Next Article in Journal
The Chemical Deformation of a Thermally Cured Polyimide Film Surface into Neutral 1,2,4,5-Benzentetracarbonyliron and 4,4′-Oxydianiline to Remarkably Enhance the Chemical–Mechanical Planarization Polishing Rate
Previous Article in Journal
Unlocking the Potential of Na2Ti3O7-C Hollow Microspheres in Sodium-Ion Batteries via Template-Free Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quasi-1D NbTe4 for Broadband Pulse Generation from 1.0 to 3.0 μm: Bridging the Near- and Mid-Infrared

1
State Key Laboratory of Radio Frequency Heterogeneous Integration, College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
2
Key Laboratory of Optoelectronic Devices and Systems of Ministry of Education and Guangdong Province, College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
3
College of Medical Engineering and Technology, Xinjiang Medical University, Urumqi 830017, China
4
International Collaborative Laboratory of 2D Materials for Optoelectronic Science & Technology of Ministry of Education, Institute of Microscale Optoelectronics (IMO), Shenzhen University, Shenzhen 518060, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(6), 424; https://doi.org/10.3390/nano15060424
Submission received: 9 February 2025 / Revised: 6 March 2025 / Accepted: 8 March 2025 / Published: 10 March 2025
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)

Abstract

:
Quasi-one-dimensional (quasi-1D) transition metal chalcogenides (TMCs), a subclass of low-dimensional materials, have attracted significant attention due to their unique optical and electronic properties, making them promising candidates for nonlinear photonics. In this work, NbTe4, a quasi-1D transition metal tetrachalcogenide, was synthesized and employed for the first time as a broadband saturable absorber (SA) for pulsed laser applications. The nonlinear optical (NLO) properties of NbTe4 were systematically characterized at 1.0 μm, 2.0 μm, and 3.0 μm, revealing saturation intensities of 59.53 GW/cm2, 14 GW/cm2, and 6.8 MW/cm2, with corresponding modulation depths of 17.4%, 5.3%, and 21.5%. Utilizing NbTe4-SA, passively Q-switched (PQS) pulses were successfully generated in the 1.0 μm and 2.0 μm bands, achieving pulse durations of 86 ns and 2 μs, respectively. Furthermore, stable mode-locked operation was demonstrated in an Er-doped fluoride fiber laser at 3.0 μm, yielding a pulse duration of 19 ps. These results establish NbTe4 as a highly promising broadband SA material for next-generation ultrafast photonic devices and pave the way for the development of other quasi-1D materials in nonlinear optics.

1. Introduction

Low-dimensional materials have attracted considerable attention in recent academic research due to their strong confinement effects and tunable physical properties, especially for quantum dot (QD) materials, two-dimensional (2D) materials, and quasi-one-dimensional (quasi-1D) materials, which enable pronounced interactions between light and matter. Over the past few decades, extensive efforts have been devoted to exploring these low-dimensional materials and their potential applications in various fields, including optoelectronics, bioelectronics, laser technology, energy storage devices, and optical modulators. Among them, QDs exhibit strong quantum confinement effects, and their size-dependent optical properties often lead to inhomogeneous broadening, which can limit their stability in certain applications. Meanwhile, two-dimensional (2D) materials have been widely studied, encompassing graphene [1], black phosphorus [2], topology insulators (e.g., Bi2Te3) [3,4], indium sulfide (e.g., In2S3) [5], transition metal dichalcogenides (e.g., MoS2 and WS2) [6,7,8], IV-VI semiconductors (e.g., SnS2 and SnSe2) [9,10], and Mxenes [11,12,13,14]. These materials have been effectively employed as broadband saturable absorbers (SAs) in lasers, facilitating the generation of Q-switched and mode-locked pulses with remarkable nonlinear absorption properties, but their interaction with light is often constrained to the plane of the material, limiting their effective nonlinear coefficient.
In contrast to conventional 2D materials, quasi-one-dimensional (quasi-1D) van der Waals (vdW) materials exhibit strong covalent bonds along the 1D chain direction, while weak vdW interactions exist between adjacent chains, as shown in Appendix A.1. The presence of additional bonding between chains facilitates the self-organization of 1D chains into 2D sheets. Similar to other 2D structures, these sheets are stacked into three-dimensional (3D) bulk crystals via weak vdW forces. These unique structural characteristics endow quasi-1D vdW materials with a combination of 1D and 2D properties [15]. Consequently, these materials exhibit diverse and intriguing electronic characteristics, including tunable bandgaps, charge density waves (CDWs), and superconductivity (SC), as well as highly anisotropic optical, thermoelectric, and nonlinear optical properties. For instance, the unique quasi-1D structure of single-walled carbon nanotubes (SWCNTs) grants them exceptional mechanical, electrical, and thermal properties, while their nonlinear excitonic optical behavior has been extensively studied for SA applications [16]. In solar cell applications, Sb2S3 exhibits carrier mobility and conductivity along the [001] direction that is approximately two orders of magnitude higher than along the [100] and [010] orientations due to its quasi-1D crystal structure [17]. Similarly, Ca(BO2)2, which consists of well-aligned 1D anion chains of corner-connected planar sp2-hybridized BO3 groups, displays significant anisotropy in polarizability, leading to large birefringence [18]. Transition metal trichalcogenides (TMTCs), commonly referred to as MX3 compounds (M = Ti, Zr, Hf, V, Nb, and Ta; X = S, Se, and Te), are typical quasi-1D vdW materials. Among them, ZrTe3 and NbSe3 exhibit unique CDW transitions and superconductivity, making them promising candidates for radio frequency nanoelectronic devices, as well as for applications in information processing and quantum computing [14]. Furthermore, owing to their strong in-plane anisotropy, ZrS3 nanosheets enable broadband and polarized photodetection with high tunability [19].
In recent years, there has been growing interest in chalcogen-rich transition metal chalcogenides, particularly tetrachalcogenides MX4 (M = V, Nb, and Ta; X = S, Se, and Te) [20]. For example, VS4, which exhibits a unique 1D atomic-chain structure, has been identified as a promising electrode material for high-performance batteries [21,22,23]. Additionally, metallic transition metal tetrachalcogenides, such as MTe4 (M = V, Nb, and Ta), have attracted substantial attention as a platform for investigating fundamental physical properties due to their distinctive quasi-1D structures [24]. The emergence of superconductivity in quasi-1D TaTe4 under pressure, observed through high-pressure electrical transport measurements, suggests that increased charge carrier density plays a crucial role in this phenomenon [25]. Moreover, NbTe4, as a phase-change material, has demonstrated exceptional switching performance in memory cells, achieving rapid switching speeds (30 ns) and exhibiting a Set/Reset resistance window exceeding two orders of magnitude [26]. Additionally, NbTe4 undergoes successive CDW transitions and superconductivity at a temperature of 2.2 K under hydrostatic pressure [27]. These studies highlight the significant potential of MTe4 (M = V, Nb, and Ta) in various technological fields. However, their application as SAs in solid-state lasers remains largely unexplored. Investigating the nonlinear optical (NLO) properties of MTe4 and developing solid-state pulsed lasers based on these materials would be highly beneficial for advancing photonic technologies.
NbTe4, as a quasi-1D material, maintains a well-defined electronic structure along its chain-like configuration, ensuring more stable and predictable nonlinear optical responses. Moreover, NbTe4 benefits from its quasi-1D structure, which enhances light–matter interaction along its chain direction while retaining high anisotropy, making it suitable for polarization-sensitive applications and achieving broadband NLO properties through symmetry-breaking distortions. Meanwhile, NbTe4 has a combination of strong nonlinear response, high carrier mobility, and structural stability. We anticipate that NbTe4 will serve as a promising candidate for ultrafast optical modulators and nonlinear frequency converters.
In this study, we investigate the broadband nonlinear absorption behavior of NbTe4 nanosheets, synthesized using the ultrasound-assisted liquid-phase exfoliation (LPE) technique. The structural characteristics of the prepared NbTe4 nanosheets are examined using X-ray diffraction (XRD), atomic force microscopy (AFM), and high-resolution transmission electron microscopy (HRTEM). Their optical properties are analyzed through Raman spectroscopy and transmission spectra measurements. For the first time, we fabricated a novel SA based on NbTe4, demonstrating pronounced NLO properties at 1.0 μm, 2.0 μm, and 3.0 μm, with saturation intensities of 59.53 GW/cm2, 14 GW/cm2, and 6.8 MW/cm2 and modulation depths of 17.4%, 5.3%, and 21.5%, respectively. By incorporating the fabricated NbTe4-SA, we successfully demonstrate passively Q-switched (PQS) solid-state lasers at 1.0 μm and 2.0 μm, as well as a mode-locked Er-doped fluoride fiber laser at 3.0 μm, achieving pulse durations of 86 ns, 2 μs, and 19 ps, respectively. These findings present a new avenue for utilizing quasi-1D tetrachalcogenides in laser and photonics applications.

2. Materials and Methods

2.1. Synthesis of NbTe4 Powder

NbTe4 powder was synthesized using a high-temperature solid-state reaction while maintaining a stoichiometric ratio of the constituent elements. The synthesis process involved the following steps: (1) weighing and mixing Nb and Te powders in the appropriate stoichiometric ratio, followed by loading the mixture into a graphite crucible, which was then placed inside a silica tube; (2) evacuating the silica tube to a vacuum of 10−3 Pa and sealing it using a flame; (3) placing the sealed silica tube inside a muffle furnace; (4) gradually heating the tube to 800 °C at a rate of 20 °C h−1 and maintaining this temperature for 100 h, followed by slow cooling to room temperature. Following this procedure, polycrystalline NbTe4 powder was obtained.

2.2. Preparation of NbTe4 Nanosheets and NbTe4-Based Saturable Absorber

As shown in Figure 1, to obtain a large quantity of few-layer NbTe4 nanosheets for the fabrication of a NbTe4-based saturable absorber (NbTe4-SA), the LPE method was employed. First, the synthesized NbTe4 powder was mechanically milled and subsequently dispersed in isopropyl alcohol (IPA). The dispersion was then subjected to ultrasonic treatment (200 W; 8 h), with careful control of sonication duration to prevent structural degradation due to overheating. Following this, the mixture was centrifuged at 6000 rpm for 15 min, and the supernatant, containing few-layer nanosheets, was extracted. The obtained nanosheets were uniformly deposited onto a 10 × 10 × 0.2 mm3 glass substrate and a gold mirror with a diameter of 25.4 mm. Finally, the sample was dried at 60 °C for 2 h in a vacuum oven to remove residual IPA, yielding the NbTe4-SA.

2.3. Nonlinear Optical Response of NbTe4-Based Saturable Absorber

In this study, the nonlinear optical properties of NbTe4-SA were investigated using an open-aperture (OA) Z-scan measurement system and a P-scan measurement system. The nonlinear optical response of NbTe4-SA arises from its saturable absorption characteristics. For the principle of saturable absorption, please refer to Appendix A.2. The OA Z-scan system, depicted in Figure 2a, was employed for measurements at 1.0 μm and 2.0 μm. A femtosecond laser source with a central wavelength of 1030 nm was utilized, delivering pulses with a duration of 500 fs at a repetition rate of 25 kHz. For 2.0 μm, we used a femtosecond laser source with a wavelength of 2000 nm, a pulse width of 225 fs, and a repetition rate of 10 kHz. The laser output was split by a 50:50 beam splitter into two beams: the reflected beam served as the reference, while the transmitted beam was used for testing. The reference beam was further divided into two paths for waveform monitoring and power measurement. The test beam was focused onto the sample using a lens with a focal length of 150 mm. The transmitted laser pulse was subsequently measured using a power meter (D2) and a photodetector (PD2).
The NbTe4 sample was placed in a cuvette with a thickness of 1 mm. By adjusting the electric displacement platform along the z-axis, transmission pulse parameters under varying conditions were obtained, enabling the construction of the OA Z-scan curve [28]. Based on this curve, the relationship between transmittance and input laser intensity was derived. The nonlinear absorption fitting curve is described by the following equation [29]:
T I = 1 T × e x p I I S α n s / T 0
where T I is the normalized transmittance, T is the modulation depth, I is the input light intensity, I S is the saturable absorber intensity, α n s is the unsaturated absorption intensity, and T 0 is the linear transmittance. As shown in Figure 2c,d, the corresponding modulation depth of 1.0 μm and 2.0 μm is 17.4% and 5.3%, respectively. The saturation intensity is 59.53 GW/cm2 at 1.0 μm and 14 GW/cm2 at 2.0 μm. A typical saturable absorption behavior was observed, where the transmittance increased with increasing incident light intensity before reaching saturation.
Due to the limitations of experimental conditions, we constructed a P-scan experimental setup to quantify the saturable absorption of NbTe4-SA at a wavelength of 3.0 μm. As illustrated in Figure 2b, a 3.0 μm semiconductor saturable absorber mirror (SESAM)-based passively mode-locked fiber laser, with a pulse duration of 10 ps, was employed to measure the reflectivity curve of NbTe4-SA under varying incident pulse fluxes. A detailed description of the setup is provided in Appendix B.1. The reflectivity curves corresponding to different peak pulse powers are presented in Figure 2e, and the experimental data were fitted using the following equation:
R I = 1 Δ R × exp I / I s a t R n s
where R I represents the reflectivity, R is the modulation depth, I denotes the input pulse energy, I S a t is the saturation energy, and R n s corresponds to the non-saturable loss. The calculated saturable power of NbTe4-SA was determined to be 6.8 MW/cm2, with a modulation depth of 21.5%.
The results demonstrate that NbTe4 nanosheets exhibit significant nonlinear absorption across the 1.0–3.0 μm range, highlighting their potential as broadband saturable absorbers for pulsed laser applications. A higher saturation intensity corresponds to greater saturable absorption loss, while modulation depth plays a crucial role in optimizing pulse duration. In passively Q-switched lasers, higher saturation intensity and modulation depth enhance energy storage, enabling high-energy, narrow-pulse outputs. However, increased pump power is required to reach saturation, which reduces conversion efficiency. In passively mode-locked lasers, the modulation capability of the saturable absorber should align with the laser structure. A lower saturation intensity facilitates mode-locked initiation at lower pump power, while sufficient modulation depth is essential for suppressing gain competition and noise perturbations, thereby ensuring stable self-starting mode-locked operation. Experimental results confirm that the 1.0 μm and 2.0 μm bands are well-suited for Q-switched operation, whereas the 3.0 μm band is more suitable for mode-locked operation, necessitating distinct resonator designs and output parameter optimization.

2.4. Laser Application of NbTe4-SA in Broadband Absorption

To investigate the modulation performance of NbTe4-SA in a 1.0 μm laser system, a compact solid-state laser cavity was designed, achieving Q-switching operation by incorporating multiple layers of NbTe4 thin films into the cavity. The experimental setup is illustrated in Figure 3a. A Nd:YAG crystal (3 × 3 × 4 mm3; 1.2 at.%) cut along the [111] plane was used as the gain medium. A commercial 808 nm fiber-coupled diode laser (fiber core diameter: 400 μm; NA: 0.22) served as the pump source, delivering a maximum output power of 30 W. A 1:0.5 collimation focusing system was employed to precisely focus the pump beam onto the crystal with a spot diameter of 200 μm. The gain crystal, wrapped in indium foil, was mounted on a custom-designed water-cooled copper block, with the circulating water temperature maintained at 17 °C for efficient heat dissipation. The input mirror (M1) was coated with a dichroic film (HT@808 nm; HR@1064 nm; R = ∞) to optimize the pump light coupling, while the film at the opposite end (HR@808 nm; AR@1064 nm) enhanced absorption efficiency and minimized the influence of residual pump light on NbTe4-SA. The output coupler (OC) was a concave mirror with a 50 mm radius of curvature and 5% transmittance at 1064 nm. M1 and OC formed a plano-concave cavity with a length of 25 mm.
To evaluate the applicability of NbTe4-SA in 2.0 μm laser systems, a similar single-mirror flat-cavity configuration was developed, utilizing a Tm: YLF crystal (3 × 3 × 10 mm3, 3.0 at.%; A-cut) as the gain medium. The experimental setup for the 2.0 μm PQS laser is also depicted in Figure 3a. Further details regarding the experimental design are provided in Appendix B.2.
To examine the saturable absorption properties of NbTe4-SA at 3.0 μm, a linear fiber laser cavity was constructed to enable both Q-switching and mode-locking operations. The experimental configuration is shown in Figure 3b. A commercially available fiber-coupled laser diode with an emission wavelength of 976 nm was employed as the pump source, delivering a maximum output power of 60 W. An aspheric lens (L1) with an 18 mm focal length was positioned after the diode output to collimate the pump light. A dichroic mirror with 95% transmittance at 976 nm and >99.5% reflectivity at 2780 nm was placed behind L1 to separate the pump beam from the laser output. The gain medium consisted of a 4 m-long double-clad Er-doped fluoride fiber with a doping concentration of 70,000 ppm. The fiber core had a diameter of 15 μm and a numerical aperture (NA) of 0.12, while the circular inner cladding had a diameter of 260 μm. To couple the pump light into the inner cladding of the gain fiber, an anti-reflection-coated plano-convex lens (L2) with a 25 mm focal length was employed. The fiber end facing the pump beam was perpendicularly cleaved to provide a 4% Fresnel reflection, functioning as an output coupler, while the opposite end was cleaved at an 8° angle to suppress parasitic oscillations. The fiber ends were coated with a graphite film and mounted on a copper V-groove translation stage for efficient heat dissipation. To collimate and focus the output laser beam onto a gold reflector, a pair of plano-convex lenses (L3 and L4) with 25 mm focal lengths was used. The NbTe4-SA was deposited on the gold reflector to modulate the laser output. A long-pass filter was placed at the laser system output to block residual 976 nm pump light and background noise. Additional details regarding the instrumentation used in these laser experiments can be found in Appendix B.3 and Appendix B.4.

3. Results and Discussion

3.1. Characterization of NbTe4 Nanosheets

To evaluate the purity of the samples we collected, powder XRD analysis was performed using a Bruker D2 X-ray diffractometer equipped with Cu Kα radiation (λ = 1.5418 Å) at room temperature. The diffraction data were collected over a 2θ range of 10–70° with a step size of 0.02° and a fixed counting time of 1 s per step. As shown in Figure 4a, the XRD pattern of NbTe4 powder exhibits distinct diffraction peaks at (100), (002), (211), and (213) within the measured range. The experimental XRD pattern closely matches the calculated reference pattern, confirming the high purity of the synthesized NbTe4. As illustrated in Figure 4b, Raman spectroscopy shows characteristic peaks at 83 cm−1 and 129 cm−1, corresponding to Nb–C and Nb–O bonds, respectively, consistent with previous studies. Energy dispersive X-ray spectroscopy (EDS) mapping, as shown in Figure 4c, shows the characteristic peaks of Nb and Te, and their element ratio is close to 1:4, which is consistent with the stoichiometric ratio of NbTe4. This result further verifies the chemical purity and composition uniformity of the sample. The thickness of the NbTe4 nanosheets was determined using AFM, as depicted in Figure 4d,e. In agreement with previous reports, monolayer NbTe4 nanosheets exhibit an approximate thickness of 2.0 nm, which is consistent with the results of nanosheet thickness measured in Figure 4f. A Statistical analysis indicates that monolayers constitute approximately 54.9% of the total NbTe4 nanosheets. The morphology of the NbTe4 nanosheets was characterized using transmission electron microscopy (TEM), as depicted in Figure 4g,h. An HRTEM image, shown in Figure 4h, reveals the well-defined crystalline lattice structure (inset: corresponding diffraction image). UV-Vis-NIR spectroscopy was conducted on the NbTe4 dispersion. Figure 4i reveals a gradual decrease in absorption from 240 to 1200 nm. The absence of a distinct absorption peak suggests that NbTe4 exhibits broadband optical properties, similar to graphene.

3.2. Performance Characterization of Passively Q-Switched Lasers at 1.0 μm and 2.0 μm

For 1.0 μm laser operation, experimental results are shown in Figure 5. As seen in Figure 5a, the measured average output power exhibits a linear increase with pump power. The laser pump threshold is 3 W, and the maximum average output power reaches 44 mW at a pump power of 5.42 W. The inset depicts the stability of the PQS laser over a 2 h measurement period at a pump power of 5.42 W, with an average output power of 43.09 mW and a root-mean-square deviation (RMSD) of 0.439. The variation in pulse duration and repetition frequency with increasing pump power is illustrated in Figure 5b. As the pump power increases, the pulse duration decreases from 329 ns to 88 ns, while the repetition frequency rises from 263.9 kHz to 420.6 kHz. The shortest pulse width obtained in the experiment is 86 ns, with the corresponding pulse train shown in Figure 5d. Based on the above experimental data, the single pulse energy and peak power were calculated, as presented in Figure 5c. Both parameters exhibit a rising trend with increasing pump power: the peak power varies from 0.07 W to 1.15 W, while the single pulse energy ranges from 0.02 μJ to 0.1 μJ. The emission spectrum of the PQS laser at maximum output power is shown in Figure 5e, revealing a central wavelength (λc) of 1065.7 nm and a full width at half maximum (FWHM) of 1.3 nm. Continuous pumping conditions can lead to an unstable PQS laser pulse output. To achieve a more stable laser pulse output and generate a pulse train at a defined frequency, pulse pumping was employed to optimize the output characteristics of the PQS laser. Figure 5f presents the radio frequency (RF) spectrum corresponding to the shortest pulse width of the PQS laser under pulse-pumping conditions. At a pulse-pumping frequency of 1 kHz and a pump power of 7.14 W, the output laser repetition frequency is 1 kHz, with an output power of 0.34 mW, a minimum pulse width of 60 ns, and a peak power of 5.66 W. The corresponding signal-to-noise ratio (SNR) is approximately 30 dBm. Notably, the pulse width obtained under pulse-pumping conditions is shorter than that observed under continuous pumping.
To further verify the modulation performance of NbTe4 at 2.0 μm, the fabricated sample was incorporated into a Tm:YLF laser to achieve Q-switched operation. As illustrated in Figure 6a, the maximum average output power of the PQS laser is 13.87 mW, with a slope efficiency of 0.67%. The inset depicts the stability measurement of the maximum average output power of NbTe4 over two hours (range: 1.4 mW; RMSD: 0.303), demonstrating the excellent stability of NbTe4-SA in 2.0 μm PQS lasers. The relationship between pulse width, pulse repetition frequency, and pump power is shown in Figure 6b. As the pump power increases from 3.7 W to 4.5 W, the pulse width decreases from 6.3 μs to 2.45 μs, while the pulse repetition frequency increases from 13.8 kHz to 25.7 kHz. The shortest pulse width obtained in the experiment is 2 μs, with the corresponding pulse sequence illustrated in Figure 6d. Additionally, Figure 6c shows the trends of single pulse energy and peak power as a function of pump power. At a pump power of 4.5 W, the highest recorded pulse energy is 0.388 μJ, with a corresponding peak power of 0.16 W. Figure 6e shows the output spectrum of the PQS laser, indicating a central wavelength of 1910.8 nm and an FWHM of 0.8 nm. To achieve a more stable laser pulse output and generate a pulse train at a specific frequency, pulse pumping was used to optimize the output of the PQS laser. Figure 6f displays the minimum pulse width, the corresponding pulse train, the RF spectrum, and the associated SNR of the PQS laser under pulse-pumping conditions. At a pump power of 8.38 W and a pump pulse repetition frequency of 1 kHz, the output laser repetition frequency is 1 kHz, with an output power of 2.4 mW, a minimum pulse width of 1.58 μs, a peak power of 1.52 W, and an SNR of approximately 35 dBm.

3.3. Performance Characterization of Mode-Locked Lasers at 3.0 μm

The broadband saturable absorption properties of NbTe4-SA were investigated in the 3.0 μm spectral region. In an Er-doped fluoride fiber laser cavity, NbTe4-SA was deposited onto a gold mirror to enable mode-locked pulse generation. At a pump power of 0.3 W, the laser operated in the continuous-wave (CW) regime. When the incident pump power exceeded 1.2 W, the laser transitioned to Q-switched and subsequently to Q-switched mode-locked (QML) operation. Upon further increasing the pump power to 1.8 W, stable continuous-wave mode-locking (CWML) was achieved, yielding an average output power of 125 mW, as illustrated in Figure 7a. For pump powers ranging from 1.8 W to 3.5 W, the maximum output power reached 257 mW. The stable mode-locked operation demonstrated the high damage threshold of NbTe4-SA. A further increase in pump power led to instability in the mode-locked pulses. Figure 7b shows a typical stable mode-locked pulse train observed on both nanosecond and microsecond time scales. The pulses exhibit a well-defined and stable profile, indicating the effective modulation capability of NbTe4-SA. As illustrated in Figure 7c, the emission spectrum exhibits an FWHM of 3.5 nm, centered at 2794.8 nm. At the maximum average output power, an autocorrelator was used to measure the pulse profile. Under the assumption of a Gaussian pulse shape, the measured pulse duration was 19 ps, as shown in Figure 7d. The RF spectrum of the laser at the maximum output power is presented in Figure 7e. The observed repetition rate of 23.62 MHz corresponds to an effective cavity length of 4.8 m. The RF spectrum exhibits an SNR of approximately 60 dB, confirming stable mode-locked operation. Owing to the limited bandwidth of the detector (250 MHz), the amplitudes of higher-order harmonics attenuate rapidly, as shown in the inset of Figure 7e. The long-term stability of the maximum average output power was assessed over two hours, as depicted in Figure 7f. The recorded data indicate an average output power of 256.8 mW with a fluctuation of approximately 0.85%, demonstrating stable laser operation over two hours. Furthermore, no degradation or damage to the NbTe4-SA was observed throughout the experiment.
The experimental results show that NbTe4 exhibits different saturable absorption properties across specific wavelength bands (e.g., near-infrared and mid-infrared). These differences primarily arise from three factors: the material’s energy band structure, its nonlinear optical response, and the wavelength dependence of the absorption cross-section. When the incident photon energy (E = hν) is close to the material’s band gap (Eg) or corresponds to a specific energy level transition, the light–electron interaction is significantly enhanced. Moreover, saturable absorption is a nonlinear optical phenomenon governed by third-order nonlinear susceptibility (χ(3)). The intensity of this effect depends on the material’s nonlinear response and the spatial distribution of light intensity. This allows different spectral regions to induce distinct nonlinear interactions, such as two-photon absorption and free-carrier absorption. At the same time, the absorption cross-section will vary with wavelength, influencing the overall saturable absorption behavior. The interaction of various factors led to the different modulation capacities of NbTe4 from 1.0 μm to 3.0 μm.
Table 1 compares different PQS lasers utilizing broadband absorption SAs at 1.0 μm and 2.0 μm, as well as the PQS and mode-locked pulse parameters of these SAs at 3.0 μm. As shown in Table 1, the pulse width of the NbTe4-SA-based laser is 86 ns at 1.0 μm, 2 μs at 2.0 μm, and 19 ps at 3.0 μm. Notably, the pulse width of the 3.0 μm laser is less than half that of BP and graphene, while its peak power is comparable to BP and is higher than graphene. Furthermore, NbTe4 outperforms most conventional 2D materials at 1.0 μm and 3.0 μm. These findings demonstrate that NbTe4-SA-based lasers achieve narrower pulse widths, highlighting the exceptional pulse modulation capability of the material, particularly its ability to generate mode-locked pulses at 3.0 μm. Additionally, Q-switched and mode-locked operation remains achievable even after exposing NbTe4-SA to room temperature conditions for one month. In summary, NbTe4 is a promising SA device for lasers, exhibiting excellent broadband modulation capability and long-term stability.

4. Conclusions

In this study, NbTe4 powder was synthesized using a high-temperature solid-state method. Subsequently, high-performance NbTe4-SA was fabricated using the LPE method, and its optical parameters were characterized. The nonlinear optical characteristics of the synthesized SA, including broadband saturable absorption, were analyzed using a combination of Z-scan and P-scan techniques. The measured saturation intensities were 59.53 GW/cm2 at 1.0 μm with a modulation depth of 17.4%. It is 14 GW/cm2 and 5.3% at 2.0 μm and 6.8 MW/cm2 and 21.5% at 3.0 μm. To the best of our knowledge, this study represents the first demonstration of NbTe4-SA-based lasers operating in the 1.0, 2.0, and 3.0 μm bands. PQS was successfully achieved at 1.0 μm and 2.0 μm, yielding pulse durations of 86 ns and 2 μs, respectively. A mode-locked pulse with a duration of 19 ps was generated at 3.0 μm, confirming picosecond operation. These findings highlight the potential of NbTe4 for nonlinear optical modulation and pave the way for the development of other quasi-1D broadband saturable absorber materials.

Author Contributions

Conceptualization, Z.C. and W.Z.; methodology, Z.C.; validation, Z.C., Q.K., H.H. and X.X.; formal analysis, Z.C. and W.Z.; investigation, Q.K.; resources, W.Z. and H.H.; writing—original draft preparation, Z.C.; writing—review and editing, Z.C., W.Z. and Q.W.; supervision, Q.W.; project administration, Q.W.; funding acquisition, S.L. and Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported by the National Natural Science Foundation of China (NSFC; 6217030813); Guangdong Basic and Applied Basic Research Foundation (2021A1515010964); the Shenzhen Science and Technology Program (SGDX20190919094803949, JCYJ20200109105810074, and JCYJ20220531103016036); the State Key Laboratory of Precision Measurement Technology and Instruments (023PMTI02); Medical–Engineering Interdisciplinary Research Foundation of ShenZhen University (2023YG033); and the Shenzhen Key Laboratory of 2D Meta-materials for Information Technology (ZDSYS201707271014468).

Data Availability Statement

The results presented in this paper are not publicly available but can be obtained from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
quasi-1DQuasi one dimensional
2DTwo dimensional
TMCsTransition metal chalcogenides
SASaturable absorber
PQSPassively Q-switched
NLONonlinear optical
vdWvan der Waals
CDWsCharge density waves
SCSuperconductivity
SWCNTsSingle-walled carbon nanotubes
TMTCsTransition metal trichalcogenides
LPELiquid-phase exfoliation
XRDX-ray diffraction
EDSEnergy dispersive X-ray spectroscopy
AFMAtomic force microscopy
TEMTransmission electron microscopy
HRTEMHigh-resolution transmission electron microscopy
IPAIsopropyl alcohol
OAOpen aperture
OCOutput coupler
NANumerical aperture
RMSDRoot-mean-square deviation
FWHMFull width at half maximum
RFRadio frequency
SNRSignal-to-noise ratio
CWContinuous wave
QMLQ-switched mode-locked
CWMLContinuous-wave mode-locking

Appendix A

Appendix A.1

The van der Waals interactions between adjacent chains of a 2D material (e.g., NbTe2) compared to a quasi-1D material (e.g., NbTe4) (Figure A1).
Figure A1. Crystal structures of NbTe2 and NbTe4. (a) NbTe₆ coordination in NbTe2; (b) two-dimensional layered structure of NbTe2; (c) multilayer stacking in NbTe2 with an interlayer spacing of 2.83 Å; (d) NbTe₈ coordination in NbTe4; (e) one-dimensional chain structure of NbTe4; (f) multichain stacking in NbTe4 with an interchain spacing of 1.53 Å.
Figure A1. Crystal structures of NbTe2 and NbTe4. (a) NbTe₆ coordination in NbTe2; (b) two-dimensional layered structure of NbTe2; (c) multilayer stacking in NbTe2 with an interlayer spacing of 2.83 Å; (d) NbTe₈ coordination in NbTe4; (e) one-dimensional chain structure of NbTe4; (f) multichain stacking in NbTe4 with an interchain spacing of 1.53 Å.
Nanomaterials 15 00424 g0a1

Appendix A.2

Due to variations in interlayer coupling strength and the effects of quantum confinement, exfoliated NbTe4 nanosheets typically exhibit semiconducting characteristics. The saturable absorption mechanism of NbTe4 can be understood based on the Pauli blocking principle, as illustrated in Figure A2. When the incident light intensity exceeds the bandgap energy (ΔE), it can excite electrons from the valence band (pink) to the conduction band (blue). Within a few hundred femtoseconds, these excited carriers undergo rapid thermalization, forming a hot Fermi–Dirac distribution. As a result, the generation of electron–hole pairs suppresses further interband photon transitions at energy level E. Subsequently, due to intraband phonon scattering, the carriers undergo cooling, establishing a new equilibrium between electron–hole pairs and enabling linear optical transitions at low light intensities. Under high-intensity illumination, the carriers excited by photons quickly occupy energy states near the conduction and valence band edges. This leads to additional absorption saturation, a phenomenon governed by the Pauli blocking principle.
Figure A2. Schematic illustration of the saturable absorption mechanism in NbTe4.
Figure A2. Schematic illustration of the saturable absorption mechanism in NbTe4.
Nanomaterials 15 00424 g0a2

Appendix B

Appendix B.1. P-Scan Experimental Setup

The P-scan experiment was conducted using a homemade 2790 nm SESAM-based passively mode-locked fiber laser, which has a pulse duration of 10 ps and a repetition frequency of 23.62 MHz. A 50:50 beam splitter, positioned at a 45° angle, was used to divide the laser beam into two paths. The reflected beam was directed to a power meter (Dynamometer 1) to serve as a reference. Meanwhile, the transmitted beam was focused onto the surface of the NbTe4-coated gold mirror, which was slightly tilted to reflect the beam toward a second power meter (Dynamometer 2). By varying the laser output power, a series of reflectance measurements were obtained.

Appendix B.2. 2.0 μm Laser Experimental Setup

For the 2.0 μm laser experiment, a laser system incorporating a Tm:YLF crystal (3 × 3 × 10 mm3; 3.0 at.%; A-cut) as the gain medium was designed. The experimental setup is illustrated in Figure 3a. A commercial 793 nm fiber-coupled diode laser (fiber core diameter: 200 μm; NA: 0.22) served as the pump source, providing a maximum output power of 20 W. A 1:0.8 collimation focusing system was employed to precisely focus the pump beam onto the crystal with a spot diameter of approximately 160 μm. The crystal was coated at both ends with different dielectric coatings: the input facet (M1) was coated for high transmittance at 793 nm (HT@793 nm) and high reflectance at 1908 nm (HR@1908 nm), while the output facet had an anti-reflective coating at 1908 nm (AR@1908 nm). The output coupler was a concave mirror with a radius of curvature of 50 mm, featuring 10% transmittance in the 1900–2100 nm range. The total resonator length was 25 mm.

Appendix B.3. Measurement Instruments for 1.0 μm and 2.0 μm PQS Laser Characterization

The experimental setups for PQS lasers based on NbTe4-SA at 1.0 μm and 2.0 μm are depicted in Figure 3a. The 1.0 μm laser output was measured using a power meter (30A-P-17, Ophir Optronics Solutions Ltd., Jerusalem, Israel), while the 2.0 μm laser output was recorded using a different power meter (LPE-1601, Physcience Opto-Electronics Ltd., Beijing, China). Pulse train signals for the 1.0 μm and 2.0 μm PQS lasers were detected using InAsSb and InGaAs photodetectors (DET10A/M and PDA10D2, Thorlabs, Inc., Newton, NJ, USA), respectively. A high-speed oscilloscope (DPO4104B, Tektronix, Inc., Beaverton, OR, USA) with a 1 GHz bandwidth and a 5 GHz sampling rate was employed to capture the pulse train signals. The spectral characteristics of the lasers were measured using two spectrometers: an Ocean Optics USB4000-VIS-NIR for the 1.0 μm wavelength range and a Wavescan MIR (Angewandte Physik & Elektronik GmbH, Berlin, Germany) for the 2.0 μm range.

Appendix B.4. Measurement Setup for 3.0 μm Mode-Locked Fiber Laser Characterization

For the characterization of the 3.0 μm fiber laser, the output pulse was measured through a germanium broadband filter (WG91050-C9, Thorlabs), which exhibited a measured transmittance of approximately 89% at 2780 nm. The output power was recorded using a power meter (3A-P, Ophir Optronics Solutions Ltd., Jerusalem, Israel). The mode-locked pulse train was detected using a HgCdTe photoconductive detector (PCI-9, VIGO System, Warsaw, Poland) with a 250 MHz bandwidth and was recorded using an oscilloscope (MDO3104, Tektronix Inc., Beaverton, OR, USA) with a 1 GHz bandwidth. The corresponding mode-locked spectrum was obtained using an optical spectrum analyzer (Ocean Optics SIR 5000) with a resolution of 2 nm.

References

  1. Zhu, G.; Zhu, X.; Wang, F.; Xu, S.; Li, Y.; Guo, X.; Balakrishnan, K.; Norwood, R.; Peyghambarian, N. Graphene Mode-Locked Fiber Laser at 2.8 μm. IEEE Photonics Technol. Lett. 2016, 28, 7–10. [Google Scholar] [CrossRef]
  2. Zhang, R.; Zhang, Y.; Yu, H.; Zhang, H.; Yang, R.; Yang, B.; Liu, Z.; Wang, J. Broadband Black Phosphorus Optical Modulator in the Spectral Range from Visible to Mid-Infrared. Adv. Opt. Mater. 2015, 3, 1787–1792. [Google Scholar] [CrossRef]
  3. Pan, Y.; Jia, K.; Xiong, Y.; Pan, H.; Cheng, S.; Gao, S. 1.45 Ps All-Solid-State Q-Switched Mode-Locked Laser Based on New Material Bi2Te3/Sb2Te3. Opt. Laser Technol. 2025, 182, 112228. [Google Scholar] [CrossRef]
  4. Liu, D.; He, C.; Chen, L.; Li, W.; Zu, Y. The Nonlinear Absorption Effects and Optical Limiting Properties of Bi2Te3/rGO Thin Films. Opt. Mater. 2021, 111, 110634. [Google Scholar] [CrossRef]
  5. Ma, X.; Chen, W.; Tong, L.; Liu, S.; Dai, W.; Ye, S.; Zheng, Z.; Wang, Y.; Zhou, Y.; Zhang, W.; et al. In2S3-Based Saturable Absorber for Passively Harmonic Mode-Locking in 2 μm Region. Opt. Laser Technol. 2022, 145, 107476. [Google Scholar] [CrossRef]
  6. Chen, B.; Zhang, X.; Wu, K.; Wang, H.; Wang, J.; Chen, J. Q-Switched Fiber Laser Based on Transition Metal Dichalcogenides MoS2, MoSe2, WS2, and WSe2. Opt. Express 2015, 23, 26723–26737. [Google Scholar] [CrossRef] [PubMed]
  7. Mohanraj, J.; Velmurugan, V.; Sivabalan, S. Transition Metal Dichalcogenides Based Saturable Absorbers for Pulsed Laser Technology. Opt. Mater. 2016, 60, 601–617. [Google Scholar] [CrossRef]
  8. Ma, P.; Lin, W.; Zhang, H.; Xu, S.; Yang, Z. High-Power Large-Energy Raman Soliton Generations Within a Mode-Locked Yb-Doped Fiber Laser Based on High-Damage-Threshold CVD-MoS2 as Modulator. Nanomaterials 2019, 9, 1305. [Google Scholar] [CrossRef]
  9. Dong, X.; Li, Z.; Pang, C.; Dong, N.; Jiang, H.; Wang, J.; Chen, F. Q-Switched Mode-Locked Nd:GGG Waveguide Laser with Tin Disulfide as Saturable Absorber. Opt. Mater. 2020, 100, 109702. [Google Scholar] [CrossRef]
  10. Liu, J.; Li, X.; Guo, Y.; Qyyum, A.; Shi, Z.; Feng, T.; Zhang, Y.; Jiang, C.; Liu, X. SnSe2 Nanosheets for Subpicosecond Harmonic Mode-Locked Pulse Generation. Small 2019, 15, 1902811. [Google Scholar] [CrossRef]
  11. Jiang, X.; Liu, S.; Liang, W.; Luo, S.; He, Z.; Ge, Y.; Wang, H.; Cao, R.; Zhang, F.; Wen, Q.; et al. Broadband Nonlinear Photonics in Few-Layer MXene Ti3C2Tx (T = F, O, or OH). Laser Photonics Rev. 2018, 12, 1700229. [Google Scholar] [CrossRef]
  12. Jhon, Y.; Koo, J.; Anasori, B.; Seo, M.; Lee, J.; Gogotsi, Y.; Jhon, Y. Metallic MXene Saturable Absorber for Femtosecond Mode-Locked Lasers. Adv. Mater. 2017, 29, 1702496. [Google Scholar] [CrossRef]
  13. Xia, X.; Ma, C.; Chen, H.; Khan, K.; Tateen, A.; Xiao, Q. Nonlinear Optical Properties and Ultrafast Photonics of 2D BP/Ti3C2 Heterostructures. Opt. Mater. 2021, 112, 110809. [Google Scholar] [CrossRef]
  14. Su, C.; Liu, Y.; Feng, T.; Qiao, W.; Zhao, Y.; Liang, Y.; Li, T. Optical Modulation of the MXene Ti3C2Tx Saturable Absorber for Er: Lu2O3 Laser. Opt. Mater. 2021, 115, 110949. [Google Scholar] [CrossRef]
  15. Chen, M.; Li, L.; Xu, M.; Li, W.; Zheng, L.; Wang, X. Quasi-One-Dimensional van Der Waals Transition Metal Trichalcogenides. Research 2023, 6, 0066. [Google Scholar] [CrossRef] [PubMed]
  16. Ortega-Mendoza, J.; Kuzin, E.; Zaca-Morán, P.; Chávez, F.; Gómez-Pavón, L.; Padilla-Vivanco, A.; Toxqui-Quitl, C.; Luis-Ramos, A. Photodeposition of SWCNTs onto the Optical Fiber End to Assemble a Q-Switched Er3+-Doped Fiber Laser. Opt. Laser Technol. 2017, 91, 32–35. [Google Scholar] [CrossRef]
  17. Shah, U.; Chen, S.; Khalaf, G.; Jin, Z.; Song, H. Wide Bandgap Sb2S3 Solar Cells. Adv. Funct. Mater. 2021, 31, 2100265. [Google Scholar] [CrossRef]
  18. Chen, X.; Zhang, B.; Zhang, F.; Wang, Y.; Zhang, M.; Yang, Z.; Poeppelmeier, K.; Pan, S. Designing an Excellent Deep-Ultraviolet Birefringent Material for Light Polarization. J. Am. Chem. Soc. 2018, 140, 16311–16319. [Google Scholar] [CrossRef]
  19. Chen, F.; Liu, G.; Xiao, Z.; Zhou, H.; Fei, L.; Wan, S.; Liao, X.; Yuan, J.; Zhou, Y. Quasi-One-Dimensional ZrS3 Nanoflakes for Broadband and Polarized Photodetection with High Tuning Flexibility. ACS Appl. Mater. Interfaces 2023, 15, 16999–17008. [Google Scholar] [CrossRef]
  20. Sokolov, M.; Fedin, V. Chalcogenide Clusters of Vanadium, Niobium and Tantalum. Coord. Chem. Rev. 2004, 248, 925–944. [Google Scholar] [CrossRef]
  21. Wang, Y.; Liu, Z.; Wang, C.; Yi, X.; Chen, R.; Ma, L.; Hu, Y.; Zhu, G.; Chen, T.; Tie, Z.; et al. Highly Branched VS4 Nanodendrites with 1D Atomic-Chain Structure as a Promising Cathode Material for Long-Cycling Magnesium Batteries. Adv. Mater. 2018, 30, 1802563. [Google Scholar] [CrossRef]
  22. Luo, L.; Li, J.; Asl, H.; Manthiram, A. In-Situ Assembled VS4 as a Polysulfide Mediator for High-Loading Lithium-Sulfur Batteries. ACS Energy Lett. 2020, 5, 1177–1185. [Google Scholar] [CrossRef]
  23. Xing, L.; Owusu, K.; Liu, X.; Meng, J.; Wang, K.; An, Q.; Mai, L. Insights into the Storage Mechanism of VS4 Nanowire Clusters in Aluminum-Ion Battery. Nano Energy 2021, 79, 105384. [Google Scholar] [CrossRef]
  24. Zwick, F.; Berger, H.; Grioni, M.; Margaritondo, G.; Forró, L.; LaVeigne, J.; Tanner, D.; Onellion, M. Coexisting One-Dimensional and Three-Dimensional Spectral Signatures in TaTe4. Phys. Rev. B 1999, 59, 7762–7766. [Google Scholar] [CrossRef]
  25. Yuan, Y.; Wang, W.; Zhou, Y.; Chen, X.; Cu, C.; An, C.; Zhou, Y.; Zhang, B.; Chen, C.; Zhang, R.; et al. Pressure-Induced Superconductivity in Topological Semimetal Candidate TaTe4. Adv. Electron. Mater. 2020, 6, 1901260. [Google Scholar] [CrossRef]
  26. Shuang, Y.; Chen, Q.; Kim, M.; Wang, Y.; Saito, Y.; Hatayama, S.; Fons, P.; Ando, D.; Kubo, M.; Sutou, Y. NbTe4 Phase-Change Material: Breaking the Phase-Change Temperature Balance in 2D Van Der Waals Transition-Metal Binary Chalcogenide. Adv. Mater. 2023, 35, 2303646. [Google Scholar] [CrossRef] [PubMed]
  27. de Lima, B.; Chaia, N.; Grant, T.; de Faria, L.; Canova, J.; de Oliveira, F.; Abud, F.; Machado, A. Large Enhancement of the CDW Resistivity Anomaly and Traces of Superconductivity in Imperfect Samples of NbTe4. Mater. Chem. Phys. 2019, 226, 95–99. [Google Scholar] [CrossRef]
  28. Sheik-Bahae, M.; Said, A.A.; Wei, T.H.; Hagan, D.J.; Van Stryland, E.W. Van Stryland Sensitive Measurement of Optical Nonlinearities Using a Single Beam. IEEE J. Quantum Electron. 1990, 26, 760–769. [Google Scholar] [CrossRef]
  29. Zhu, C.; Wang, F.; Meng, Y.; Yuan, X.; Xiu, F.; Luo, H.; Wang, Y.; Li, J.; Lv, X.; He, L.; et al. A Robust and Tuneable Mid-Infrared Optical Switch Enabled by Bulk Dirac Fermions. Nat. Commun. 2017, 8, 14111. [Google Scholar] [CrossRef]
  30. Chu, Z.; Liu, J.; Guo, Z.; Zhang, H. 2 μm Passively Q-Switched Laser Based on Black Phosphorus. Opt. Mater. Express 2016, 6, 2374–2379. [Google Scholar] [CrossRef]
  31. Qin, Z.; Xie, G.; Zhao, C.; Wen, S.; Yuan, P.; Qian, L. Mid-Infrared Mode-Locked Pulse Generation with Multilayer Black Phosphorus as Saturable Absorber. Opt. Lett. 2016, 41, 56–59. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, S.; Yu, H.; Zhang, H.; Wang, A.; Zhao, M.; Chen, Y.; Mei, L.; Wang, J. Broadband Few-Layer MoS2 Saturable Absorbers. Adv. Mater. 2014, 26, 3538–3544. [Google Scholar] [CrossRef]
  33. Ge, P.; Liu, J.; Jiang, S.; Xu, Y.; Man, B. Compact Q-Switched 2 μm Tm:GdVO4 Laser with MoS2 Absorber. Photonics Res. 2015, 3, 256–259. [Google Scholar] [CrossRef]
  34. Wang, S.; Tang, Y.; Yang, J.; Zhong, H.; Fan, D. MoS2 Q-Switched 2.8 μm Er:ZBLAN Fiber Laser. Laser Phys. 2019, 29, 025101. [Google Scholar] [CrossRef]
  35. Zhang, W.; Liang, Y.; Gan, Y.; Huang, H.; Liang, G.; Kang, Q.; Leng, X.; Jing, Q.; Wen, Q. VTe2: Broadband Saturable Absorber for Passively Q-Switched Lasers in the Near- and Mid-Infrared Regions. ACS Appl. Mater. Interfaces 2023, 15, 57475–57485. [Google Scholar] [CrossRef] [PubMed]
  36. Lin, H.; Liu, X.; Huang, X.; Xu, Y.; Meng, X.; Xiong, F. Monolayer Graphene-Based Passively Q-Switched Nd:YAG Laser. Optik 2016, 127, 243–245. [Google Scholar] [CrossRef]
  37. Wang, Q.; Teng, H.; Zou, Y.; Zhang, Z.; Li, D.; Wang, R.; Gao, C.; Lin, J.; Guo, L.; Wei, Z. Graphene on SiC as a Q-Switcher for a 2 μm Laser. Opt. Lett. 2012, 37, 395–397. [Google Scholar] [CrossRef]
  38. Li, P.; Zhang, G.; Zhang, H.; Zhao, C.; Chi, J.; Zhao, Z.; Yang, C.; Hu, H.; Yao, Y. Q-Switched Mode-Locked Nd:YVO4 Laser by Topological Insulator Bi2Te3 Saturable Absorber. IEEE Photonics Technol. Lett. 2014, 26, 1912–1915. [Google Scholar] [CrossRef]
  39. Gao, P.; Huang, H.; Wang, X.; Liu, H.; Huang, J.; Weng, W.; Dai, S.; Li, J.; Lin, W. Passively Q-Switched Solid-State Tm:YAG Laser Using Topological Insulator Bi2Te3 as a Saturable Absorber. Appl. Opt. 2018, 57, 2020–2024. [Google Scholar] [CrossRef]
  40. Yin, K.; Jiang, T.; Zheng, X.; Yu, H.; Cheng, X.; Hou, J. Mid-Infrared Ultra-Short Mode-Locked Fiber Laser Utilizing Topological Insulator Bi2Te3 Nano-Sheets as the Saturable Absorber. arXiv 2015, arXiv:1505.06322. [Google Scholar]
  41. Zhang, W.; Huang, H.; Wang, Z.; Kang, Q.; Liu, S.; Lu, S.; Wen, Q. Ternary Chalcogenide Ta2NiTe5 as Broadband Saturable Absorber for Passively Q-Switched Pulse Generation. Adv. Opt. Mater. 2024, 12, 2301261. [Google Scholar] [CrossRef]
Figure 1. Preparation process of the NbTe4-SA.
Figure 1. Preparation process of the NbTe4-SA.
Nanomaterials 15 00424 g001
Figure 2. Nonlinear optical characterization of NbTe4. (a) Experimental setup of the OA Z-scan system. (b) Experimental device of the P-scan system. Relationship between the normalized transmittance and input laser intensity at (c) 1030 nm and (d) 2000 nm, respectively. (Inset: corresponding OA Z-scan curve.) (e) Saturable absorption cure at 2790 nm.
Figure 2. Nonlinear optical characterization of NbTe4. (a) Experimental setup of the OA Z-scan system. (b) Experimental device of the P-scan system. Relationship between the normalized transmittance and input laser intensity at (c) 1030 nm and (d) 2000 nm, respectively. (Inset: corresponding OA Z-scan curve.) (e) Saturable absorption cure at 2790 nm.
Nanomaterials 15 00424 g002
Figure 3. Experimental setups of the lasers based on NbTe4-SA. (a) PQS solid-state lasers operating at 1.0 μm and 2.0 μm. (b) Mode-locked Er-doped fluoride fiber laser at 3.0 μm.
Figure 3. Experimental setups of the lasers based on NbTe4-SA. (a) PQS solid-state lasers operating at 1.0 μm and 2.0 μm. (b) Mode-locked Er-doped fluoride fiber laser at 3.0 μm.
Nanomaterials 15 00424 g003
Figure 4. Characterization of the NbTe4. (a) Simulated and experimental XRD patterns of the NbTe4 polycrystalline powder. (b) Raman spectrum of NbTe4 nanosheets. (c) EDS spectrum. (d) AFM and (e) corresponding 3D morphology of NbTe4 nanosheets. (f) Thickness distribution of NbTe4 nanosheets. (g,h) HRTEM images of exfoliated NbTe4 nanosheets at scales of 50 nm and 5 nm, respectively. (i) The transmission spectrum of the NbTe4 nanosheets.
Figure 4. Characterization of the NbTe4. (a) Simulated and experimental XRD patterns of the NbTe4 polycrystalline powder. (b) Raman spectrum of NbTe4 nanosheets. (c) EDS spectrum. (d) AFM and (e) corresponding 3D morphology of NbTe4 nanosheets. (f) Thickness distribution of NbTe4 nanosheets. (g,h) HRTEM images of exfoliated NbTe4 nanosheets at scales of 50 nm and 5 nm, respectively. (i) The transmission spectrum of the NbTe4 nanosheets.
Nanomaterials 15 00424 g004
Figure 5. Characterization of PQS Nd:YAG laser. (a) Dependence of output power on pump power (inset: output power stability over two hours). (b) Pulse width and repetition rate as functions of pump power. (c) Dependence of peak power and single-pulse energy on pump power. (d) Minimum pulse duration (inset: corresponding stable pulse trains). (e) Emission spectrum of the PQS laser. (f) Minimum pulse duration and RF spectrum of the PQS laser under pulse-pumping operation.
Figure 5. Characterization of PQS Nd:YAG laser. (a) Dependence of output power on pump power (inset: output power stability over two hours). (b) Pulse width and repetition rate as functions of pump power. (c) Dependence of peak power and single-pulse energy on pump power. (d) Minimum pulse duration (inset: corresponding stable pulse trains). (e) Emission spectrum of the PQS laser. (f) Minimum pulse duration and RF spectrum of the PQS laser under pulse-pumping operation.
Nanomaterials 15 00424 g005
Figure 6. Characterization of PQS Tm:YLF laser. (a) Dependence of output power on pump power (inset: output power stability over two hours). (b) Pulse width and repetition rate as functions of pump power. (c) Dependence of peak power and single-pulse energy on pump power. (d) Minimum pulse duration (inset: corresponding stable pulse trains). (e) Emission spectrum of the PQS laser. (f) Minimum pulse duration and RF spectrum of the PQS laser under pulse-pumping operation.
Figure 6. Characterization of PQS Tm:YLF laser. (a) Dependence of output power on pump power (inset: output power stability over two hours). (b) Pulse width and repetition rate as functions of pump power. (c) Dependence of peak power and single-pulse energy on pump power. (d) Minimum pulse duration (inset: corresponding stable pulse trains). (e) Emission spectrum of the PQS laser. (f) Minimum pulse duration and RF spectrum of the PQS laser under pulse-pumping operation.
Nanomaterials 15 00424 g006
Figure 7. Characterization of the mode-locked Er-doped fluoride fiber laser. (a) Output power as a function of pump power. (b) Stable mode-locked pulse trains at different time scales. (c) Optical spectrum of the mode-locked pulses. (d) Autocorrelation trace with Gaussian fitting. (e) RF spectrum of the mode-locked pulses (inset: broadband RF spectrum). (f) Stability of the average output power over two hours.
Figure 7. Characterization of the mode-locked Er-doped fluoride fiber laser. (a) Output power as a function of pump power. (b) Stable mode-locked pulse trains at different time scales. (c) Optical spectrum of the mode-locked pulses. (d) Autocorrelation trace with Gaussian fitting. (e) RF spectrum of the mode-locked pulses (inset: broadband RF spectrum). (f) Stability of the average output power over two hours.
Nanomaterials 15 00424 g007
Table 1. Performance comparison of lasers based on different broadband saturable absorbers.
Table 1. Performance comparison of lasers based on different broadband saturable absorbers.
SAλ
(μm)
Operation ModePulse Width (ns)Output Power (mW)Repetition Rate
(kHz)
Single
Pulse
Energy
(μJ)
Peak Power (W)Ref.
BP1.0PQS495223120.070.14[2]
2.0PQS178015119.257.844.4[30]
3.0CWML0.04261324,2700.0255607[31]
MoS21.0PQS9702277320.310.32[32]
2.0PQS80010048.092.082.6[33]
3.0PQS8061407022.48[34]
VTe21.0PQS195883500.251.29[35]
2.0PQS56323762.53.796.73
3.0PQS749402.8129.83.14.14
Graphene1.0PQS7532664360.610.81[36]
2.0PQS20803818.11.740.84[37]
3.0CWML0.0421825,4000.000716.87[1]
Bi2Te31.0PQS2000183151.51.210.6[38]
2.0PQS38227257.674.812[39]
3.0CWML0.02136042,4300.008400[40]
Ta2NiTe51.0PQS3302731921.424.36[41]
2.0PQS26007802235.113.5
3.0PQS8501821251.461.71
NbTe41.0PQS8643.09420.60.11.15This work
2.0PQS200013.8725.70.3880.16
3.0CWML0.01925723,6200.011570
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cai, Z.; Zhang, W.; Kang, Q.; Huang, H.; Xiang, X.; Lu, S.; Wen, Q. Quasi-1D NbTe4 for Broadband Pulse Generation from 1.0 to 3.0 μm: Bridging the Near- and Mid-Infrared. Nanomaterials 2025, 15, 424. https://doi.org/10.3390/nano15060424

AMA Style

Cai Z, Zhang W, Kang Q, Huang H, Xiang X, Lu S, Wen Q. Quasi-1D NbTe4 for Broadband Pulse Generation from 1.0 to 3.0 μm: Bridging the Near- and Mid-Infrared. Nanomaterials. 2025; 15(6):424. https://doi.org/10.3390/nano15060424

Chicago/Turabian Style

Cai, Zian, Wenyao Zhang, Qi Kang, Hongfu Huang, Xin Xiang, Shunbin Lu, and Qiao Wen. 2025. "Quasi-1D NbTe4 for Broadband Pulse Generation from 1.0 to 3.0 μm: Bridging the Near- and Mid-Infrared" Nanomaterials 15, no. 6: 424. https://doi.org/10.3390/nano15060424

APA Style

Cai, Z., Zhang, W., Kang, Q., Huang, H., Xiang, X., Lu, S., & Wen, Q. (2025). Quasi-1D NbTe4 for Broadband Pulse Generation from 1.0 to 3.0 μm: Bridging the Near- and Mid-Infrared. Nanomaterials, 15(6), 424. https://doi.org/10.3390/nano15060424

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop