Skip to content
BY 4.0 license Open Access Published by De Gruyter February 10, 2021

Enhanced broadband nonlinear optical response of TiO2/CuO nanosheets via oxygen vacancy engineering

  • Li Dong , Hongwei Chu ORCID logo EMAIL logo , Xiao Wang , Ying Li , Shengzhi Zhao and Dechun Li ORCID logo EMAIL logo
From the journal Nanophotonics

Abstract

Cupric oxide (CuO), as a transition metal oxide (TMO) semiconductor, has attracted tremendous attention for various applications. In the present work, we synthesize the CuO nanosheets modified by TiO2 nanoparticles via a facile, non-toxic two-step method. Subsequently, the morphology and the structures of CuO and TiO2/CuO nanocomposites are investigated. By utilizing the common Z-scan technology, broadband nonlinear optical (NLO) properties of the as-prepared CuO nanosheets and TiO2/CuO nanocomposites are demonstrated, elucidating the enhancement on the NLO response via the TiO2 dopant, which is attributed to the more oxygen vacancies and the formed p-n junctions. Furthermore, CuO nanosheets and TiO2/CuO nanocomposites are implemented to the passively Q-switched bulk lasers operating in the near-infrared (NIR) region, generating broadband ultrastable pulses. Ultimately, TiO2/CuO nanocomposites were intergrated in a passive mode-locking bulk laser for the first time, achieving stable mode-locked pulses and verifying its ultrafast optical response potential. Our results illustrate the tremendous prospects of the CuO nanosheets modified by oxygen vacancy engineering as a broadband NLO material in ultrafast photonics field.

1 Introduction

Two-dimensional (2D) nanomaterials have aroused unprecedented research enthusiasm on accounts of their eminent mechanical, electrical and optical properties [1], [2]. 2D materials have been extensively applied in the fields of electrodes, photovoltaics, catalysis and nonlinear optics including photonic coupling, plasmonics, optical modulators and switches [3], [4], [5], [6], [7], [8], [9], [10], [11], [12]. Besides, these nanomaterials have also successfully acted as saturable absorbers (SAs) for the generation of ultrafast pulses laser [13], [14], [15], benefiting from their distinguished third order optical nonlinearity, ultrafast carrier dynamics, and prominent broadband response characteristics [16], [17]. Lately, transition metal oxides (TMOs) have launched a new upsurge of research on nanomaterials owing to their strong phonon–electron coupling and electron–electron correlation, diverse optical and electrical properties [18]. Especially in the nonlinear optics, TMOs present itself by the high thermal and chemical stability, wide absorption bandwidth and strong optical nonlinearity [19], [20]. Among the multitudinous TMOs, cupric oxide (CuO) attracts numerous attentions owing to its remarkable superconductivity, high temperature stability, non-toxicity, fast recovery time of picosecond order and high third order nonlinearity coefficient [21], [22], [23]. Moreover, CuO has a monoclinic crystal structure with a narrower bandgap (Eg) of 1.2–1.4 eV [24], making it more appropriate in the near-infrared (NIR) region. Despite CuO nanomaterials possess strong NLO properties, the research on the optical modulation and nonlinear saturation absorption is rare [25], remaining blank in the broadband nonlinear photoresponse nature of CuO nanomaterials yet. In addition, no efforts are made to optimize the CuO nanomaterials in order to obtain a more stable and excellent modulation effect.

In a perfect CuO structure, the Cu atom is embraced by four coplanar O atoms, forming an approximately rectangular parallelogram structure [26]. However, during the preparation of TMOs, there will be oxygen vacancy (OV) defects and other point defects, disturbing the original physical structures and electronic states essentially. The OV defects commonly capture electrons by acting as the shallow donors to increase the density of carriers, thus improving the electrical conductivity, gas sensing performance and optical properties [27], [28], [29], [30]. Even more amazing is that the crystal structure is still extremely stable with these OV defects [27]. Based on these, some endeavors have been put to enhance the properties of the TMOs by manipulating and optimizing their defect structures and density artificially [30], [31]. Owing to the induced OV defects, the bandgap of the semiconducting TMOs would decrease, because the OV defects act as donors which can generate new energy levels in the band gap. Thus, much smaller photon energies are required to realize the electron transition from the valence band to the conduction band [31]. This provides an inspiration for us to reduce the band gap and improve the broad spectral band response performances in the CuO nanomaterials and other TMOs. Titanium (IV) oxide (TiO2) as a multifunctional material of TMOs, has been extensively extended to the photocatalyst, photovoltaic devices, and biosensors [32], [33], [34]. More importantly, studies have demonstrated that the Ti3+ in TiO2 can bring more OV defects [35], leading to the decrease of the bandgap [36], [37]. Hence, it is rational to anticipate that the composite of TiO2 nanoparticles on CuO hierarchical substrates can introduce more OV defects, reduce the Eg effectively, and improve the broadband response characteristics. In addition, TiO2 is a typical n-type semiconductor while CuO is a p-type semiconducting material. During the compositing process, the formation of p-n heterojunction would lead to electron–hole pairs separation effectively and increase the density of photogenerated carriers resulting to the variation of nonlinear absorption [38]. What’s more, TiO2 nanomaterials possess ultrafast carrier recovery time [36], high damage threshold [39], and giant third-order nonlinearity [40]. Accordingly, the formation of TiO2/CuO nanocomposites was expected to improve the ultrafast response and enhance the nonlinear characteristics of CuO nanosheets thanks to the stronger quantum local effect of TiO2 nanoparticles [41].

In this work, CuO nanosheets were fabricated by a facile water bath strategy and then TiO2/CuO nanocomposites were synthesized with CuO nanosheets as the precursor via the titanium tetrafluoride (TiF4) solution etching process. The uniformity of the dispersions and effectiveness of the composites were characterized. Then the broadband nonlinear optical absorption properties were elucidated via the open-aperture Z-scan technique at different operating wavelengths in NIR region. The results confirmed that the NLO properties were improved by TiO2 dopant in CuO nanosheets. Based on the large NLO responses, we implemented the CuO and TiO2/CuO nanocomposites as SAs in passively Q-switched (PQS) lasers operating at 1064, 1342 and 1878 nm to ulteriorly investigate the broadband nonlinear optical response characteristics. Finally, we utilized TiO2/CuO nanocomposites as a passive modulator in a mode-locked bulk laser for the first time, obtaining stable mode-locked pulses with a signal-to-noise (SNR) of ∼55 dB. Our results might open the gate to investigate the boosting effects of the oxygen vacancy defects in TMOs on NLO properties and broaden its application in nonlinear and ultrafast photonics.

2 Fabrication and characterization

In our experiment, we fabricated the CuO nanosheets by the method of a simple water bath under an alkaline atmosphere. First, 1.52 g Cu(NO3)2·3H2O and 0.443 g hexamethylene tetramine (HMTA) were independently dissolved into 60 mL pure deionized (DI) water, respectively, and the mixtures were vigorously stirred. Then we quickly poured the HMTA solution into the Cu(NO3)2 solution with the magnetic stirring at 50 °C for 1 h to make them mix uniformly. Subsequently, 12 mL 2.5 M sodium hydroxide (NaOH) solution was added dropwise into the above homogeneous mixed solution with vigorous magnetic stirring for 2 h at 80 °C, alone with the formation of black precipitates. Afterward, collecting the formed precipitates after several times centrifugation and washing in the DI water and ethanol, then drying it in a vacuum oven for 48 h at 60 °C. In this step, owing to the reducing HMTA, Cu2+ was produced as well as Cu+, leading to the OV defects in the prepared CuO powder.

After that, we took the prepared CuO powder (0.025 g) and dissolved the powder into 60 mL DI water. Then the mixture was ultrasonicated for 1 h and subsequently magnetically stirred for 0.5 h in order to get the well-distributed dispersion. Thereafter, 0.4 mL 0.02 M titanium tetrafluoride (TiF4) aqueous solution was dropped into the above solution with vigorous magnetic stirring for 2 h. Here, TiF4 not only worked as the precursors to hydrolyze the TiO2, but also served as the corrodent to etch the CuO nanosheets forming a rough surface, facilitating the attachment of the TiO2 nanoparticles. During the hydrolysis of TiF4, reducing H atom will change tetravalent Ti4+ to trivalent Ti3+ [42], generating abundant OV defects. Subsequently, we transferred the above mixture into a 150 mL sealed Teflon-lined autoclave under the atmosphere of 200 °C for 2 h. The ending products were collected, centrifuged and washed with DI water and ethanol for several times. Then the powder was dried in a vacuum oven for 48 h at 60 °C. By then, the pure CuO and the TiO2/CuO powders were successfully synthetized.

To fabricate the CuO nanosheets and the TiO2/CuO nanocomposites as SAs, the as-prepared powders were dissolved into alcohol and then the mixtures were ultrasonically shaken for 2 h at a speed of 5000 rpm. After that, decanting the supernatants before the centrifugation for 30 min at 10,000 rpm. Finally, 10 mL supernatant dispersions were spin coated onto 150 μm thick quartz glass substrate, and then dried in a vacuum oven for 24 h at 60 °C. It is worth noting that the operation consistencies of the CuO and the TiO2/CuO nanocomposites processing have been strictly guaranteed throughout all the process.

The crystalline structure of CuO is shown in Figure 1a, in which the Cu atom is neighbored with four O atoms. The scanning electron microscope (SEM) image indicated the as-prepared CuO nanosheets exhibited a shape of willow leaves with a length of ∼2 µm, as shown in Figure 1b. After the introduction of TiF4, the surface of the composites showed the morphology of “snowflakes falling on the leaves” (Figure 1c), which would generate massive nanoscale p-n junctions and change the characteristics of the carrier migration [42]. Moreover, the energy dispersive spectra (EDS) patterns (Figure 1d) were employed to determine the accurate chemical composition and the high purity of our above-prepared CuO and TiO2/CuO. Figure 1e showed the transmission electron microscopy (TEM) image of the pristine CuO, further verifying the elongated nanosheets morphology with a smooth surface, exactly coinciding with the SEM results. On the contrary, from Figure 1f, we could see that the surface of the composites TiO2/CuO had been roughened, which was conductive to the attachment of TiO2 nanoparticles with the average size of ∼10 nm. As well as the evidence of TiO2/CuO p-n heterojunction in TEM image, the resistances of CuO and TiO2/CuO were measured as 1.3 and 14.8 MΩ, respectively. The increased resistance of TiO2/CuO was attributed to the depletion layer during the formation of the p-n junction.

Figure 1: Microstructure and morphology characterization of cupric oxide (CuO) and TiO2/CuO nanocomposites.(a) Crystal structure of monoclinic CuO. (b) Scanning electron microscope (SEM) image of CuO. (c) SEM of TiO2/CuO. (d) Energy dispersive spectra (EDS) images of CuO and TiO2/CuO nanocomposites. (e) Transmission electron microscopy (TEM) of CuO. (f) TEM of TiO2/CuO nanocomposites.
Figure 1:

Microstructure and morphology characterization of cupric oxide (CuO) and TiO2/CuO nanocomposites.

(a) Crystal structure of monoclinic CuO. (b) Scanning electron microscope (SEM) image of CuO. (c) SEM of TiO2/CuO. (d) Energy dispersive spectra (EDS) images of CuO and TiO2/CuO nanocomposites. (e) Transmission electron microscopy (TEM) of CuO. (f) TEM of TiO2/CuO nanocomposites.

The crystal structures and phases of the pristine CuO and the composites TiO2/CuO were characterized by the X-ray diffraction (XRD), as illustrated in Figure 2. The XRD patterns of the as-synthesized pure CuO of the red line presented the strong diffraction peaks of (111), (111), (202), (113), (311), (113) planes, located at about 35.5°, 38.7°, 48.7°, 61.5°, 66.2° and 67.9°, respectively, agreeing well with the monoclinic CuO (JPCDS: 72-0629). For the composites of the purple line, in addition to the above-mentioned XRD patterns of CuO, there was a small diffraction peak of (101) located at 24.8°, which could be indexed to JPCDS No. 21-1272 of the anatase TiO2. The absence of other TiO2 diffraction peaks in nanocomposites could be ascribed to the low TiO2 concentration. There were no other peaks in the XRD spectra, revealing the high purity of the crystal phases.

Figure 2: X-ray diffraction (XRD) patterns of the prepared (a) CuO and (b) TiO2/CuO nanomaterials.
Figure 2:

X-ray diffraction (XRD) patterns of the prepared (a) CuO and (b) TiO2/CuO nanomaterials.

To further characterize the difference between the OV defects contained in the prepared CuO nanosheets and TiO2/CuO nanocomposites, the electron paramagnetic resonance (EPR) spectroscopy measurements were performed. As shown in Figure 3a, for both nanomaterials, the room-temperature EPR spectrums were centered at the magnetic field strength of 3516 Gs with g = 2.002, which could be ascribed to the surface unpaired electrons trapped on oxygen vacancies [43]. Moreover, the TiO2/CuO nanocomposites exhibited stronger EPR intensity than pure CuO nanosheets, confirming the more existence of OV defects [44]. To further elaborate the influence of the OV defects on the bandgap of CuO nanosheets, the UV-Vis-NIR spectrophotometer tests were performed. As depicted in Figure 3b, the bandgap could be reckoned by the horizontal intercepts of the tangents to the (αhν)1/2 versus the incident photon energy (hν). By extrapolating the linear part of the curve to the hν axis (vertical axis equal to zero), the Eg of CuO nanosheets was calculated to be 1.36 eV. Compared with CuO nanosheets, the bandgap of TiO2/CuO nanocomposites was exactly red-shifted to 1.24 eV due to the induced more OV defects.

Figure 3: (a) Electron paramagnetic resonance (EPR) spectrum. (b) Optical absorption spectrum of cupric oxide (CuO) nanosheets and TiO2/CuO nanomaterials.
Figure 3:

(a) Electron paramagnetic resonance (EPR) spectrum. (b) Optical absorption spectrum of cupric oxide (CuO) nanosheets and TiO2/CuO nanomaterials.

3 Experimental results and discussion

3.1 Broadband NLO responses

To scrutinize the broadband nonlinear optical properties of the CuO nanosheets and TiO2/CuO nanocomposites, the open-aperture (OA) Z-scan technique was performed at 1.06, 1.34 and 1.87 μm, respectively (see the details of the measurements in the Supporting Information). As shown in Figure 4, all normalized transmittances gradually increased to the maximum as the sample closed to the focus point (Z = 0) symmetrically, exhibiting the prominent saturable absorption properties of both CuO and TiO2/CuO nanocomposites. Notably, under the same experimental conditions, the empty substrates did not show any nonlinear response. For a two-level energy system, the experimental data can be theoretically fitted with the following equation [45]:

(1)T=m=0[q0(z,0)]m(m+1)1.5,mNq0(z,0)=βeffLeffI0(1+Z2/(Z02))
Figure 4: The nonlinear transmission curves and saturable absorption curves (the insets) of cupric oxide (CuO) and TiO2/CuO nanocomposites at (a) 1.06 μm, (b) 1.34 μm and (c) 1.87 μm.
Figure 4:

The nonlinear transmission curves and saturable absorption curves (the insets) of cupric oxide (CuO) and TiO2/CuO nanocomposites at (a) 1.06 μm, (b) 1.34 μm and (c) 1.87 μm.

In which Leff = (1 − e0)/α0 was the effective length while L was the actual length of the samples, I0 was the on-axis peak intensity. Moreover, α0 and βeff were the linear absorption and the nonlinear absorption coefficient, constituting the absorption model: α(I)=α0+βeffI. In addition, employing the prevalent single photon absorption model of SAs, we fitted the nonlinear transmittance evolutions curves as shown in the insets of Figure 4. It intuitively presented that the transmittance increased as the incident laser intensity increased until saturation, which was examined a typical feature of saturable absorbers. During the fitting, we assumed that the nanomaterials possessed instantaneous response with the following relationship [45]:

(2)T=1ΔTexp(I/IS)Tns

where, ΔT was the modulation depth, Is was the saturation intensity, and Tns was the nonsaturable losses, which could be fitted by the experimental data as summarized in Table 1. The deviations of the measured Is and βeff of the prepared CuO nanosheets from the previous reports [46], could be ascribed to the different thickness, size, layers and quality of the nanosheets determined by different synthetic techniques, which has also been elaborated in MoS2 samples [47]. Compared with the pure CuO nanosheets, the synthesized TiO2/CuO nanocomposites exhibited larger ΔT and βeff under the same experimental techniques and conditions. The enhancement of the nonlinearity would be ascribed to the strong quantum confinement and local field effects from dielectric confinement and saturable excitonic resonance [48] of the introduced TiO2 nanoparticles. Furthermore, the formed p-n junctions in TiO2/CuO nanocomposites would cause more separations of electron–hole pairs, leading to the increase of the carrier concentration [38]. Therefore, under laser irradiation, the nanomaterials needed to absorb more photons to reach the bleached state, which increased the nonlinear absorption coefficient and enhanced the nonlinearity. Moreover, by comparing with the NLO properties of several typical nanomaterials at ∼1, ∼1.3 and 2 μm, TiO2/CuO nanocomposites exhibited strong nonlinear absorption coefficient and lower saturation intensity. In summary, the above-mentioned results definitely manifested the potential of the CuO and TiO2/CuO nanocomposites for constructing broadband modulated pulsed lasers in NIR wavebands.

Table 1:

Nonlinear absorption characteristics of CuO and TiO2/CuO nanocomposites and the comparison with different nanomaterials.

Materialλ (nm)ΔT (%)Tns (%)Is (MW/cm2)βeff (cm/MW)Ref
CuO10649.3 ± 0.817.7 ± 1.20.56 ± 0.02−0.32 ± 0.01This work
134210.8 ± 1.010.1 ± 0.90.55 ± 0.01−0.41 ± 0.03
187811.4 ± 1.17.7 ± 0.80.76 ± 0.05−0.71 ± 0.09
TiO2/CuO106421.7 ± 1.518.4 ± 1.30.68 ± 0.04−0.35 ± 0.02This work
134211.8 ± 1.19.4 ± 0.70.4 ± 0.03−0.45 ± 0.05
187815.3 ± 1.310.3 ± 0.90.92 ± 0.07−0.83 ± 0.04
Tellurene10642440.29−0.058[49]
InSe13403.7826.061.21−0.28[50]
InSe19857.946.230.31−0.88[50]

3.2 Broadband modulation characteristics in all-solid PQS state laser

The marvelous broadband saturable absorption properties endowed CuO and TiO2/CuO nanocomposites with the potential application in passively Q-switched lasers. In this section, the Nd:GdVO4 (4F3/2 → 2I11/2), Nd:GdVO4 (4F3/2 → 2I13/2) and Tm:YLF crystals were utilized as the gain mediums to evaluate the capabilities for pulsed lasers at 1.06, 1.34, 1.87 μm, respectively. The detailed information of the schematic setups was illustrated in Figure S2 of the Supporting Information. In our case, all experimental conditions ensured strict consistency of CuO and TiO2/CuO nanomaterials.

3.2.1 Q-switching at 1.06 μm

For the Q-switching at 1.06 μm, stable operations were achieved with either CuO-SA or TiO2/CuO-SA integrated into the cavity. By removing the SAs from the resonator, only continuous wave (CW) operation was generated, and the threshold pump power was 0.49 W. On inserting the CuO and TiO2/CuO nanomaterials, the Q-switching thresholds were optimized to be 1.08 and 0.98 W, respectively. The PQS lasers characteristics for both cases were sketched in Figure 5. The output power of the CW laser and PQS lasers of CuO-SA or TiO2/CuO-SA increased quasi-linearly with the incident pump power, corresponding to the slope efficiencies of 51.8, 21.4 and 21.6%, respectively. Besides, the stabilities of these two PQS lasers were detected for 40 min at the maximum pump power of 1.96 W and both instabilities were <2% in root mean-squared error (RMSE). In addition, the TiO2/CuO-SA PQS laser exhibited a marginally higher stability. Higher slope efficiency and modulation stability can be foreseen by further cooling the SA and crystal with nitrogen and optimizing the mirror coating process.

Figure 5: Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.06 μm.(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.
Figure 5:

Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.06 μm.

(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.

Figure 5b depicted the pulse duration and repetition rate versus the incident pump power in detail. The pulse width decreased with the enhancement of the pump power, while the repetition rate displayed a contradictory tendency, which was typical of SAs in modulation characteristics and agreed well with the PQS theory. At the identical maximal pump power of 1.96 W, the TiO2/CuO-SA yielded shorter pulses of 120 ns at a higher repetition rate of 243 kHz, which demonstrated that the CuO doped with a small amount of TiO2 could compress the pulse duration more effectively and generate higher repetition rate at the same pump power level. Previous studies have shown that the carrier recovery time in CuO is about ∼50 ps [24], while TiO2 has an ultrafast saturation recovery time of 1.5 ps [36], which is more conducive to achieving high repetition rate output as an optical modulator. Compared with pure CuO (Figure 5c), the introduction of TiO2 accelerates the carrier recombination process and plays an important role on the faster Q-switching. Shorter Q-switching pulse duration could be anticipated with a larger modulation depth of TiO2/CuO-SA, by synthetizing felicitous number of layers and thickness [51]. Then we calculated the single pulse energy and the peak power, as depicted in Figure S3a and both of them presented the upward trends with increasing of the pump power. The temporal pulse trains of about 500 μs and single pulses were displayed in Figure 5d. No satellite pulses were observed and the pulse-to-pulse instability was calculated to be less than 3% RMSE, demonstrating the high stable modulation characteristics in 1 μm. In addition, the transverse movement of the saturable absorber could hardly impact the output performance, demonstrating that the spin-coating CuO nanosheets and TiO2/CuO nanocomposites on the substrates were rather homogeneous.

3.2.2 Q-switched laser at 1.34 μm

For the 4F3/2 → 2I13/2 transition of the Nd:GdVO4 operating at 1.34 μm, the Q-switched performances of pure CuO and TiO2/CuO nanocomposites as SAs were shown in Figure 6. Compared to 1.06 μm, the threshold pump power of the CW laser increased to 0.59 W at 1.3 μm, which due to the weaker emission cross section of Nd:GdVO4 at 1.3 μm. As shown in Figure 6a, at the highest incident pump power of 2.74 W, the maximum PQS average output power was 225 mW by utilizing TiO2/CuO-SA, corresponding to a higher slope efficiency of 22.6%. Under this pump power level, the minimum pulse duration of 158 ns and the highest repetition rate of 270.8 kHz was obtained with TiO2/CuO as SA (Figure 6b), corresponding to the single pulse energy of 830 nJ and the peak power of 5.25 W, as shown in Figure S3b in the Supporting information. Moreover, the transient pulse profiles were depicted in Figure 6c and d, clearly demonstrating that both CuO and TiO2/CuO nanocomposites could realize ultra-high stability modulation in 1.34 μm.

Figure 6: Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.34 μm.(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.
Figure 6:

Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.34 μm.

(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.

3.2.3 Q-switching at 1.9 µm

Subsequently, to further testify the capabilities of pulse response characteristics of CuO and TiO2/CuO nanocomposites at longer wavelengths in the mid-infrared region, we studied the CW and Q-switching regimes at ∼2 μm. The measured CW output spectrum was shown in Figure S4 with a central wavelength of 1878 nm. Under the incident pump power of 3.03 W, the obtained maximum PQS output powers were 155.9 and 166.8 mW with CuO and TiO2/CuO as SAs, respectively, illustrated in Figure 7a. Compared with that in the above-mentioned NIR laser, the reason for the lower output power here was that the water vapor has a strong absorption peak in the 2 μm band, causing the great losses in the laser resonator. Moreover, the gain of the Tm:YLF crystal is much lower owing to the smaller emission cross section. By increasing the incident pump power from 1.23 to 3.03 W, the pulse duration was compressed from 1286 to 210 ns with CuO-SA, while from 1321 to 179 ns with TiO2/CuO nanocomposites as the SA (Figure 7b), indicating that the composites CuO doped with little TiO2 had better pulse performance. The temporal pulses trains were displayed in Figure 7c and d, demonstrating the good amplitude stability. However, compared with that at 1.06 μm, the pulse stability was worse. Because the absorption of the Tm:YLF crystal was lower, the residual pump laser in the cavity resulted in a stronger thermal effect, affecting the stability of the modulation.

Figure 7: Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.87 μm.(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.
Figure 7:

Experimental performances of the passively Q-switched (PQS) lasers with cupric oxide (CuO) and TiO2/CuO nanocomposites at 1.87 μm.

(a) Continuous wave (CW) and Q-switched average output power. (b) Pulse width and repetition rate versus the incident pump power. (c), (d) Typical Q-switched temporal pulse profile and trains of CuO and TiO2/CuO lasers, respectively.

The results clearly showed that the PQS lasers based on TiO2/CuO nanocomposites exhibited more prominent modulation performance compared with the pristine CuO nanomaterials. The enhancement of the optical saturable absorption properties was attributed not only to the formed p-n junctions with the introduction of TiO2 nanoparticles, but also to the increased massive OV defects. Indeed, p-n heterojunctions in TiO2/CuO nanocomposites would cause electron–hole pairs separation, and increase the carrier concentration. While the massive OVs would modify the band gap, leading to the red shift of the absorption spectrum. The OV defects involved energy transitions were definitely beneficial to the NIR or even MIR optical absorption [52]. The synergy of the two effects effectively enhanced the nonlinear optical properties of TiO2/CuO nanocomposite in the near-infrared band. Therefore, the nonlinear optical absorption and modulation characteristics at 1.34 and 1.88 μm might come from the massive intermediate trap states.

Table 2 summarized the PQS bulk lasers performances with different NLO materials as SAs. Exactly as presented, compared with other nanomaterials, the synthesized TiO2/CuO nanocomposites featured better optical modulated properties and more potential to generate shorter pulse width and higher repetition rate. More importantly, the higher stability of the output pulses further evinced its remarkable manifestation in nonlinear photonics.

Table 2:

Broadband performance comparison of PQS bulk lasers with different NLO materials.

MaterialLaser crystalSA propertiesPulse output propertiesRef
ΔT (%)Isλ (nm)Width (ns)Rate (kHz)Power (mW)
InSeNd:YVO43.781.21 MW/cm21340520150420[50]
MoS2Tm,Ho:YAP210.8 MW/cm2212943555275[53]
BPNd:GdVO435.486.14 GW/cm2106449531222[54]
Bi2Se3Nd:YAG14.74.6 kW/cm2106457628.57150[55]
Bi2Se3/GTm:YAP6.60.44 MW/cm220002381082340[56]
TiO2/CuONd:GdVO421.70.68 MW/cm21064120243213This work
Nd:GdVO411.80.4 MW/cm21342158270225
Tm:YLF15.30.92 MW/cm21878215100156

4 Ultrafast photonics application

Sparked by the abovementioned prominent nonlinearity and laser performances of the TiO2/CuO nanocomposites, an ultrafast mode-locked laser established on a TiO2/CuO nanocomposites became a fascinating proposition and constituted the major spotlight of this section. As shown in Figure 8a, the construction of the mode-locked Nd:GdVO4 laser was a 1.44 m long Z-type resonator and the details were illustrated in the Supplementary materials.

Figure 8: (a) Experimental setup of the continuous wave mode-locking (CWML) laser based on TiO2/CuO nanocomposites at 1.06 μm. (b) Output power of CW and CWML regime versus the incident pump power.
Figure 8:

(a) Experimental setup of the continuous wave mode-locking (CWML) laser based on TiO2/CuO nanocomposites at 1.06 μm. (b) Output power of CW and CWML regime versus the incident pump power.

Without TiO2/CuO nanocomposites in the cavity, only continuous wave (CW) laser could be obtained and the threshold pump power of 0.59 W. When inserting the TiO2/CuO nanocomposites and adjusting the mirrors carefully, the stable CW mode-locking (CWML) operation was achieved. Figure 8b depicted the relationships of the CW and CWML output power versus the incident pump power. As shown in Figure 8b, when the incident pump power is larger than 4.8 W, the CWML operation can be self-started. Under the incident pump power of 6.7 W, the maximum CWML output power was 0.44 W. When increasing the pump power, the accumulated heat inside the sample would lead to the instability of mode-locking. Thereby, we did not further enhance the pump power during the experiments. Importantly, when reducing the pump power to the original level, stable mode-locked operation could be restored to the origin state, revealing the good thermal and chemical stability of the TiO2/CuO nanocomposites.

The detailed CWML output characteristics were illustrated in Figure 9. The interval between two adjacent pulses was 9.6 ns and the repetition rate was 103 MHz (Figure 9a), coinciding well with the total cavity length of 1.44 m. Assuming a Gauss intensity profile, the pulse duration of a single pulse was fitted to be 541 ps (Figure 9b). The spectrum of the CWML was centered at 1063.2 nm with the FWHM of 0.14 nm (Figure 9c). The time-bandwidth product (TBP) was calculated to be 20.1, great than the typical value of Gauss pulse profile (TBP = 0.44), manifesting the large frequency chirping in the cavity. Accordingly, shorter mode-locked pulse could be anticipated by adding prism pairs to compensate the group delay dispersion (GDD) in the cavity and fabricating the materials with larger modulation depth. For the radio frequency (RF) spectrum, the fundamental peak was located at 103 MHz with a signal-to-noise ratio (SNR) of ∼55 dB (Figure 9d), matching well with the cavity length, demonstrating the great potential of the ultrafast optical modulation properties of the fabricated TiO2/CuO nanocomposites.

Figure 9: Experimental performances of the continuous wave mode-locking (CWML) lasers with TiO2/CuO nanocomposites at 1.06 μm.(a) Pulse train. (b) Single pulse profile. (c) Spectrum. (d) RF spectrum with a span of 800 MHz.
Figure 9:

Experimental performances of the continuous wave mode-locking (CWML) lasers with TiO2/CuO nanocomposites at 1.06 μm.

(a) Pulse train. (b) Single pulse profile. (c) Spectrum. (d) RF spectrum with a span of 800 MHz.

5 Conclusion

In conclusion, the pure CuO nanosheets and TiO2/CuO nanocomposites were synthesized based on a facile hydrothermal method and a TiF4 solution etching process, respectively. Subsequently, the Z-scan experimental results demonstrated that both CuO and TiO2/CuO nanomaterials possessed outstanding saturable absorption performances in NIR spectra range, but ensuring the consistency of the experimental conditions, the TiO2/CuO nanocomposites had the larger nonlinear absorption coefficient and modulation depth. In virtue of these, the broadband PQS lasers and ultrafast mode-locking laser utilizing CuO and TiO2/CuO nanocomposites as SAs were constructed for the first time, which output results once again confirmed the prominent characteristics of the TiO2/CuO nanocomposites. The promoted modulation performances could be assigned to massive OV defects and the formed p-n junctions, resulting to the decreased band gap, the red shift of absorption wavelength and strong broadband NLO response characteristics. Our work is likely to trigger more thinking about the effects of the OV defects on the bandgap or the modulation characteristics of the nanomaterials, especially the TMOs.


Corresponding authors: Hongwei Chu and Dechun Li, School of Information Science and Engineering, Shandong University, Qingdao266237, China, E-mail: ,

Funding source: Fundamental Research Fund of Shandong University

Award Identifier / Grant number: 2018TB044

Award Identifier / Grant number: 12004213

Award Identifier / Grant number: 21872084

Award Identifier / Grant number: 62001189

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: The National Natural Science Foundation of China (12004213, 21872084, 62001189); Fundamental Research Fund of Shandong University (2018TB044). H. C. would like to thank the financial support from the Young Scholar Program of Shandong University.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] F. Bonaccorso, Z. Sun, T. Hasan, et al.., “Graphene photonics and optoelectronics,” Nat. Photonics., vol. 4, pp. 611–622, 2010, https://doi.org/10.1038/nphoton.2010.186.Search in Google Scholar

[2] Y. Song, Z. Liang, X. Jiang, et al.., “Few-layer antimonene decorated microfiber: ultra-short pulse generation and all-optical thresholding with enhanced long term stability,” 2D Materials, vol. 4, p. 045010, 2017, https://doi.org/10.1088/2053-1583/aa87c1.Search in Google Scholar

[3] Z. Xiao, Z. Yang, L. Zhou, et al.., “Highly conductive porous transition metal dichalcogenides via water steam etching for high-performance lithium–sulfur batteries,” ACS Appl. Mater. Interfaces, vol. 9, pp. 18845–18855, 2017, https://doi.org/10.1021/acsami.7b04232.Search in Google Scholar PubMed

[4] M. Buscema, D. Groenendijk, G. Steele, et al.., “Photovoltaic effect in few-layer black phosphorus PN junctions defined by local electrostatic gating,” Nat. Commun., vol. 5, pp. 4651–4656, 2014, https://doi.org/10.1038/ncomms5651.Search in Google Scholar PubMed

[5] T. Jiang, R. Miao, J. Zhao, et al., “Electron-phonon coupling in topological insulator Bi2Se3 thin films with different substrates,” Chin. Opt. Lett., vol. 17, p. 020005, 2019.10.3788/COL201917.020005Search in Google Scholar

[6] Y. Song, Y. Chen, X. Jiang, et al.., “Nonlinear few-layer antimonene-based all-optical signal processing: ultrafast optical switching and high-speed wavelength conversion,” Adv. Opt. Mater., vol. 6, p. 1701287, 2018, https://doi.org/10.1002/adom.201701287.Search in Google Scholar

[7] L. Zeng, D. Wu, J. Jie, et al.., “Van der Waals epitaxial growth of mosaic‐like 2D platinum ditelluride layers for room‐temperature mid‐infrared photodetection up to 10.6 µm,” Adv. Mater., vol. 32, p. 2004412, 2020, https://doi.org/10.1002/adma.202004412.Search in Google Scholar PubMed

[8] L. Zeng, D. Wu, S. Lin, et al.., “Controlled synthesis of 2D palladium diselenide for sensitive photodetector applications,” Adv. Funct. Mater., vol. 29, p. 1806878, 2019, https://doi.org/10.1002/adfm.201806878.Search in Google Scholar

[9] L. Zeng, S. Lin, Z. Li, et al.., “Fast, self‐driven, air‐stable, and broadband photodetector based on vertically aligned PtSe2/GaAs heterojunction,” Adv. Funct. Mater., vol. 28, p. 1705970, 2018, https://doi.org/10.1002/adfm.201705970.Search in Google Scholar

[10] L. Zeng, Q. Chen, Z. Zhang, et al.., “Multilayered PdSe2/perovskite Schottky junction for fast, self‐powered, polarization‐sensitive, broadband photodetectors, and image sensor application,” Adv. Sci., vol. 6, p. 1901134, 2019, https://doi.org/10.1002/advs.201901134.Search in Google Scholar PubMed PubMed Central

[11] D. Wu, J. Guo, J. Du, et al.., “Highly polarization-sensitive, broadband, self-powered photodetector based on graphene/PdSe2/Germanium heterojunction,” ACS Nano, vol. 13, pp. 9907–9917, 2019, https://doi.org/10.1021/acsnano.9b03994.Search in Google Scholar PubMed

[12] L. Zeng, S. Lin, Z. Lou, et al.., “Ultrafast and sensitive photodetector based on a PtSe2/silicon nanowire array heterojunction with a multiband spectral response from 200 to 1550 nm,” NPG Asia Mater., vol. 10, pp. 352–362, 2018, https://doi.org/10.1038/s41427-018-0035-4.Search in Google Scholar

[13] W. Liu, M. Liu, J. Yin, et al.., “Tungsten diselenide for all-fiber lasers with the chemical vapor deposition method,” Nanoscale, vol. 10, pp. 7971–7977, 2018, https://doi.org/10.1039/c8nr00471d.Search in Google Scholar

[14] L. Li, R. Lv, J. Wang, et al.., “Optical nonlinearity of ZrS2 and applications in fiber laser,” Nanomaterials, vol. 9, pp. 315–324, 2019, https://doi.org/10.3390/nano9030315.Search in Google Scholar PubMed PubMed Central

[15] H. Mu, S. Lin, Z. Wang, et al.., “Black phosphorus–polymer composites for pulsed lasers,” Adv. Opt. Mater., vol. 3, pp. 1447–1453, 2015, https://doi.org/10.1002/adom.201500336.Search in Google Scholar

[16] S. Tani, F. Blanchard, and K. Tanaka, “Ultrafast carrier dynamics in graphene under a high electric field,” Phys. Rev. Lett., vol. 109, p. 166603, 2020.10.1103/PhysRevLett.109.166603Search in Google Scholar PubMed

[17] L. Du, D. Lu, J. Li, et al.., “Broadband nonlinear optical response of single-crystalline bismuth thin film,” ACS Appl. Mater. Interfaces, vol. 11, pp. 35863–35870, 2019, https://doi.org/10.1021/acsami.9b10354.Search in Google Scholar PubMed

[18] H. Fan, G. You, Y. Li, et al.., “Shape-controlled synthesis of single-crystalline Fe2O3 hollow nanocrystals and their tunable optical properties,” J. Phys. Chem. C, vol. 113, pp. 9928–9935, 2009, https://doi.org/10.1021/jp9020883.Search in Google Scholar

[19] L. Dong, H. Chu, Y. Li, et al.., “Few-layered α-Fe2O3 nanoflake saturable absorber for stable MIR pulse generation,” Opt. Mater. Express, vol. 10, pp. 2313–2320, 2020, https://doi.org/10.1364/ome.404316.Search in Google Scholar

[20] M. Ahmadi, R. Younesi, and M. Guinel, “Synthesis of tungsten oxide nanoparticles using a hydrothermal method at ambient pressure,” J. Mater. Res., vol. 29, pp. 1424–1430, 2014, https://doi.org/10.1557/jmr.2014.155.Search in Google Scholar

[21] F. Gao, L. Zhu, H. Li, and H. Xie, “Hierarchical flower-like CuO film: one-step room temperature synthesis, formation mechanism and excellent optoelectronic properties,” Mater. Res. Bull., vol. 93, pp. 342–351, 2017, https://doi.org/10.1016/j.materresbull.2017.05.033.Search in Google Scholar

[22] A. Chen, G. Yang, H. Long, et al.., “Optical limiting properties in copper oxide thin films under a high-repetition-rate femtosecond laser,” Mater. Lett., vol. 91, pp. 319–322, 2013, https://doi.org/10.1016/j.matlet.2012.09.111.Search in Google Scholar

[23] H. Shanshool, M. Yahaya, W. Yunus, and I. Abdullah, “Influence of CuO nanoparticles on third order nonlinearity and optical limiting threshold of polymer/ZnO nanocomposites,” Opt. Quant. Electron., vol. 49, pp. 18–21, 2017, https://doi.org/10.1007/s11082-016-0830-5.Search in Google Scholar

[24] B. Born, S. Gagnon, J. Krupa, et al.., “Ultrafast charge-carrier dynamics of copper oxide nanocrystals,” ACS Photonics, vol. 3, pp. 2475–2481, 2016, https://doi.org/10.1021/acsphotonics.6b00717.Search in Google Scholar

[25] S. Abdulhameed, H. Wadi, and A. Hadi, “Ultrashort pulse generation with an erbium-doped fiber laser ring cavity based on a copper oxide saturable absorber,” Appl. Opt., vol. 57, pp. 5180–5185, 2018.10.1364/AO.57.005180Search in Google Scholar PubMed

[26] W. Ching, Y. Xu, and K. Wong, “Ground-state and optical properties of Cu2O and CuO crystals,” Phys. Rev. B, vol. 40, p. 7684, 1989, https://doi.org/10.1103/physrevb.40.7684.Search in Google Scholar PubMed

[27] G. Wang, Y. Ling, and Y. Li, “Oxygen-deficient metal oxide nanostructures for photoelectrochemical water oxidation and other applications,” Nanoscale, vol. 4, pp. 6682–6691, 2012, https://doi.org/10.1039/c2nr32222f.Search in Google Scholar PubMed

[28] L. Zhang, Q. Fang, Y. Huang, et al.., “Oxygen vacancy enhanced gas-sensing performance of CeO2/graphene heterostructure at room temperature,” Anal. Chem., vol. 90, pp. 9821–9829, 2018, https://doi.org/10.1021/acs.analchem.8b01768.Search in Google Scholar PubMed

[29] D. Wang, J. Sun, X. Cao, et al.., “High-performance gas sensing achieved by mesoporous tungsten oxide mesocrystals with increased oxygen vacancies,” J. Mater. Chem. A, vol. 17, pp. 8653–8657, 2013, https://doi.org/10.1039/c3ta11506b.Search in Google Scholar

[30] M. Jeem, L. Zhang, J. Ishioka, et al.., “Tuning optoelectrical properties of ZnO nanorods with excitonic defects via submerged illumination,” Nano Lett., vol. 17, pp. 2088–2093, 2017, https://doi.org/10.1021/acs.nanolett.7b00324.Search in Google Scholar PubMed

[31] F. Gu, C. Li, D. Han, and Z. Wang, “Manipulating the defect structure (VO) of In2O3 nanoparticles for enhancement of formaldehyde detection,” Appl. Mater. Interfaces, vol. 10, pp. 933–942, 2018, https://doi.org/10.1021/acsami.7b16832.Search in Google Scholar PubMed

[32] C. Grigori and C. Sim, “Study on TiO2 thin films grown by advance pulsed laser deposition on ITO,” Thin Solid Films, vol. 518, pp. 1314–1317, 2019.10.1016/j.tsf.2009.05.049Search in Google Scholar

[33] J. Liu, X. Fu, D. Cao, et al.., “Stacked graphene-TiO2 photoanode via electrospray deposition for highly efficient dye-sensitized solar cells,” Org. Electron., vol. 23, pp. 158–163, 2015, https://doi.org/10.1016/j.orgel.2015.04.021.Search in Google Scholar

[34] D. Zhang, J. Liu, C. Jiang, et al.., “High-performance sulfur dioxide sensing properties of layer-by-layer self-assembled titania-modified graphene hybrid nanocomposite,” Sens. Actuators B Chem., vol. 245, pp. 560–567, 2017, https://doi.org/10.1016/j.snb.2017.01.200.Search in Google Scholar

[35] X. Chen, L. Liu, Z. Liu, et al.., “Properties of disorder-engineered black titanium dioxide nanoparticles through hydrogenation,” Sci. Rep., vol. 3, pp. 1510–1516, 2013, https://doi.org/10.1038/srep01510.Search in Google Scholar PubMed PubMed Central

[36] H. Elim, W. Ji, A. Yuwono, et al.., “Ultrafast optical nonlinearity in poly (methylmethacrylate)-TiO2 nanocomposites,” Appl. Phys. Lett., vol. 82, pp. 2691–2693, 2003, https://doi.org/10.1063/1.1568544.Search in Google Scholar

[37] M. Mehta, N. Kodan, S. Kuma, et al.., “Hydrogen treated anatase TiO2: a new experimental approach and further insights from theory,” J. Mater. Chem. A, vol. 4, pp. 2670–2681, 2016, https://doi.org/10.1039/c5ta07133j.Search in Google Scholar

[38] T. Atabaev, D. Hun, and N. Hong, “Fabrication of TiO2/CuO photoelectrode with enhanced solar water splitting activity,” Funct. Mater. Lett., vol. 10, p. 1750084, 2017, https://doi.org/10.1142/s1793604717500849.Search in Google Scholar

[39] J. Yao, Z. Fan, Y. Jin, et al., “Investigation of damage threshold to TiO2 coatings at different laser wavelength and pulse duration,” Thin Solid Films, vol. 516, pp. 1237–1241, 2018.10.1016/j.tsf.2007.06.058Search in Google Scholar

[40] S. Das, C. Schwanke, A. Pfuch, et al.., “Highly efficient THG in TiO2 nanolayers for third-order pulse characterization,” Opt. Express, vol. 19, p. 16985, 2011, https://doi.org/10.1364/oe.19.016985.Search in Google Scholar

[41] N. Herron and D. Thorn, “Nanoparticles: uses and relationships to molecular cluster compounds,” Adv. Mater., vol. 10, p. 1173, 1998, https://doi.org/10.1002/(sici)1521-4095(199810)10:15<1173::aid-adma1173>3.0.co;2-6.10.1002/(SICI)1521-4095(199810)10:15<1173::AID-ADMA1173>3.0.CO;2-6Search in Google Scholar

[42] X. Wang, Y. Li, Z. Li, et al.., “Highly sensitive and low working temperature detection of trace triethylamine based on TiO2 nanoparticles decorated CuO nanosheets sensors,” Sens. Actuators B Chem., vol. 301, p. 127019, 2019, https://doi.org/10.1016/j.snb.2019.127019.Search in Google Scholar

[43] H. Kaftelen, K. Ocakoglu, R. Thomann, et al.., “EPR and photoluminescence spectroscopy studies on the defect structure of ZnO nanocrystals,” Phys. Rev. B, vol. 86, p. 9, 2012, https://doi.org/10.1103/physrevb.86.014113.Search in Google Scholar

[44] F. Lei, Y. Sun, K. Liu, et al.., “Oxygen vacancies confined in ultrathin indium oxide porous sheets for promoted visible-light water splitting,” J. Am. Chem. Soc., vol. 136, pp. 6826–6829, 2014, https://doi.org/10.1021/ja501866r.Search in Google Scholar

[45] H. Pan, H. Chu, X. Wang, et al.., “Nonlinear optical features of zeolitic imidazolate framework-67 nanocrystals for mid-infrared pulse generation,” Cryst. Growth Des., vol. 20, pp. 6683–6690, 2020, https://doi.org/10.1021/acs.cgd.0c00842.Search in Google Scholar

[46] A. Chen, G. Yang, H. Long, et al., “Nonlinear optical properties of laser deposited CuO thin films,” Thin Solid Films, vol. 7, pp. 4277–4280, 2009.10.1016/j.tsf.2008.11.139Search in Google Scholar

[47] R. Woodward, E. Kelleher, R. Howe, et al.., “Tunable Q-switched fiber laser based on saturable edge-state absorption in few-layer molybdenum disulfide (MoS2),” Opt. Express, vol. 22, pp. 31113–31122, 2014, https://doi.org/10.1364/oe.22.031113.Search in Google Scholar

[48] D. Chemla and D. Miller, “Mechanism for enhanced optical nonlinearities and bistability by combined dielectric–electronic confinement in semiconductor microcrystallites,” Opt. Lett., vol. 11, pp. 522–524, 1986, https://doi.org/10.1364/ol.11.000522.Search in Google Scholar

[49] F. Zhang, G. Liu, Z. Wang, et al.., “Broadband nonlinear absorption properties of two-dimensional hexagonal tellurene nanosheets,” Nanoscale, vol. 11, p. 17058, 2019, https://doi.org/10.1039/c9nr03701b.Search in Google Scholar

[50] H. Pan, L. Cao, H. Chu, et al.., “Broadband nonlinear optical response of InSe nanosheets for the pulse generation from 1 to 2 μm,” ACS Appl. Mater. Interfaces, vol. 11, pp. 48281–48289, 2019, https://doi.org/10.1021/acsami.9b18632.Search in Google Scholar

[51] J. Zhang, T. Jiang, T. Zhou, et al.., “Saturated absorption of different layered Bi2Se3 films in the resonance zone,” Photonics Res., vol. 6, pp. C8–C14, 2018, https://doi.org/10.1364/prj.6.0000c8.Search in Google Scholar

[52] G. Wang, H. Wang, Y. Ling, et al.., “Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water splitting,” Nano Lett., vol. 11, pp. 3026–3033, 2011, https://doi.org/10.1021/nl201766h.Search in Google Scholar PubMed

[53] C. Luan, X. Zhang, K. Yang, et al.., “High-peak power passively Q-switched 2-µm laser with MoS2 saturable absorber,” IEEE J. Sel. Top. Quantum Electron., vol. 23, p. 1600105, 2017, https://doi.org/10.1109/jstqe.2016.2537980.Search in Google Scholar

[54] R. Zhang, Y. Zhang, H. Yu, et al.., “Broadband black phosphorus optical modulator in the spectral range from visible to mid‐infrared,” Adv. Opt. Mater., vol. 3, pp. 1787–1792, 2015, https://doi.org/10.1002/adom.201500298.Search in Google Scholar

[55] Y. Sun, C. Lee, J. Xu, et al., “Passively Q-switched tri-wavelength Yb3+:GdAl3(BO3)4 solid-state laser with topological insulator Bi2Te3 as saturable absorber,” Photonics Res., vol. 3, p. A97, 2015. https://doi.org/10.1364/prj.3.000a97.Search in Google Scholar

[56] Z. You, Y. Sun, D. Sun, et al.., “High performance of a passively Q-switched mid-infrared laser with Bi2Te3/graphene composite SA,” Opt. Lett., vol. 42, pp. 871–874, 2017, https://doi.org/10.1364/ol.42.000871.Search in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2020-0649).


Received: 2020-12-10
Accepted: 2021-01-25
Published Online: 2021-02-10

© 2021 Li Dong et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2020-0649/html
Scroll to top button