. Lett. 32: 26.
  17. 17 Tokita, S., Murakami, M., Shimizu, S., Hashida, M., and Sakabe, S. (2009). Liquid‐cooled 24 W mid‐infrared Er:ZBLAN fiber laser. Opt. Lett. 34: 3062.
  18. 18 Tokita, S., Hirokane, M., Murakami, M., Shimizu, S., Hashida, M., and Sakabe, S. (2010). Stable 10 W Er:ZBLAN fiber laser operating at 2.71–2.88 μm. Opt. Lett. 35: 3943.
  19. 19 Faucher, D., Bernier, M., Androz, G., Caron, N., and Vallée, R. (2011). 20 W passively cooled single‐mode all‐fiber laser at 2.8 μm. Opt. Lett. 36: 1104.
  20. 20 Aydin, Y.O., Fortin, V., Maes, F., Jobin, F., Jackson, S.D., Vallée, R., and Bernier, M. (2017). Diode‐pumped mid‐infrared fiber laser with 50% slope efficiency. Optica 4: 235.
  21. 21 Fortin, V., Bernier, M., Bah, S.T., and Vallée, R. (2015). 30 W fluoride glass all‐fiber laser at 2.94 μm. Opt. Lett. 40: 2882.
  22. 22 Fortin, V., Maes, F., Bernier, M., Bah, S.T., D’Auteuil, M., and Vallée, R. (2016). Watt‐level erbium‐doped all‐fiber laser at 3.44 μm. Opt. Lett. 41: 559.
  23. 23 Maes, F., Fortin, V., Bernier, M., and Vallée, R. (2017). 5.6 W monolithic fiber laser at 3.55 μm. Opt. Lett. 42: 2054.
  24. 24 Henderson‐Sapir, O., Jackson, S.D., and Ottaway, D.J. (2016). Versatile and widely tunable mid‐infrared erbium doped ZBLAN fiber laser. Opt. Lett. 41: 1676.
  25. 25 Majewski, M.R. and Jackson, S.D. (2016). Highly efficient mid‐infrared dysprosium fiber laser. Opt. Lett. 41: 2173.
  26. 26 Dianov, E.M., Bufetov, I.A., Mashinsky, V.M., Shubin, A.V., Medvedkov, O.I., Rakitin, A.E., Melkumov, M.A., Khopin, V.F., and Gur’yanov, A.N. (2005). Raman fiber lasers based on heavily GeO2‐doped fibers. Quantum Electron. 35: 435.
  27. 27 Cumberland, B.A., Popov, S.V., Taylor, J.R., Medvedkov, O.I., Vasiliev, S.A., and Dianov, E.M. (2007). 2.1 μm continuous‐wave Raman laser in GeO2 fiber. Opt. Lett. 32: 1848.
  28. 28 Fortin, V., Bernier, M., Carrier, J., and Vallée, R. (2011). Fluoride glass Raman fiber laser at 2185 nm. Opt. Lett. 36: 4152.
  29. 29 Fortin, V., Bernier, M., Faucher, D., Carrier, J., and Vallée, R. (2012). 3.7 W fluoride glass Raman fiber laser operating at 2231 nm. Opt. Express 20: 19412.
  30. 30 Jackson, S.D. and Anzueto‐Sánchez, G. (2006). Chalcogenide glass Raman fiber laser. Appl. Phys. Lett. 88: 221106.
  31. 31 Bernier, M., Fortin, V., Caron, N., El‐Amraoui, M., Messaddeq, Y., and Vallée, R. (2013). Mid‐infrared chalcogenide glass Raman fiber laser. Opt. Lett. 38: 127.
  32. 32 Fortin, V., Bernier, M., Caron, N., Faucher, D., El Amraoui, M., Messaddeq, Y., and Vallée, R. (2013). Towards the development of fiber lasers for the 2 to 4 μm spectral region. Opt. Eng. 52: 054202.
  33. 33 Bernier, M., Fortin, V., El‐Amraoui, M., Messaddeq, Y., and Vallée, R. (2014). 3.77 μm fiber laser based on cascaded Raman gain in a chalcogenide glass fiber. Opt. Lett. 39: 2052.
  34. 34 Eichhorn, M. and Jackson, S.D. (2007). High‐pulse‐energy actively Q‐switched Tm3+‐doped silica 2 μm fiber laser pumped at 792 nm. Opt. Lett. 32: 2780.
  35. 35 Kadwani, P., Modsching, N., Sims, R.A., Leick, L., Broeng, J., Shah, L., and Richardson, M. (2012). Q‐switched thulium‐doped photonic crystal fiber laser. Opt. Lett. 37: 1664.
  36. 36 Eichhorn, M. and Jackson, S.D. (2008). High‐pulse‐energy, actively Q‐switched Tm3+,Ho3+‐codoped silica 2‐μm fiber laser. Opt. Lett. 33: 1044.
  37. 37 Zhao, J., Ouyang, D., Zheng, Z., Liu, M., Ren, X., Li, C., Ruan, S., and Xie, W. (2016). 100 W dissipative soliton resonances from a thulium‐doped double‐clad all‐fiber‐format MOPA system. Opt. Express 24: 12072.
  38. 38 Hu, T., Hudson, D.D., and Jackson, S.D. (2012). Actively ?‐switched 2.9 μm Ho3+Pr3+‐doped fluoride fiber laser. Opt. Lett. 37: 2145.
  39. 39 Tokita, S., Murakami, M., Shimizu, S., Hashida, M., and Sakabe, S. (2011). 12 W Q‐switched Er:ZBLAN fiber laser at 2.8 μm. Opt. Lett. 36: 2812.
  40. 40 Wei, C., Zhu, X., Wang, F., Xu, Y., Balakrishnan, K., Song, F., Norwood, R.A., and Peyghambarian, N. (2013). Graphene Q‐switched 2.78 μm Er3+‐doped fluoride fiber laser. Opt. Lett. 38: 3233.
  41. 41 Zhu, G., Zhu, X., Balakrishnan, K., Norwood, R.A., and Peyghambarian, N. (2013). Fe2+:ZnSe and graphene Q‐switched singly Ho3+‐doped ZBLAN fiber lasers at 3 μm. Opt. Mater. Express 3: 1365.
  42. 42 Li, J.F., Luo, H.Y., He, Y.L., Liu, Y., Zhang, L., Zhou, K.M., Rozhin, A.G., and Turistyn, S.K. (2014). Semiconductor saturable absorber mirror passively Q‐switched 2.97 μm fluoride fiber laser. Laser Phys. Lett. 11: 065102.
  43. 43 Li, J., Luo, H., Wang, L., Zhao, C., Zhang, H., Li, H., and Liu, Y. (2015). 3‐μm mid‐infrared pulse generation using topological insulator as the saturable absorber. Opt. Lett. 40: 3659.
  44. 44 Imeshev, G. and Fermann, M.E. (2005). 230‐kW peak power femtosecond pulses from a high power tunable source based on amplification in Tm‐doped fiber. Opt. Express 13: 7424.
  45. 45 Lee, K.F., Mohr, C., Jiang, J., Schunemann, P.G., Vodopyanov, K.L., and Fermann, M.E. (2015). Midinfrared frequency comb from self‐stable degenerate GaAs optical parametric oscillator. Opt. Express 23: 26596.
  46. 46 Haboucha, A., Fortin, V., Bernier, M., Genest, J., Messaddeq, Y., and Vallée, R. (2014). Fiber Bragg grating stabilization of a passively mode‐locked 2.8 μm Er3+: fluoride glass fiber laser. Opt. Lett. 39: 3294.
  47. 47 Hu, T., Hudson, D.D., and Jackson, S.D. (2014). Stable, self‐starting, passively mode‐locked fiber ring laser of the 3 μm class. Opt. Lett. 39: 2133.
  48. 48 Hu, T., Jackson, S.D., and Hudson, D.D. (2015). Ultrafast pulses from a mid‐infrared fiber laser. Opt. Lett. 40: 4226.
  49. 49 Duval, S., Bernier, M., Fortin, V., Genest, J., Piché, M., and Vallée, R. (2015). Femtosecond fiber lasers reach the mid‐infrared. Optica 2: 623.
  50. 50 Duval, S., Olivier, M., Fortin, V., Bernier, M., Piché, M., and Vallée, R. (2016). 23‐kW peak power femtosecond pulses from a mode‐locked fiber ring laser at 2.8 μm. Proc. SPIE 9728: 972 802.
  51. 51 Duval, S., Gauthier, J.‐C., Robichaud, L.‐R., Paradis, P., Olivier, M., Fortin, V., Bernier, M., Piché, M., and Vallée, R. (2016). Watt‐level fiber‐based femtosecond laser source tunable from 2.8 to 3.6 μm. Opt. Lett. 41: 5294.
  52. 52 Fermann, M.E. and Hartl, I. (2013). Ultrafast fibre lasers. Nat. Photonics 7: 868.
  53. 53 Nelson, L.E., Ippen, E.P., and Haus, H.A. (1995). Broadly tunable sub‐500 fs pulses from an additive‐pulse mode‐locked thulium‐doped fiber ring laser. Appl. Phys. Lett. 67: 19.
  54. 54 Sharp, R.C., Spock, D.E., Pan, N., and Elliot, J. (1996). 190‐fs passively mode‐locked thulium fiber laser with a low threshold. Opt. Lett. 21: 881.
  55. 55 Bethge, J., Jiang, J., Mohr, C., Fermann, M., and Hartl, I. (2012). Optically referenced Tm‐fiber‐laser frequency comb, Technical Digest. In: Advanced Solid State Photonics. Washington, DC: Optical Society of America, Paper AT5A.3.
  56. 56 Sotor, J., Pawliszewska, M., Sobon, G., Kaczmarek, P., Przewolka, A., Pasternak, I., Cajzl, J., Peterka, P., Honzátko, P., Kašík, I., Strupinski, W., and Abramski, K. (2016). All‐fiber Ho‐doped mode‐locked oscillator based on a graphene saturable absorber. Opt. Lett. 41: 2592.
  57. 57 Wei, C., Zhu, X., Norwood, R.A., and Peyghambarian, N. (2012). Passively continuous‐wave mode‐locked Er3+‐doped ZBLAN fiber laser at 2.8 μm. Opt. Lett. 37: 3849.
  58. 58 Li, J., Hudson, D.D., Liu, Y., and Jackson, S.D. (2012). Efficient 2.87 μm fiber laser passively switched using a semiconductor saturable absorber mirror. Opt. Lett. 37: 3747.

Note

  1. 1 http://www.ipgphotonics.com.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
13.59.82.167