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

Band-gap modulation of two-dimensional saturable absorbers for solid-state lasers

Open Access Open Access

Abstract

Due to the manifestation of fascinating physical phenomena and materials science, two-dimensional (2D) materials have recently attracted enormous research interest with respect to the fields of electronics and optoelectronics. There have been in-depth investigations of the nonlinear properties with respect to saturable absorption, and many 2D materials show potential application in optical switches for passive pulsed lasers. However, the Eigen band-gap determines the responding wavelength band and constrains the applications. In this paper, based on band-gap engineering, some different types of 2D broadband saturable absorbers are reviewed in detail, including molybdenum disulfide (MoS2), vanadium dioxide (VO2), graphene, and the Bi2Se3 topological insulator. The results suggest that the band-gap modification should play important roles in 2D broadband saturable materials and can provide some inspiration for the exploration and design of 2D nanodevices.

© 2015 Chinese Laser Press

1. INTRODUCTION

Lasers, since first demonstrated by a ruby crystal in 1960 [1], have shown tremendous progress so far. On account of the advantages of large pulse energy, short pulse duration, and high peak power, pulse lasers, including mode-locking and Q-switching, receive more widespread attention [2,3]. Pulsed lasers with the pulse width from nanosecond to subpicosecond scales play a key role in the industrial machining, remote sensing, and military defense areas [4]. Ultrashort femtosecond lasers have significant applications in the fields of the in-situ detection, complex reaction dynamics, medical survey, and telecommunication [3]. Passive pulse modulation could be operated with compact and cost-effective structures, such that it has become an influential research direction in the laser field. Taking advantage of saturable absorption of saturable absorbers, that is part of the third-order nonlinear properties, passive pulse modulation in the Q-switched or mode-locking lasers could be easily implemented. Traditional saturable absorbers, such as ion-doped crystals (Cr:YAG, Cr:ZnSe, V:YAG, and so on) [57], GaAs [8], and color-center crystals (for example F2:LiF) [9], are wavelength-sensitive and are only applied in their respective corresponding wavelength bands. Although the commercial semiconductor saturable absorber mirrors (SESAMs) could be fabricated for the specific wavelength from 0.4 to 2.5 μm by using different semiconductor material systems [2,10], the relatively complex structural design may increase the product cost and the difficulty of preparation. In addition, the quasi-phase-matching technique in optical superlattice, which is feasible for a wide wavelength range as long as it is in the transparent spectral region of the optical superlattice material, is also a promising method to obtain passive mode-locking lasers [1113]. Nevertheless, this technique requires the high-efficiency generation of second-harmonic generation and elaborate design of the resonant cavity and the second-harmonic crystal. Therefore, the development of low-cost and robust broadband saturable absorbers is still an intense focus of research nowadays [2].

Graphene, as the representative of the two-dimensional (2D) materials, since found by K. Geim and K. S. Novoselov, has profoundly promoted and broadened the research areas of 2D materials in the fields of electronics and optoelectronics due to an abundance of fantastic physics [1423]. With in-depth studies, many other novel 2D materials are discovered and are brought into people’s horizons, such as transition metal dichalcogenides (TMDCs) (typically MoS2, MoSe2, WS2, and so on) [16,17,19,24,25], transition metal oxides (for instance MoO3 and TiO2) [17,24,26], topological insulators (Bi2Se3, Bi2Te3, and Sb2Te3) [16,17,24,27], as well as silicone [28], germanane [16,28], black phosphorus [29,30], and other graphene analogues (typically h-BN) [19,24]. For the major 2D materials, the layered structure results in the strong in-plane coupling and the weak van der Waals coupling between layers. Therefore monolayer or few-layer 2D perfect samples could be easily fabricated by mechanical exfoliation or chemical exfoliation [18,31]. In addition, aiming at the growth habits and atomic arrangements for different 2D materials, many more efforts are made to explore the economical and practical growth methods, such as chemical vapor deposition, and epitaxial growth on adaptive substrates [21,25]. The layered 2D materials exhibit numerous exotic physical, chemical, and mechanical properties such that they are rapidly developed. Many potential functional applications in terms of 2D materials in some aspects will be realized in the near future, including optical modulators, photodetectors, logic transistors, high-frequency transistors, energy storage, and sensor devices [1618,24,31].

Most 2D semiconductor materials show a simple two energy band structure of the conduction band and the valence band. Light that is of higher energy than the gap energy can excite carriers from the valance band to the conduction band. If the excitation has stronger intensity, all possible initial states are depleted and the final states are partially occupied in accordance with the Pauli blocking effect such that the absorption will be saturated [2]. Thus the 2D semiconductor materials give some opportunities for the fabrication of cost-effective and flexible broadband saturable absorbers, and some 2D materials have realized this goal [4,3234]. Saturable absorption is the modulation of nonlinear absorption. However, the bandgap of a semiconductor determines the responding wavelength and every semiconductor has its specialized bandgap. Based on the photoelectric effect, the semiconductors materials with a narrow bandgap have broad responding wavelength bands. Therefore, besides the narrow bandgap, for 2D materials such as graphene, the modulation of the bandgap should be important, especially for the materials with large bandgaps such as MoS2 [13,35]. In this review, what is discussed will mainly concentrate on the state-of-the-art research exploration and development about the saturable absorbers for several 2D materials. The content consists of the band-gap modification by impurity of MoS2, band-gap modulation by phase transition of VO2, and broadband modulation by inherent bandgaps for graphene and Bi2Se3 topological insulator. This work can show some guiding designing functions for the investigation of other 2D layered optoelectronic materials.

2. BAND-GAP ENGINEERING FOR MoS2

Layered MoS2, a typical representative of the layered TMDCs family, is composed of the sandwiches combined by van der Waals interactions, and each sandwich consists of covalently bound S–Mo–S trilayers with two hexagonal layers of S atoms and an intermediate hexagonal plane of Mo atoms [17,25,3638]. The monolayer MoS2 has a direct band-gap at the K point of the Brillouin zone with a gap of 1.8 eV (0.7 μm), and the bulk MoS2 is of an indirect band-gap between the Γ point (valence band) and K point (conduction band) of the Brillouin zone with a gap of ranging from 0.86 eV (1.4 μm) to 1.29 eV (1.0 μm) [3944]. Besides the change of thickness, the bandgap of MoS2 was also modified by the other engineering technologies, such as the external electric field and the external strain [4548]. Atomic-layered MoS2 exhibits intriguing physical properties distinct from its bulk state, for example, the dramatical increasing of luminescence quantum efficiency for MoS2 from the bulk state to single layer. Therefore, layered MoS2 attracts more interest in the fabrication of ultrathin and flexible nanoelectronic devices [16,25], and was brought into various research fields, such as lubrication, hydrodesulfurization catalysis, supercapacitors, and field-effect transistors [3638,49].

Monolayer MoS2 has been proven with strong optical response, such as enhanced photon absorption and large photocurrent caused by van Hove singularities or band nesting [50,51]. In 2012, using the spatially and temporally resolved pump-probe technique with a 390 nm pump source and a 660 nm probe source, R. Wang et al. measured and discussed the ultrafast carrier dynamics of atomically thin MoS2 [52]. One year later, more nonlinear investigations were carried out [5355], and it is significant that ultrafast saturable absorption of MoS2 nanosheets was demonstrated around 800 nm by K. P. Wang et al. [56].

Most of the studies were in pursuit of the high-quality layered MoS2 samples in order to obtain remarkable electronic and optical properties [3739]; however, the generation of defects, which deviate from a perfect crystal structure, was inevitable in the preparation of MoS2 samples. The probable defects could localize electronic states and change the energy level, leading to some fantastic physical phenomena, such as the Mott transition and Anderson localization [57].

Recently, S. X. Wang et al. systematically analyzed the band-gap change of multilayer MoS2 in a theoretical analysis with the ratio (R) between Mo and S atoms slightly deviating from 1:2 [34]. The first-principle theoretical calculations were performed by using the plane-wave basis Vienna ab initio simulation package [58,59], and the atomic arrangement of MoS2 was the universal AB stacking (2HMoS2). As shown in Fig. 1, the bandgap is about 1.08 eV (1.2 μm) for multilayered MoS2 with the stoichiometric ratio, which corresponds well to the previous calculated value for bulk MoS2 [40]. The other results with different R values are displayed in Fig. 2. When R is larger than 2.09, MoS2 exhibits metallic character, and the generation of saturable absorption requiring suitable patterns or distances between the metal units becomes difficult. When the R is located in the range 1.89–2 and 2–2.09, few-layered MoS2 showed an indirect semiconductor property and its energy gap is 0.08 eV (15.5 μm), 0.23 eV (5.4 μm), 0.48 eV (2.6 μm), 0.63 eV (2.0 μm), and 0.26 eV (4.7 μm) for the R with a value of 2.09, 2.04, 1.97, 1.94, and 1.89, respectively. Thus, it can be concluded that the band-gap of MoS2 would be reduced by the introduction of some Mo or S defects in a suitable range. The smaller band-gap means the broadband saturable absorption of defective MoS2 becomes possible.

 figure: Fig. 1.

Fig. 1. Brillouin zone (left) and calculated band structure (blue lines, right) of bulk MoS2 with stoichiometric ratio. Selected from Ref. [34].

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Theoretical band gap of MoS2 samples; (a) AB stacked MoS2 observed from the top (left) and side (right); (b) calculated band structure of bulk MoS2 with R=12.12; (c) calculated band structure of bulk MoS2 with R=12.09; (d) calculated band structure of bulk MoS2 with R=12.04; (e) calculated band structure of bulk MoS2 with R=11.97; (f) calculated band structure of bulk MoS2 with R=11.94; (g) calculated band structure of bulk MoS2 with R=11.89. Selected from Ref. [34].

Download Full Size | PDF

Then the MoS2 samples were fabricated by the pulse laser deposition (PLD) method, which is an efficient technique to produce the S imperfections for MoS2 samples, since low-mass S is evaporated more easily than high-mass Mo [60]. The R lies in the range 1.89–1.97 moving from the center to the edge of the sample surveyed by X-ray photoelectron spectrometry. As demonstrated in Fig. 3, the measured absorbance of the MoS2 sample decreases with an increase of wavelength [34]. For the R with the value of 1.89, the energy gap for MoS2 samples is only 0.26 eV, corresponding to a wavelength of 4.7 μm. In contrast to the absorption range of standard stoichiometric MoS2, the prepared MoS2 samples are more likely to have broadband saturable absorption.

 figure: Fig. 3.

Fig. 3. Measured absorption spectrum of MoS2 sample. Selected from Ref. [34].

Download Full Size | PDF

Saturable absorption of MoS2 was investigated employing the two-level saturable absorber model widely used for 2D quantum wells [61,62]

α*=αS*1+NNS+αNS*,
N=α*Iτω,
where α*(N) is the absorption coefficient, and αS* and αNS* are the saturable and nonsaturable absorption coefficient, respectively. N is the photoinduced electron hole density, and NS, the saturation density, is the value of N for which the absorption falls to one-half of its initial value. I is the incident light intensity of a continuous-wave (cw) or long-pulse excitation source, τ is the carrier recombination time, and ω is the light frequency. Combined with the Eqs. (1) and (2), the saturable absorption equation can therefore be expressed as [32,61]
α*=αS*1+IIS+αNS*,
where the saturation intensity IS is a most important parameter to describe how easy or difficult that the photoinduced electron hole pairs are fully saturated in the saturable absorption process. With the exponential relationship between absorption and transmission, the power-dependent nonlinear transmittance T could be fitted with the following [63,64]
T=Aexp(δT1+IIS),
where T is the transmission, A is a normalization constant, and δT is the absolute modulation depth.

Saturable absorption of the as-grown MoS2 sample was measured with a picosecond pulse laser with a pulse width of 40 ps at 1.06 μm, and the data was analyzed with Eq. (4). The saturation intensity of MoS2 was calculated with 2.45GW/cm2, a value larger than that of graphene (0.610.71MW/cm2) and comparable to that of Bi2Se3 (4.3GW/cm2). The saturable carrier density Ns is about 1.4×1017cm2, a magnitude 2–3 orders larger than graphene (5.84×1013 to 8.16×1014cm2) [32,34,63].

Using the MoS2 samples as saturable absorbers, passively Q-switched lasers at wavelengths of 1.06, 1.42, and 2.1 μm were successfully operated with Nd:GdVO4, Nd:Y3Ga5O12 (Nd:YGG), and Tm:Ho:Y3Ga5O12 (Tm:Ho:YGG) crystals, respectively, as gain materials. The average output power and the repetition rate with the increase of pump power are demonstrated in Fig. 4, and the typical pulses at the wavelengths of 1.06, 1.42, and 2.1 μm are shown in Fig. 5. Q-switched lasers for the broadband MoS2 saturable absorber, ranging from 1.06 to 2.1 μm were obtained in this experiment with sub-microsecond temporal widths and maximum output powers of up to a few hundred milliwatts.

 figure: Fig. 4.

Fig. 4. Average output power and repetition rate of passively Q-switched laser; (a) passively Q-switched Nd:GdVO4 laser performance at 1.06 μm; (b) passively Q-switched Nd:YGG laser performance at 1.42 μm; (c) passively Q-switched Tm:Ho:YGG laser performance at 2.1 μm; (d) relation between the carrier density (N) and pulse repetition rate at the laser wavelength of 1.42 μm. Selected from Ref. [34].

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Passively Q-switched laser spectra and pulses; (a) passively Q-switched Nd:GdVO4, Nd:YGG, and Tm:Ho:YGG laser spectra of at center wavelengths of 1.06, 1.42, and 2.1 μm, respectively; (b) passively Q-switched Nd:GdVO4, Nd:YGG, and Tm:Ho:YGG laser spectra with the pulse width of 970, 729, and 410 ns, respectively. Selected from Ref. [34].

Download Full Size | PDF

Many research results about the pulse modulation with MoS2 saturable absorbers have been demonstrated so far. It has shown remarkable mode-locking or Q-switched performance in the different gain materials, such as Yb-doped fibers (around 1 μm) [65,66], Er-doped fibers (1.5 to 1.6 μm) [6770], Tm-doped fiber (around 2 μm) [71], Nd:GdVO4 [34], Nd:YGG [34], and Tm:Ho:YGG [34]. For the perfectly single-layered MoS2, saturable absorption is hardly acquired when the wavelength is above 0.7 μm due to the large band-gap (1.8 eV). Even if MoS2 is in the bulk state (1.08 eV), the modulation wavelength would still be not longer than 1.2 μm [34]. Employing band-gap engineering, which is similar to the fabrication of SESAMs [13], broadband pulse modulation for few-layer MoS2 was successfully obtained with a range 1.06–2.1 μm. This approach may benefit the exploitation and development of new 2D optoelectronic devices for some 2D materials with a large inherent band-gap.

3. PHASE-TRANSITION MODULATION FOR VO2

Phase transitions are ubiquitous in nature, such as superconductive phase transitions [72], ferroelectric and antiferroelectric phase transitions [73], and insulator–metal transitions (IMTs) [7481]. The complex physical science contained in the transition process is intriguing with respect to the study of phase-transition kinetics in an attempt to well-understand the changes of energy structure, electron correlation, electron-lattice coupling, and so on [79]. VO2, an archetypal IMT, exhibits a reversible IMT at about 340 K accompanied by a transition from a low-temperature insulating phase with a monoclinic space group of P21/c to a high-temperature rutile metallic phase with a tetragonal space group of P42/mnm [75,77,79]. Therefore, it worth noting that the IMT phase transition of VO2 could change the bandgap and broaden the responding wavelength band with the increase of the temperature.

The IMT of VO2 could be considered as the Coulomb interactions between 3d electrons of V4+ ions and the degeneracy of their energy levels. For the rutile metallic phase of VO2, each V4+ ion is located at the octahedral site surrounded by six O2 ions. Based on the Goodenough’s model [82], the d levels of the V4+ ions are first split into low-energy triply degenerate t2g states and high-energy doubly degenerate egσ states. The egσ states lie 3.5 eV above the Fermi level and are empty bands [75,83]. The t2g multiplet then are separated into an a1g singlet (d state) and an egπ doublet (π state and π* state). The a1g(d) orbitals are parallel to the c-axis with σ bonding of V–V pairs, and the egπ doublet (π and π*) are strongly hybridized with the O 2p orbitals. The d orbitals and π* orbitals partly overlap and both of them partially occupy the Fermi level. The monoclinic insulating phase is identified as a Peierls–Mott insulator, and in that phase, the dimerization and tilting of the V–V pairs could lead to the d band along the c-axis is split into a lower-energy bonding combination and a higher-energy anti-bonding one (d*) [8385]. Moreover, the egπ states (π* state) are shifted to the upper position because of the increase of hybridization between the egπ doublet with O 2p orbitals. Ultimately, for the monoclinic phase, the bonding d orbitals are completely occupied, but the anti-bonding d(d*) and π* orbitals are empty due to both states being above the Fermi level [7880,8385].

IMT of VO2 could be triggered with different stimuli, such as temperature, light, electric field, mechanical strain, and magnetic field [7678,8688]. In terms of linear optical response, VO2 undergoes a large and rapid change, so it has potential use in ultrafast switching and sensing [89]. However the nonlinear optical response of VO2 with respect to pulse modulation in the IMT process, which is vital in terms of photonic and optoelectronic applications, has not reported up to now.

Using the first-principle theoretical calculations and the atomic structural parameters previously reported [58,59,80], the density of states (DOSs) for VO2 in the different states were calculated, consisting of the insulating phase, phases near the IMT point, and metallic phase [90]. As presented in Fig. 6, the room-temperature monoclinic phase behaves as an insulator with a band-gap of 0.68 eV, coinciding well with the result measured by photoemission spectroscopy [83]. In addition, accompanied by the IMT being in progress, the VO2 band-gap decreases and its phase enters a tetragonal metallic state from the monoclinic insulating state. The calculation of DOS for two phases at the IMT temperature point indicates that VO2 has an optical response to light with a wavelength shorter than 1.85 μm (0.68 eV), corresponding to the band-gap of the pure insulating phase.

 figure: Fig. 6.

Fig. 6. Theoretical DOS for VO2 in different phases; (a) DOS in the monoclinic phase at room temperature with a band-gap of 0.68 eV; (b) DOS in the monoclinic phase near the IMT point with a band-gap of 0.36 eV; (c) DOS in the tetragonal phase near the IMT point; (d) DOS in the tetragonal phase and final metallic states. Selected from Ref. [90].

Download Full Size | PDF

As displayed in Fig. 7, the linear optical response of the VO2 sample fabricated by the PLD method was investigated at 1.06 μm under different temperatures. It is seen that both the transmission and reflection decrease and the absorption coefficient increases with increasing temperature. When the temperature decreases, all the linear optical properties show hysteresis because of the hysteretic twisting of V–V pairs in VO2. The experimental results show good agreement with the previous literature [74].

 figure: Fig. 7.

Fig. 7. Linear optical response of VO2 layer; (a) reflection and transmission of VO2 at different temperatures and light wavelength of 1.06 μm measured by increasing (↑) and decreasing (↓) the temperature; (b) linear absorption coefficient of VO2 at different temperatures and light wavelength of 1.06 μm measured by increasing (↑) and decreasing (↓) the temperature. Selected from Ref. [90].

Download Full Size | PDF

Since light is capable of inducing the IMT process, investigating the nonlinear response of VO2 is difficult for the traditional Z-scan technique by using picosecond or femtosecond pulses. Therefore, the VO2 sample was employed as a saturable modulator to generate pulse lasers, and the kinetic process of the nonlinear response for VO2 during IMT was analyzed from the pulsed laser performance. The passively Q-switched laser at a wavelength of 1.06 μm was demonstrated with a Nd:GdVO4 crystal as the gain medium and the VO2 sample as the saturable absorber. On the basis of the analysis of Fig. 8(a), it is worth mentioning that the output power during an increasing and a decreasing of the pump power also shows hysteresis. The IMT of VO2 occurs in the pump range from 2.6 to 5.4 W, corresponding to the temperatures 310 and 360 K, respectively. When the pump power is below 2.6 W, VO2 is a pure insulator, and the pulse modulation is generated on account of the small band-gap (0.68 eV). When the pump power is above 5.4 W, VO2 fully becomes a metal material, and the pulse modulation is caused by the saturable properties of the surface plasmon resonance absorption [9194].

 figure: Fig. 8.

Fig. 8. Laser performance with VO2 as an optical switch; (a) average output power and central temperature of the laser beam in VO2 sample recorded during increasing (↑) and decreasing (↓) pump power. Inset: laser patterns achieved with CCD; (b) repetition rate and pulse width during increase (↑) and decrease (↓) of pump power; (c) peak power during increase (↑) and decrease (↓) of pump power; (d) modulation depth with increase (↑) and decrease (↓) of central temperature generated by the pump power. Selected from Ref. [90].

Download Full Size | PDF

As displayed in Fig. 8(b), it is surprising that a minimum pulse width of 43 ns is obtained at a pump power of 4.4 W, where a break point appeared in the repetition rate, corresponding to a central temperature of 347 K. The result is much better than those of the topological insulator (pulse width of 660 ns) and MoS2 (pulse width of 970 ns) in the same experimental setup [34,63]. As demonstrated in Fig. 8(d), the modulation depth ΔT could be theoretically calculated by [95]

ΔT=eαS(ω)leα(ω)l=3.52TRτ
where l=130nm is the thickness of VO2 thin film, TR=0.54ns is the cavity round-trip time, and τ is the pulse width. In addition, as shown in Fig. 9, the variation of saturation intensity and the third-order nonlinear absorption coefficient with changing temperature are elucidated. The largest saturation intensity occurs at 335 K with 21kW/cm2, a value much smaller than those of graphene (710kW/cm2), Bi2Se3 topological insulators (4.3GW/cm2), and MoS2 (2.45GW/cm2) [32,34,63]. The lower saturation intensity of VO2 indicates the sensitive nonlinear optical responses, and the promising application for VO2 in optoelectronic sensors.

 figure: Fig. 9.

Fig. 9. Nonlinear optical response of VO2 layer; (a) saturation intensity with increase (↑) and decrease (↓) of central temperature induced by pump power in the IMT process; (b) nonlinear absorption coefficient with increase (↑) and decrease (↓) of central temperature induced by pump power in the IMT process. Selected from Ref. [90].

Download Full Size | PDF

In recent years, VO2 has received more attention in some areas, including electronics, ultrafast techniques, energy technology, optical storage, and ionic gating [7476,8992,96]. By studying the kinetics of the third-order optical nonlinearity during the IMT of VO2, it is worth noting that VO2 is of a broad response band (below 1.85 μm) and has fantastic saturable absorption properties in the IMT process. The remarkable Q-switched laser with a minimum pulse width of 43 ns and the sensitive nonlinear optical responses all show the utility of phase-transition materials in the development of potential optoelectronic devices.

4. BROADBAND MODULATION BY INHERENT BAND-GAPS

A. Graphene

Graphene is of a very stable honeycomb hexagonal structure assembled by sp2-hybridized carbon atoms [18,21]. Notwithstanding the significant template for fabricating other allotropes of carbon, such as fullerenes, carbon nanotubes, and graphite, graphene was not discovered until 2004 [97], and ironically, the date is later than all these allotropes [22]. For the sp2 hybridization of graphene, the σ bond is formed by one s orbital and two p orbital (px and py) electrons between two carbon atoms in the planar structure. The π electron that is the electron in the unaffected p orbital (pz) is perpendicular to the 2D honeycomb plane. The inter-couplings between the electron wave functions of each two nearest π electrons can lead to the formation of a half-filled π band or a half-filled π* band, corresponding to the valence band and conduction band, respectively [15,22,23].

Zero band-gap of graphene possesses linear dispersion near the Dirac point. Furthermore, this material is insensitive to wavelength, and shows relatively large absorption of visible-to-infrared light (about 2.3% for a monolayer), which all were beneficial for the acquisition of broadband optical response [98100], As for the carrier dynamics of graphene, the intraband carrier relaxation time and the interband relaxation time are about 10–150 fs and 0.4–1.7 ps, respectively [101103]. The longer interband relaxation can act as the modulation switching by changing the intrinsic electron and hole carrier densities of graphene. In addition, graphene is also of superior thermal conductivity (5300 W/m/K for a monolayer) [104], large electron mobility of 105cm2/V/s [18,105], and remarkable elastic properties (a Young’s modulus of 1.0 TPa) [106]. As a consequence, graphene is a candidate broadband saturable absorber for the generation of mode-locking or Q-switched pulses.

In 2009, 5 years after the discovery of graphene, the saturable absorption of graphene was first investigated by Q. L. Bao et al. [32]. The saturation intensity IS ranges from 0.61 to 0.71MW/cm2 for different layers of graphene, corresponding to the saturation density ranging from 8.16×1013 to 5.84×1013 [32,107]. This research group also demonstrated the first saturable modulation of graphene in the Er-doped fiber mode-locking laser. The laser oscillation wavelength is about 1.57 μm and the minimum pulse width is about 756 fs. In 2010, the first graphene-based Q-switched laser was presented by Z. Q. Luo et al. in the Er-doped fiber with a pulse width of 3.7 μs [108]. Employing a Nd:YAG ceramic medium and a graphene modulator in 2010, W. D. Tan et al. reported the first diode-pumped solid-state bulk mode-locking laser with the obtained pulse width of 4 ps at 1.06 μm [64]. Since then, graphene has set off a new wave of research in the mode-locking and Q-switched pulse laser fields. The preparing method causes the prepared graphene saturable absorber to have a low damage threshold and consequently cannot be used in the Q-switching lasers with larger pulse energy than mode-locking lasers.

In the same year of 2010, by means of using graphene epitaxially grown on SiC with high refractive index as the saturable absorber and as the output coupler [109], we obtained the stable passively Q-switched laser for Nd:YAG at 1.06 μm, and it is the first time that the graphene saturable absorber was applied in the Q-switched solid-state pulse laser field. The pulse profiles under different pump power are demonstrated in Fig. 10 and the minimum pulse width is 161 ns. A similar experiment was implemented in the Nd:LuVO4 gain crystal at 1.06 μm, and as shown in Fig. 11, the minimum pulse width is about 56.2 ns [110].

 figure: Fig. 10.

Fig. 10. (a) Display of the cw laser recorded by a digital oscilloscope; (b) Q-switched pulse profile under the pump power of 11.2 W; (c) Q-switched pulse profile under the pump power of 12.9 W; (d) Q-switched pulse profile under the pump power of 14.6 W; (e) Q-switched pulse profile under the pump power of 16.5 W. Selected from Ref. [109].

Download Full Size | PDF

 figure: Fig. 11.

Fig. 11. Pulse profile with the width of 56.2 ns. Selected from Ref. [110].

Download Full Size | PDF

In addition, recently, graphene was used as a pulse modulator for the generation of a pulsed optical vortex at a wavelength of 1.36 μm in the gain medium of Nd:LYSO [111]. In that experiment, the thermal effect in the Nd:LYSO crystal caused by the large quantum defect of 40.5% is used to do mode-matching selection among different Laguerre–Gaussian (LGp,l) modes. As exhibited in Fig. 12, the different LGp,l with corresponding topological charge of 1 (l=0, 1, and 2) are directly obtained by controlling the pump power. Corresponding Hermite–Gaussian (HGm,n) modes are easily transformed with using two identical cylindrical lenses. Importantly, no topological charge is lost in the whole process suggesting that there is no angular momentum transfer between graphene and vortex pulses such that graphene shows the potential application of generating pulse optical vortices.

 figure: Fig. 12.

Fig. 12. (a) Continuous wave and pulsed output power of LGp,l modes versus the absorbed pump power; (b) transverse pattern of the laser beam. Top row, achieved LGp,l modes. Bottom row, the converted HGm,n modes corresponding the LGp,l modes. Selected from Ref. [111].

Download Full Size | PDF

B. Bi2Se3 Topological Insulator

In recent years, topological insulators (Bi2Se3, Bi2Te3, and Sb2Te3) has drawn great attention in the frontier of science research, due to the exotic physical properties [27,112118]. Topological insulators show a rhombohedral crystal structure and belong to the space group of D3d5(R3¯m). Each layer for this kind of materials is of a quintuple-layered sandwich structure along the direction of the 3-fold rotation axis of symmetry; for instance, the layered configuration of Bi2Se3 is Se1–Bi1–Se2–Bi1’–Se1’. The coupling between two atomic layers of one quintuple layer is stronger than that between two quintuple layers held together predominantly by the van der Waals forces [113,114].

Two energy states simultaneously exist in topological insulators, and they are the bulk state and the metallic surface state, respectively. Bi2Se3 topological insulators have the largest energy gap of the bulk state with a value of about 0.3 eV, and no impurity states exist in the gap [113,115]. The metallic surface state is topological protection with the combination of spin–orbit coupling and time-reversal symmetry against probable scattering [27,116,118]. Similar to graphene, the surface state consists of massless Dirac fermions with a linear energy–momentum relationship. Graphene has two spin-degenerate Dirac cones located at the K and K points, respectively. Nevertheless, for the Bi2Se3 class of topological insulators materials, there is only a single surface Dirac cone around the Γ point with no spin degeneracy [113115]. Some splendid physical properties were discovered with the gradually in-depth investigations of topological insulators, such as the electromagnetic effect, superconductivity, and quantum anomalous Hall effect [112115,119,120].

Along with the active research of electronic and magnetic phenomena of topological insulators, its optical characteristic also aroused research interest. Considering the small energy gap of the bulk state for topological insulators, the saturable absorption was expected to generate under strong excitation with the wavelength not longer than 4.13 μm [114,115]. In 2012, F. Bernard et al. discovered that the Bi2Te3 topological insulator has saturable absorption at a wavelength of 1.55 μm [121], and then, using the saturable absorber fabricated by Bi2Te3 topological insulator and the gain medium of Er-doped fiber, C. J. Zhao et al. first obtained the mode-locking lasers with a pulse width of 1.2 ps at 1.56 μm [33].

In 2013, our group demonstrated the stable passively Q-switched laser for Nd:GdVO4 at 1.06 μm with the Bi2Se3 topological insulator as the pulse modulator [63]. On the strength of a Z-scan technique, saturable absorption of Bi2Se3 saturable modulator was measured and was fitted with Eq. (4). The saturation intensity was determined to be 4.3GW/cm2, a value that is much larger than that of graphene (0.610.71MW/cm2) [32,63]. The output power and pulse profiles under different pump power are demonstrated in Fig. 13, and the minimum pulse width is about 666 ns. Employing the same Bi2Se3 saturable absorber, a dual-wavelength simultaneously Q-switched Nd:Lu2O3 laser was also achieved [122]. As shown in Fig. 14, the shortest pulse width is about 720 ns and the laser wavelengths are 1077 and 1081 nm, respectively.

 figure: Fig. 13.

Fig. 13. (a) Average output power and pulse energy vs. increasing incident pump power; (b) pulse width and repetition rate vs. incident pump power; (c) display recorded by digital oscilloscope for lasers. One through 6 are pulse profiles of cw Nd:GdVO4 laser, and pulsed lasers under pump power of 1.19, 1.27, 1.46, 1.67, and 1.85 W, respectively. Selected from Ref. [63].

Download Full Size | PDF

 figure: Fig. 14.

Fig. 14. (a) Single-pulse profile with duration of 720 ns. Inset, corresponding pulse train with the repetition rate of 94.7 kHz; (b) laser spectrum of the dual-wavelength laser at 1077 and 1081 nm. Selected from Ref. [122].

Download Full Size | PDF

At the beginning of pulse modulation for topological insulators (Bi2Se3, Bi2Te3, and Sb2Te3), the Q-switched or mode-locking lasers are concentrated on the wavelengths from 1 to 1.6 μm in the gain media such as Nd:GdVO4 [63], Yb-doped fibers [123], Er-doped fibers [33,124126], and Er:YAG ceramic [127]. In 2014, M. W. Jung reported the mode-locked femtosecond pulse laser at 1935 nm with Bi2Te3 topological insulator as the saturable absorber and the Tm/Ho co-doped fiber as the gain medium [128]. Soon after, Z. Q. Luo et al. demonstrated a Q-switched double-clad single-mode Tm3+-doped fiber laser at 2 μm wavelength with Bi2Se3 topological insulator as the saturable modulator [129]. These two experiments confirm the broadband saturable modulation of topological insulators with the applicable wavelength expanded to 2 μm or longer wavelengths.

5. CONCLUSIONS

In this paper, on the basis of the band-gap modulation, some latest research results about 2D broadband saturable absorbers, including MoS2 and VO2, graphene, and Bi2Se3 topological insulator were systematically elucidated and reviewed. For the few-layer MoS2, the broadband pulse modulation was achieved from 1.06 to 2.1 μm by introducing appropriate S defects to reduce the intrinsic band-gap. For the VO2 IMT, the transition process from the insulator phase to metal phase resulted in some fantastic results in terms of pulse modulation, such as small pulse widths and sensitive nonlinear optical responses. In addition, with respect to graphene and Bi2Se3 topological insulator, the small inherent bandgaps leaded to the success of broadband saturated absorption. As a consequence, this work shows the excellent performance of band-gap engineering for exploring 2D broadband saturable materials, and importantly, it opens up an available application opportunity for 2D materials in photonics and optoelectronics.

ACKNOWLEDGMENTS

This work was supported by the National Natural Science Foundation of China (Nos. 51025210 and 51422205) and the China Scholarship Council in 2014 (No. 201406220045).

REFERENCES

1. T. H. Maiman, “Stimulated optical radiation in ruby,” Nature 187, 493–494 (1960). [CrossRef]  

2. U. Keller, “Recent developments in compact ultrafast lasers,” Nature 424, 831–838 (2003). [CrossRef]  

3. H. Yu, J. Liu, H. Zhang, A. A. Kaminskii, Z. Wang, and J. Wang, “Advances in vanadate laser crystals at a lasing wavelength of 1 micrometer,” Laser Photon. Rev. 8, 847–864 (2014).

4. F. Bonaccorso, Z. Sun, T. Hasan, and A. C. Ferrari, “Graphene photonics and optoelectronics,” Nat. Photonics 4, 611–622 (2010). [CrossRef]  

5. Y. Shimony, Z. Burshtein, and Y. Kalisky, “Cr4+:YAG as passive Q-switch and Brewster plate in a pulsed Nd:YAG laser,” IEEE J. Quantum Electron. 31, 1738–1741 (1995). [CrossRef]  

6. T. Y. Tsai and M. Birnbaum, “Q-switched 2-μm lasers by use of a Cr2+:ZnSe saturable absorber,” Appl. Opt. 40, 6633–6637 (2001). [CrossRef]  

7. A. M. Malyarevich, I. A. Denisov, K. V. Yumashev, V. P. Mikhailov, R. S. Conroy, and B. D. Sinclair, “V:YAG–A new passive Q-switch for diode-pumped solid-state lasers,” Appl. Phys. B 67, 555–558 (1998). [CrossRef]  

8. U. Keller, D. A. B. Miller, G. D. Boyd, T. H. Chiu, J. F. Ferguson, and M. T. Asom, “Solid-state low-loss intracavity saturable absorber for Nd:YLF lasers: An antiresonant semiconductor Fabry–Perot saturable absorber,” Opt. Lett. 17, 505–507 (1992). [CrossRef]  

9. T. T. Basiev, S. B. Mirov, and V. V. Osiko, “Room-temperature color center lasers,” IEEE J. Quantum Electron. 24, 1052–1069 (1988). [CrossRef]  

10. U. Keller and A. C. Tropper, “Passively modelocked surface-emitting semiconductor lasers,” Phys. Rep. 429, 67–120 (2006). [CrossRef]  

11. X. P. Hu, P. Xu, and S. N. Zhu, “Engineered quasi-phase-matching for laser techniques [Invited],” Photon. Res. 1, 171–185 (2013).

12. Y. F. Chen, S. W. Tsai, and S. C. Wang, “High-power diode-pumped nonlinear mirror mode-locked Nd:YVO4 laser with periodically-poled KTP,” Appl. Phys. B 72, 395–397 (2001). [CrossRef]  

13. Y. H. Liu, Z. D. Xie, S. D. Pan, X. J. Lv, Y. Yuan, X. P. Hu, J. Lu, L. N. Zhao, C. D. Chen, G. Zhao, and S. N. Zhu, “Diode-pumped passively mode-locked Nd:YVO4 laser at 1342 nm with periodically poled LiTaO3,” Opt. Lett. 36, 698–700 (2011). [CrossRef]  

14. K. S. Novoselov, V. I. Fal, L. Colombo, P. R. Gellert, M. G. Schwab, and K. Kim, “A roadmap for graphene,” Nature 490, 192–200 (2012). [CrossRef]  

15. C. N. R. Rao, A. K. Sood, K. S. Subrahmanyam, and A. Govindaraj, “Graphene: The new two-dimensional nanomaterial,” Angew. Chem.. Int. Ed. Engl., Suppl. 48, 7752–7777 (2009).

16. S. Z. Butler, S. M. Hollen, L. Cao, Y. Cui, J. A. Gupta, H. R. Gutiérrez, T. F. Heinz, S. S. Hong, J. Huang, A. F. Ismach, E. Johnston-Halperin, M. Kuno, V. V. Plashnitsa, R. D. Robinson, R. S. Ruoff, S. Salahuddin, J. Shan, L. Shi, M. G. Spencer, M. Terrones, W. Windl, and J. E. Goldberger, “Progress, challenges, and opportunities in two-dimensional materials beyond graphene,” ACS Nano 7, 2898–2926 (2013). [CrossRef]  

17. M. Xu, T. Liang, M. Shi, and H. Chen, “Graphene-like two-dimensional materials,” Chem. Rev. 113, 3766–3798 (2013). [CrossRef]  

18. A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater. 6, 183–191 (2007). [CrossRef]  

19. K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov, and A. K. Geim, “Two-dimensional atomic crystals,” Proc. Natl. Acad. Sci. USA 102, 10451–10453 (2005).

20. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov, “Two-dimensional gas of massless Dirac fermions in graphene,” Nature 438, 197–200 (2005). [CrossRef]  

21. K. P. Loh, Q. Bao, P. K. Ang, and J. Yang, “The chemistry of graphene,” J. Mater. Chem. 20, 2277–2289 (2010). [CrossRef]  

22. A. H. C. Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, “The electronic properties of graphene,” Rev. Mod. Phys. 81, 109–162 (2009). [CrossRef]  

23. P. R. Wallace, “The band theory of graphite,” Phys. Rev. 71, 622–634 (1947). [CrossRef]  

24. A. K. Geim and I. V. Grigorieva, “Van der Waals heterostructures,” Nature 499, 419–425 (2013). [CrossRef]  

25. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7, 699–712 (2012). [CrossRef]  

26. Z. He, C. Zhong, S. Su, M. Xu, H. Wu, and Y. Cao, “Enhanced power-conversion efficiency in polymer solar cells using an inverted device structure,” Nat. Photonics 6, 591–595 (2012).

27. Y. Zhang, K. He, C. Z. Chang, C. L. Song, L. L. Wang, X. Chen, J. F. Jia, Z. Fang, X. Dai, W. Y. Shan, S. Q. Shen, Q. Niu, X. L. Qi, S. C. Zhang, X. C. Ma, and Q. K. Xue, “Crossover of the three-dimensional topological insulator Bi2Se3 to the two-dimensional limit,” Nat. Phys. 6, 584–588 (2010). [CrossRef]  

28. C. C. Liu, W. Feng, and Y. Yao, “Quantum spin Hall effect in silicene and two-dimensional germanium,” Phys. Rev. Lett. 107, 076802 (2011). [CrossRef]  

29. L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen, and Y. Zhang, “Black phosphorus field-effect transistors,” Nat. Nanotechnol. 9, 372–377 (2014). [CrossRef]  

30. J. Qiao, X. Kong, Z. X. Hu, F. Yang, and W. Ji, “High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus,” Nat. Commun. 5, 4475 (2014).

31. J. N. Coleman, M. Lotya, A. O’Neill, S. D. Bergin, P. J. King, U. Khan, K. Young, A. Gaucher, S. De, R. J. Smith, I. V. Schvets, S. K. Arora, G. Stanton, H. Kim, K. Lee, G. T. Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Wang, J. F. Donegan, J. C. Grunlan, G. Moriarty, A. Shmeliov, R. J. Nicholls, J. M. Perkins, E. M. Grieveson, K. Theuwissen, D. W. McComb, P. D. Nellist, and V. Nicolosi, “Two-dimensional nanosheets produced by liquid exfoliation of layered materials,” Science 331, 568–571 (2011). [CrossRef]  

32. Q. Bao, H. Zhang, Y. Wang, Z. Ni, Y. Yan, Z. Shen, K. Loh, and D. Tang, “Atomic-layer graphene as a saturable absorber for ultrafast pulsed lasers,” Adv. Funct. Mater. 19, 3077–3083 (2009). [CrossRef]  

33. C. Zhao, H. Zhang, X. Qi, Y. Chen, Z. Wang, S. Wen, and D. Tang, “Ultra-short pulse generation by a topological insulator based saturable absorber,” Appl. Phys. Lett. 101, 211106 (2012). [CrossRef]  

34. S. Wang, H. Yu, H. Zhang, A. Wang, M. Zhao, Y. Che, L. Mei, and J. Wang, “Broadband few-layer MoS2 saturable absorbers,” Adv. Mater. 26, 3538–3544 (2014). [CrossRef]  

35. L. Ci, L. Song, C. Jin, D. Jariwala, D. Wu, Y. Li, A. Srivastava, Z. F. Wang, K. Storr, L. Balicas, F. Liu, and P. M. Ajayan, “Atomic layers of hybridized boron nitride and graphene domains,” Nat. Mater. 9, 430–435 (2010). [CrossRef]  

36. S. Helveg, J. V. Lauritsen, E. Lægsgaard, I. Stensgaard, J. K. Nørskov, B. S. Clausen, H. Topsøe, and F. Besenbacher, “Atomic-scale structure of single-layer MoS2 nanoclusters,” Phys. Rev. Lett. 84, 951–954 (2000). [CrossRef]  

37. S. Wu, J. S. Ross, G. B. Liu, G. Aivazian, A. Jones, Z. Fei, W. Zhu, D. Xiao, W. Yao, D. Cobden, and X. Xu, “Electrical tuning of valley magnetic moment through symmetry control in bilayer MoS2,” Nat. Phys. 9, 149–153 (2013). [CrossRef]  

38. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol. 6, 147–150 (2011). [CrossRef]  

39. R. A. Gordon, D. Yang, E. D. Crozier, D. T. Jiang, and R. F. Frindt, “Structures of exfoliated single layers of WS2, MoS2, and MoSe2 in aqueous suspension,” Phys. Rev. B 65, 125407 (2002).

40. H. S. S. Ramakrishna Matte, A. Gomathi, A. K. Manna, D. J. Late, R. Datta, S. K. Pati, and C. N. R. Ra, “MoS2 and WS2 analogues of graphene,” Angew. Chem. 122, 4153–4156 (2010). [CrossRef]  

41. T. Li and G. Galli, “Electronic properties of MoS2 nanoparticles,” J. Phys. Chem. C 111, 16192–16196 (2007). [CrossRef]  

42. K. K. Kam and B. A. Parkinson, “Detailed photocurrent spectroscopy of the semiconducting group VIB transition metal dichalcogenides,” J. Phys. Chem. 86, 463–467 (1982). [CrossRef]  

43. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C. Y. Chim, G. Galli, and F. Wang, “Emerging photoluminescence in monolayer MoS2,” Nano Lett. 10, 1271–1275 (2010).

44. R. Coehoorn, C. Haas, J. Dijkstra, C. J. F. Flipse, R. A. D. Groot, and A. Wold, “Electronic structure of MoSe2, MoS2, and WSe2. I. Band-structure calculations and photoelectron spectroscopy,” Phys. Rev. B 35, 6195–6202 (1987).

45. J. S. Qi, X. Li, X. F. Qian, and J. Feng, “Bandgap engineering of rippled MoS2 monolayer under external electric field,” Appl. Phys. Lett. 102, 173112 (2013). [CrossRef]  

46. H. J. Conley, B. Wang, J. I. Ziegler, R. F. Haglund Jr., S. T. Pantelides, and K. I. Bolotin, “Bandgap engineering of strained monolayer and bilayer MoS2,” Nano Lett. 13, 3626–3630 (2013).

47. A. Castellanos-Gomez, R. Roldán, E. Cappelluti, M. Buscema, F. Guinea, H. S. J. van der Zant, and G. A. Steele, “Local strain engineering in atomically thin MoS2,” Nano Lett. 13, 5361–5366 (2013).

48. Q. H. Liu, L. Z. Y. F. Li, Z. X. Gao, Z. F. Chen, and J. Lu, “Tuning electronic structure of bilayer MoS2 by vertical electric field: A first-principles investigation,” J. Phys. Chem. C 116, 21556–21562 (2012). [CrossRef]  

49. Y. Kim, J. L. Huang, and C. M. Lieber, “Characterization of nanometer scale wear and oxidation of transition metal dichalcogenide lubricants by atomic force microscopy,” Appl. Phys. Lett. 59, 3404–3406 (1991). [CrossRef]  

50. A. Carvalho, R. M. Ribeiro, and A. H. C. Neto, “Band nesting and the optical response of two-dimensional semiconducting transition metal dichalcogenides,” Phys. Rev. B 88, 115205 (2013).

51. L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle, A. Mishchenko, Y.-J. Kim, R. V. Gorbachev, T. Georgiou, S. V. Morozov, A. N. Grigorenko, A. K. Geim, C. Casiraghi, A. H. C. Neto, and K. S. Novoselov, “Strong light-matter interactions in heterostructures of atomically thin films,” Science 340, 1311–1314 (2013). [CrossRef]  

52. R. Wang, B. A. Ruzicka, N. Kumar, M. Z. Bellus, H. Y. Chiu, and H. Zhao, “Ultrafast and spatially resolved studies of charge carriers in atomically thin molybdenum disulfide,” Phys. Rev. B 86, 045406 (2012).

53. N. Kumar, S. Najmaei, Q. Cui, F. Ceballos, P. M. Ajayan, J. Lou, and H. Zhang, “Second harmonic microscopy of monolayer MoS2,” Phys. Rev. B 87, 161403 (2013).

54. J. Strait, P. Nene, H. Wang, C. Zhang, and F. Rana, “Carrier relaxation dynamics in MoS2 measured by optical/THz pump-probe spectroscopy,” in Conference on Lasers and Electro-Optics (CLEO), Technical Digest Series (OSA, 2013), paper JTh2A.37.

55. H. Wang, C. Zhang, and F. Rana, “Ultrafast carrier dynamics in single and few layer MoS2 studied by optical pump probe technique,” in Conference on Lasers and Electro-Optics (CLEO), Technical Digest Series (OSA, 2013), paper QTu1D.2.

56. K. Wang, J. Wang, J. Fan, M. Lotya, A. O’Neill, D. Fox, Y. Feng, X. Zhang, B. Jiang, Q. Zhao, H. Zhang, J. N. Coleman, L. Zhang, and W. J. Blau, “Ultrafast saturable absorption of two-dimensional MoS2 nanosheets,” ACS Nano 7, 9260–9267 (2013). [CrossRef]  

57. O. Madelung, Introduction to Solid-State Theory, Vol. 2 (Springer, 1996), Chap. 8 and 9.

58. G. Kresse and J. Furthmuller, “Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set,” Comput. Mater. Sci. 6, 15–50 (1996). [CrossRef]  

59. G. Kresse and J. Furthmüller, “Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set,” Phys. Rev. B 54, 11169–11186 (1996).

60. V. Yu. Fominski, V. N. Nevolin, R. I. Romanov, and I. Smurov, “Ion-assisted deposition of MoSx films from laser-generated plume under pulsed electric field,” J. Appl. Phys. 89, 1449–1457 (2001). [CrossRef]  

61. E. Garmire and A. Kost, “Resonant optical nonlinearities in semiconductors,” in Nonlinear Optics in Semiconductor I, R. K. Willardson and E. R. Weber, eds. (Academic, 1999), pp. 2–50.

62. E. Garmire, “Resonant optical nonlinearities in semiconductors,” IEEE J. Sel. Top. Quantum Electron. 6, 1094–1110 (2000). [CrossRef]  

63. H. Yu, H. Zhang, Y. Wang, C. Zhao, B. Wang, S. Wen, H. Zhang, and J. Wang, “Topological insulator as an optical modulator for pulsed solid-state lasers,” Laser Photon. Rev. 7, L77–L83 (2013).

64. W. D. Tan, C. Y. Su, R. J. Knize, G. Q. Xie, L. J. Li, and D. Tang, “Mode locking of ceramic Nd:yttrium aluminum garnet with graphene as a saturable absorber,” Appl. Phys. Lett. 96, 031106 (2010). [CrossRef]  

65. J. Du, Q. Wang, G. Jiang, C. Xu, C. Zhao, Y. Xiang, Y. Chen, S. Wen, and H. Zhang, “Ytterbium-doped fiber laser passively mode locked by few-layer molybdenum disulfide (MoS2) saturable absorber functioned with evanescent field interaction,” Sci. Rep. 4, 6346 (2014). [CrossRef]  

66. H. Zhang, S. B. Lu, J. Zheng, J. Du, S. C. Wen, D. Y. Tang, and K. P. Loh, “Molybdenum disulfide (MoS2) as a broadband saturable absorber for ultra-fast photonics,” Opt. Express 22, 7249–7260 (2014). [CrossRef]  

67. H. Liu, A. Luo, F. Wang, R. Tang, M. Liu, Z. Luo, W. Xu, C. Zhao, and H. Zhang, “Femtosecond pulse erbium-doped fiber laser by a few-layer MoS2 saturable absorber,” Opt. Lett. 39, 4591–4594 (2014). [CrossRef]  

68. M. Liu, X. W. Zheng, Y. L. Qi, H. Liu, A. Luo, Z. C. Luo, W. C. Xu, C. J. Zhao, and H. Zhang, “Microfiber-based few-layer MoS2 saturable absorber for 2.5 GHz passively harmonic mode-locked fiber laser,” Opt. Express 22, 22841–22846 (2014). [CrossRef]  

69. R. Khazaeizhad, S. H. Kassani, H. Jeong, D. I. Yeom, and K. Oh, “Mode-locking of Er-doped fiber laser using a multilayer MoS2 thin film as a saturable absorber in both anomalous and normal dispersion regimes,” Opt. Express 22, 23732–23742 (2014). [CrossRef]  

70. Y. Z. Huang, Z. Q. Y. Y. Luo, Y. Y. Li, M. Zhong, B. Xu, K. J. Che, H. Y. Xu, Z. P. Cai, J. Peng, and J. Weng, “Widely-tunable, passively Q-switched erbium-doped fiber laser with few-layer MoS2 saturable absorber,” Opt. Express 22, 25258–25266 (2014). [CrossRef]  

71. Z. Luo, Y. Huang, M. Zhong, Y. Li, J. Wu, B. Xu, H. Xu, Z. Cai, J. Peng, and J. Weng, “1, 1.5 and 2 μm fiber lasers Q-switched by a broadband few-layer MoS2 saturable absorber,” J. Lightwave Technol. 32, 4077–4084 (2014).

72. J. Bardeen, L. N. Cooper, and J. R. Schrieffer, “Theory of superconductivity,” Phys. Rev. 108, 1175–1204 (1957). [CrossRef]  

73. C. Haas, “Phase transitions in ferroelectric and antiferroelectric crystals,” Phys. Rev. 140, A863–A868 (1965). [CrossRef]  

74. M. M. Qazilbash, M. Brehm, B.-G. Chae, P.-C. Ho, G. O. Andreev, B.-J. Kim, S. J. Yun, A. V. Balatsky, M. B. Maple, F. Keilmann, H.-T. Kim, and D. N. Basov, “Mott transition in VO2 revealed by infrared spectroscopy and nano-imaging,” Science 318, 1750–1753 (2007). [CrossRef]  

75. N. B. Aetukuri, A. X. Gray, M. Drouard, M. Cossale, L. Gao, A. H. Reid, R. Kukreja, H. Ohldag, C. A. Jenkins, E. Arenholz, K. P. Roche, H. A. Dürr, M. G. Samant, and S. S. P. Parkin, “Control of the metal-insulator transition in vanadium dioxide by modifying orbital occupancy,” Nat. Phys. 9, 661–666 (2013). [CrossRef]  

76. M. Nakano, K. Shibuya, D. Okuyama, T. Hatano, S. Ono, M. Kawasaki, Y. Iwasa, and Y. Tokura, “Collective bulk carrier delocalization driven by electrostatic surface charge accumulation,” Nature 487, 459–462 (2012). [CrossRef]  

77. M. Liu, H. Y. Hwang, H. Tao, A. Strikwerda, K. Fan, G. R. Keiser, A. J. Sternbach, K. G. West, S. Kittiwatanakul, J. Lu, S. A. Wolf, F. G. Omenetto, X. Zhang, K. A. Nelson, and R. D. Averitt, “Terahertz-field-induced insulator-to-metal transition in vanadium dioxide metamaterial,” Nature 487, 345–348 (2012). [CrossRef]  

78. V. Eyert, “VO2: A novel view from band theory,” Phys. Rev. Lett. 107, 016401 (2011). [CrossRef]  

79. M. Imada, A. Fujimori, and Y. Tokura, “Metal–insulator transitions,” Rev. Mod. Phys. 70, 1039–1263 (1998). [CrossRef]  

80. T. Yao, X. Zhang, Z. Sun, S. Liu, Y. Huang, Y. Xie, C. Wu, X. Yuan, W. Zhang, Z. Wu, G. Pan, F. Hu, L. Wu, Q. Liu, and S. Wei, “Understanding the nature of the kinetic process in a VO2 metal–insulator transition,” Phys. Rev. Lett. 105, 226405 (2010). [CrossRef]  

81. J. Jeong, N. Aetukuri, T. Graf, T. D. Schladt, M. G. Samant, and S. S. P. Parkin, “Suppression of metal–insulator transition in VO2 by electric field-induced oxygen vacancy formation,” Science 339, 1402–1405 (2013). [CrossRef]  

82. J. B. Goodenough, “Direct cation–cation interactions in several oxides,” Phys. Rev. 117, 1442–1451 (1960). [CrossRef]  

83. T. C. Koethe, Z. Hu, M. W. Haverkort, C. Schüßler-Langeheine, F. Venturini, N. B. Brookes, O. Tjernberg, W. Reichelt, H. H. Hsieh, H.-J. Lin, C. T. Chen, and L. H. Tjeng, “Transfer of spectral weight and symmetry across the metal-insulator transition in VO2,” Phys. Rev. Lett. 97, 116402 (2006). [CrossRef]  

84. S. Biermann, A. Poteryaev, A. I. Lichtenstein, and A. Georges, “Dynamical singlets and correlation-assisted Peierls transition in VO2,” Phys. Rev. Lett. 94, 026404 (2005). [CrossRef]  

85. A. Cavalleri, T. Dekorsy, H. H. W. Chong, J. C. Kiedder, and R. W. Schoenlein, “Evidence for a structurally-driven insulator-to-metal transition in VO2: A view from the ultrafast timescale,” Phys. Rev. B 70, 161102 (2004).

86. P. Limelette, A. Georges, D. Jérome, P. Wzietek, P. Metaclf, and J. M. Honig, “Universality and critical behavior at the Mott transition,” Science 302, 89–92 (2003). [CrossRef]  

87. A. Cavalleri, C. Tóth, C. W. Siders, J. A. Squier, and F. Ráksi, “Femtosecond structural dynamics in VO2 during an ultrafast solid-solid phase transition,” Phys. Rev. Lett. 87, 237401 (2001). [CrossRef]  

88. J. Umeda, H. Kusumoto, K. Narita, and E. Yamada, “Nuclear magnetic resonance in polycrystalline VO2,” J. Chem. Phys. 42, 1458–1459 (1965). [CrossRef]  

89. J. H. Park, J. M. Coy, T. S. Kasirga, C. Huang, Z. Fei, S. Hunter, and D. H. Gobden, “Measurement of a solid-state triple point at the metal-insulator transition in VO2,” Nature 500, 431–434 (2013). [CrossRef]  

90. H. Yu, S. Wang, A. Wang, M. Zhao, H. Zhang, Y. Chen, L. Mei, and J. Wang, “Kinetics of nonlinear optical response at insulator-metal transition in vanadium dioxide,” Adv. Opt. Mater. 3, 64–70 (2015)

91. L. Wang, E. Radue, S. Kittiwatanakul, C. Clavero, J. Lu, S. A. Wolf, I. Novikova, and R. A. Lukaszew, “Surface plasmon polaritons in VO2 thin films for tunable low-loss plasmonic applications,” Opt. Lett. 37, 4335–4337 (2012). [CrossRef]  

92. M. Rini, A. Cavalleri, R. W. Schoenlein, R. López, L. C. Feldman, R. F. Haglund, L. A. Boatner, and T. E. Haynes, “Photoinduced phase transition in VO2 nanocrystals: Ultrafast control of surface-plasmon resonance,” Opt. Lett. 30, 558–560 (2005). [CrossRef]  

93. B. A. Kruger, A. Joushaghani, and J. K. S. Poon, “Design of electrically driven hybrid vanadium dioxide (VO2) plasmonic switches,” Opt. Express 20, 23598–23609 (2012). [CrossRef]  

94. Y. Zhou, A. Huang, Y. Li, S. Ji, Y. Gao, and P. Jin, “Surface plasmon resonance induced excellent solar control for VO2@SiO2 nanorods-based thermochromic foils,” Nanoscale 5, 9208–9213 (2013). [CrossRef]  

95. X. Zhang, S. Zhao, Q. Wang, Q. Zhang, L. Sun, and S. Zhang, “Optimization of Cr4+-doped saturable-absorber Q-switched lasers,” IEEE J. Quantum Electron. 33, 2286–2294 (1997). [CrossRef]  

96. C. Kübler, H. Ehrke, R. Huber, R. Lopez, A. Halabica, R. F. Haglund Jr., and A. Leitenstorfer, “Coherent structural dynamics and electronic correlations during an ultrafast insulator-to-metal phase transition in VO2,” Phys. Rev. Lett. 99, 116401 (2007). [CrossRef]  

97. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, “Electric field effect in atomically thin carbon films,” Science 306, 666–669 (2004). [CrossRef]  

98. A. B. Kuzmenko, E. V. Heumen, F. Carbone, and D. V. D. Marel, “Universal optical conductance of graphite,” Phys. Rev. Lett. 100, 117401 (2008). [CrossRef]  

99. R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim, “Fine structure constant defines visual transparency of graphene,” Science 320, 1308 (2008). [CrossRef]  

100. Z. Q. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. L. Stormer, and D. N. Basov, “Dirac charge dynamics in graphene by infrared spectroscopy,” Nat. Phys. 4, 532–535 (2008). [CrossRef]  

101. J. M. Dawlaty, S. Shivaraman, M. Chandrashekhar, F. Rana, and M. G. Spencer, “Measurement of ultrafast carrier dynamics in epitaxial graphene,” Appl. Phys. Lett. 92, 042116 (2008). [CrossRef]  

102. P. A. George, J. Strait, J. Dawlaty, S. Shivaraman, M. Chandrashekhar, F. Rana, and M. G. Spencer, “Ultrafast optical-pump terahertz-probe spectroscopy of the carrier relaxation and recombination dynamics in epitaxial graphene,” Nano Lett. 8, 4248–4251 (2008).

103. M. Breusing, C. Ropers, and T. Elsaesser, “Ultrafast carrier dynamics in graphite,” Phys. Rev. Lett. 102, 086809 (2009). [CrossRef]  

104. A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. teweldebrhan, F. Miao, and C. N. Lau, “Superior thermal conductivity of single-layer graphene,” Nano Lett. 8, 902–907 (2008).

105. S. V. Morozov, K. S. Novoselov, M. I. Katsnelson, F. Schedin, D. C. Elias, J. A. Jaszczak, and A. K. Geim, “Giant intrinsic carrier mobilities in graphene and its bilayer,” Phys. Rev. Lett. 100, 016602 (2008). [CrossRef]  

106. C. Lee, X. Wei, J. W. Kysar, and J. Hone, “Measurement of the elastic properties and intrinsic strength of monolayer graphene,” Science 321, 385–388 (2008). [CrossRef]  

107. H. Zhang, S. Virally, Q. Bao, K. P. Loh, S. Massar, N. Godbout, and P. Kockaert, “Z-scan measurement of the nonlinear refractive index of graphene,” Opt. Lett. 37, 1856–1858 (2012). [CrossRef]  

108. Z. Luo, M. Zhou, J. Weng, G. M. Huang, H. Y. Xu, C. C. Ye, and Z. P. Cai, “Graphene-based passively Q-switched dual-wavelength erbium-doped fiber laser,” Opt. Lett. 35, 3709–3711 (2010). [CrossRef]  

109. H. Yu, X. Chen, H. Zhang, X. Xu, X. Hu, Z. Wang, J. Wang, S. Zhuang, and M. Jiang, “Large energy pulse generation modulated by graphene epitaxially grown on silicon carbide,” ACS Nano 4, 7582–7586 (2010). [CrossRef]  

110. H. Yu, X. Chen, X. Hu, S. Zhuang, Z. Wang, X. Xu, J. Wang, H. Zhang, and M. Jiang, “Graphene as a Q-switcher for neodymium-doped lutetium vanadate laser,” Appl. Phys. Express 4, 022704 (2011). [CrossRef]  

111. Y. Ding, M. Xu, Y. Zhao, H. Yu, H. Zhang, Z. Wang, and J. Wang, “Thermally driven continuous-wave and pulsed optical vortex,” Opt. Lett. 39, 2366–2369 (2014). [CrossRef]  

112. C. Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y. Ou, P. Wei, L. L. Wang, Z. Q. Ji, Y. Feng, S. Ji, X. Chen, J. Jia, X. Dai, Z. Fang, S. C. Zhang, K. He, Y. Wang, L. Lu, X. C. Ma, and Q. K. Xue, “Experimental observation of the quantum anomalous Hall effect in a magnetic topological insulator,” Science 340, 167–170 (2013). [CrossRef]  

113. Y. L. Chen, J. G. Analytis, J. H. Chu, Z. K. Liu, S. K. Mo, X. L. Qi, H. J. Zhang, D. H. Lu, X. Dai, Z. Fang, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z. X. Shen, “Experimental realization of a three-dimensional topological insulator, Bi2Te3,” Science 325, 178–181 (2009). [CrossRef]  

114. H. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S. C. Zhang, “Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface,” Nat. Phys. 5, 438–442 (2009). [CrossRef]  

115. Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, “Observation of a large-gap topological-insulator class with a single Dirac cone on the surface,” Nat. Phys. 5, 398–402 (2009). [CrossRef]  

116. J. E. Moore, “The birth of topological insulators,” Nature 464, 194–198 (2010). [CrossRef]  

117. J. G. Analytis, R. D. McDonald, S. C. Riggs, J. H. Chu, G. S. Boebinger, and L. R. Fisher, “Two-dimensional surface state in the quantum limit of a topological insulator,” Nat. Phys. 6, 960–964 (2010). [CrossRef]  

118. J. Moore, “Topological insulators: The next generation,” Nat. Phys. 5, 378–380 (2009). [CrossRef]  

119. M. Veldhorst, M. Snelder, M. Hoek, T. Gang, V. K. Guduru, X. L. Wang, U. Zeitler, W. G. van der Wiel, A. A. Golubov, H. Hilgenkamp, and A. Brinkman, “Josephson supercurrent through a topological insulator surface state,” Nat. Mater. 11, 417–421 (2012). [CrossRef]  

120. Y. L. Chen, J. H. Chu, J. G. Analytis, Z. K. Liu, K. Lgarashi, H. H. Kuo, S. K. Mo, R. G. Moore, D. H. Lu, M. Hashimoto, T. Sasagawa, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z. X. Shen, “Massive Dirac fermion on the surface of a magnetically doped topological insulator,” Science 329, 659–662 (2010). [CrossRef]  

121. F. Bernard, H. Zhang, S. P. Gorza, and P. Emplit, “Towards mode-locked fiber laser using topological insulators,” in Nonlinear Photonics, OSA Digest Series (OSA, 2012) paper NTh1A.5.

122. B. Wang, H. Yu, H. Zhang, C. Zhao, S. Wen, H. Zhang, and J. Wang, “Topological insulator simultaneously Q-switched dual-wavelength Nd:Lu2O3 laser,” IEEE J. Photon. 6, 1501007 (2014).

123. Z. Luo, Y. Huang, J. Weng, H. Cheng, Z. Lin, B. Xu, Z. Cai, and H. Xu, “1.06 μm Q-switched ytterbium-doped fiber laser using few-layer topological insulator Bi2Se3 as a saturable absorber,” Opt. Express 21, 29516–29522 (2013). [CrossRef]  

124. C. Zhao, Y. Zou, Y. Chen, Z. Wang, S. Lu, H. Zhang, S. Wen, and D. Tang, “Wavelength-tunable picosecond soliton fiber laser with topological insulator: Bi2Se3 as a mode locker,” Opt. Express 20, 27888–27895 (2012). [CrossRef]  

125. Y. Chen, C. Zhao, H. Huang, S. Chen, P. Tang, Z. Wang, S. Lu, H. Zhang, S. Wen, and D. Tang, “Self-assembled topological insulator: Bi2Se3 membrane as a passive Q-switcher in an erbium–doped fiber laser,” J. Lightwave Technol. 31, 2857–2863 (2013). [CrossRef]  

126. Z. C. Luo, M. Liu, H. Liu, X. W. Zheng, A. P. Luo, C. J. Zhao, H. Zhang, S. C. Wen, and W. C. Xu, “2 GHz passively harmonic mode-locked fiber laser by a microfiber-based topological insulator saturable absorber,” Opt. Lett. 38, 5212–5215 (2013). [CrossRef]  

127. P. Tang, X. Zhang, C. Zhao, Y. Wang, H. Zhang, D. Shen, S. Wen, D. Tang, and D. Fan, “Topological insulator: Bi2Te3 saturable absorber for the passive Q-switching operation of an in-band pumped 1645-nm Er:YAG ceramic laser,” IEEE J. Photon. 5, 1500707 (2013).

128. M. Jung, J. Lee, J. Koo, J. Park, Y. W. Song, K. Lee, S. Lee, and J. H. Lee, “A femtosecond pulse fiber laser at 1935 nm using a bulk-structured Bi2Te3 topological insulator,” Opt. Express 22, 7865–7874 (2014). [CrossRef]  

129. Z. Q. Luo, C. Liu, Y. Z. Huang, D. D. Wu, J. Y. Wu, H. Y. Xu, Z. P. Cai, Z. Q. Lin, L. P. Sun, and J. Weng, “Topological-insulator passively Q-switched double-clad fiber laser at 2 μm wavelength,” IEEE J. Sel. Top. Quantum Electron. 20, 0902708 (2014).

Cited By

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

Alert me when this article is cited.


Figures (14)

Fig. 1.
Fig. 1. Brillouin zone (left) and calculated band structure (blue lines, right) of bulk MoS 2 with stoichiometric ratio. Selected from Ref. [34].
Fig. 2.
Fig. 2. Theoretical band gap of MoS 2 samples; (a) AB stacked MoS 2 observed from the top (left) and side (right); (b) calculated band structure of bulk MoS 2 with R = 1 2.12 ; (c) calculated band structure of bulk MoS 2 with R = 1 2.09 ; (d) calculated band structure of bulk MoS 2 with R = 1 2.04 ; (e) calculated band structure of bulk MoS 2 with R = 1 1.97 ; (f) calculated band structure of bulk MoS 2 with R = 1 1.94 ; (g) calculated band structure of bulk MoS 2 with R = 1 1.89 . Selected from Ref. [34].
Fig. 3.
Fig. 3. Measured absorption spectrum of MoS 2 sample. Selected from Ref. [34].
Fig. 4.
Fig. 4. Average output power and repetition rate of passively Q -switched laser; (a) passively Q -switched Nd : GdVO 4 laser performance at 1.06 μm; (b) passively Q -switched Nd:YGG laser performance at 1.42 μm; (c) passively Q -switched Tm:Ho:YGG laser performance at 2.1 μm; (d) relation between the carrier density ( N ) and pulse repetition rate at the laser wavelength of 1.42 μm. Selected from Ref. [34].
Fig. 5.
Fig. 5. Passively Q -switched laser spectra and pulses; (a) passively Q -switched Nd : GdVO 4 , Nd:YGG, and Tm:Ho:YGG laser spectra of at center wavelengths of 1.06, 1.42, and 2.1 μm, respectively; (b) passively Q -switched Nd : GdVO 4 , Nd:YGG, and Tm:Ho:YGG laser spectra with the pulse width of 970, 729, and 410 ns, respectively. Selected from Ref. [34].
Fig. 6.
Fig. 6. Theoretical DOS for VO 2 in different phases; (a) DOS in the monoclinic phase at room temperature with a band-gap of 0.68 eV; (b) DOS in the monoclinic phase near the IMT point with a band-gap of 0.36 eV; (c) DOS in the tetragonal phase near the IMT point; (d) DOS in the tetragonal phase and final metallic states. Selected from Ref. [90].
Fig. 7.
Fig. 7. Linear optical response of VO 2 layer; (a) reflection and transmission of VO 2 at different temperatures and light wavelength of 1.06 μm measured by increasing (↑) and decreasing (↓) the temperature; (b) linear absorption coefficient of VO 2 at different temperatures and light wavelength of 1.06 μm measured by increasing (↑) and decreasing (↓) the temperature. Selected from Ref. [90].
Fig. 8.
Fig. 8. Laser performance with VO 2 as an optical switch; (a) average output power and central temperature of the laser beam in VO 2 sample recorded during increasing (↑) and decreasing (↓) pump power. Inset: laser patterns achieved with CCD; (b) repetition rate and pulse width during increase (↑) and decrease (↓) of pump power; (c) peak power during increase (↑) and decrease (↓) of pump power; (d) modulation depth with increase (↑) and decrease (↓) of central temperature generated by the pump power. Selected from Ref. [90].
Fig. 9.
Fig. 9. Nonlinear optical response of VO 2 layer; (a) saturation intensity with increase (↑) and decrease (↓) of central temperature induced by pump power in the IMT process; (b) nonlinear absorption coefficient with increase (↑) and decrease (↓) of central temperature induced by pump power in the IMT process. Selected from Ref. [90].
Fig. 10.
Fig. 10. (a) Display of the cw laser recorded by a digital oscilloscope; (b)  Q -switched pulse profile under the pump power of 11.2 W; (c)  Q -switched pulse profile under the pump power of 12.9 W; (d)  Q -switched pulse profile under the pump power of 14.6 W; (e)  Q -switched pulse profile under the pump power of 16.5 W. Selected from Ref. [109].
Fig. 11.
Fig. 11. Pulse profile with the width of 56.2 ns. Selected from Ref. [110].
Fig. 12.
Fig. 12. (a) Continuous wave and pulsed output power of LG p , l modes versus the absorbed pump power; (b) transverse pattern of the laser beam. Top row, achieved LG p , l modes. Bottom row, the converted HG m , n modes corresponding the LG p , l modes. Selected from Ref. [111].
Fig. 13.
Fig. 13. (a) Average output power and pulse energy vs. increasing incident pump power; (b) pulse width and repetition rate vs. incident pump power; (c) display recorded by digital oscilloscope for lasers. One through 6 are pulse profiles of cw Nd : GdVO 4 laser, and pulsed lasers under pump power of 1.19, 1.27, 1.46, 1.67, and 1.85 W, respectively. Selected from Ref. [63].
Fig. 14.
Fig. 14. (a) Single-pulse profile with duration of 720 ns. Inset, corresponding pulse train with the repetition rate of 94.7 kHz; (b) laser spectrum of the dual-wavelength laser at 1077 and 1081 nm. Selected from Ref. [122].

Equations (5)

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

α * = α S * 1 + N N S + α NS * ,
N = α * I τ ω ,
α * = α S * 1 + I I S + α NS * ,
T = A exp ( δ T 1 + I I S ) ,
Δ T = e α S ( ω ) l e α ( ω ) l = 3.52 T R τ
Select as filters


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