Introduction

Water pollution is an important issue of concern in many countries. Many dyes in textile printing or dyeing wastewater contain aromatic compounds, which are chemically stable and harmful to human health1. Because of the adverse environmental impacts of these refractory organic compounds, it is necessary to develop new methods to degrade them. Photocatalysis is a promising technique for photodegradation of hazardous chemicals in wastewater2. Titanium dioxide (TiO2) and TiO2 based nanostructures are being used for photocatalytic degradation of organic pollutants to protect environment3. With the advantages of low cost, large specific surface area, high oxidizing power, and good chemical stability, TiO2 has become one of the most promising candidates for photocatalysis4. However, the catalytic performance of TiO2 is severely limited by the large band gap (3.2 eV), low photon utilization rate of solar energy (about 5%), and high recombination rate for photogenerated electron–hole pairs5,6. On the other hand, the industrial treatment of wastewater containing various organic pollutants using TiO2-photocatalyst is not common due to low efficiency of photodegradation7. To solve this problem, several methods have been used to improve the photocatalytic efficiency of TiO2. It has been reported that in a suitable heterostructured system, the presence of heterojunction changes the energy band positions and their inclination on the interface to accelerate the migration of photogenerated charge carriers, and eventually enhance the efficiency of photocatalyst8.

Perylenetetracarboxylic acid diimides (PDIs) are cheap organic dyes with high photothermal stability and strong absorption in the visible region9,10. As a typical n-type semiconductor, PDIs have high electron mobility and electron affinity [LUMO] due to its strong π–π stacking between the conjugated π bonds11,12. PDIs materials have been used in photocatalysis because it can improve the light absorption and decrease the photo-generated electron–hole recombination13. PDIs have been introduced into TiO2, ZnO, and other photocatalysts to improve the photocatalytic performance14,15. However, such composite materials still have some defects such as quick recombination of photoexcited electron–hole pairs and photoetching, which lead to low efficiency of visible light driven photodegradation16. Therefore, it is necessary to further improve their photocatalytic activities and stabilities.

Molybdenum disulfide (MoS2) is a rapidly rising two-dimensional layered material, which has aroused enormous scientific interest in developing new MoS2-based materials for rich potential applications17. MoS2 with different structures such as nanoribbon and nanosheet has been successfully synthesized18,19. It has unique physical properties that distinguish it from other materials and has been used in many applications such as lubricant, photovoltaic, and photocatalysis20,21. Hu et al. synthesized bulk, nano-slice, and nano-ball MoS2 by precipitation method, and found that the photocatalytic efficiency of nano-slice MoS2 towards methyl orange (90%) was higher than others22. James et al. synthesized MoS2 nanoparticles through thermally decomposing method, and the material was used for photocatalytic degradation of methylene blue (MB) with an efficiency of 30%23. Xu et al. prepared flower-like MoS2 nanopowders through hydrothermal method for photocatalytic degradation of Rhodamine B (RhB) and an efficiency of 15% was achieved24. In addition, the good conductivity of monolayer MoS2 can efficiently separate the electron–hole pairs and enhance the photocatalytic activity. Hence, MoS2 can be used in TiO2 photocatalyst to improve its photocatalytic performance. For instance, Behzad Pourabbas et al. reported the photo-oxidation of phenol by MoS2/TiO2 hybrids. Suddhasatwa Basu et al. demonstrated the ability to degrade RhB by MoS2/TiO2 nanocomposite25. Giri et al. reported the photo-oxidation of RhB by MoS2/TiO2 hybrids26. Du et al. reported a MoS2/CdS/TiO2 photocatalyst which has high photocatalytic activity and stability27. This technique produces powerful non-selective oxidants, such as superoxide radicals and hydroxyl radicals, which can degrade and mineralize a wide variety of pollutants28. The produced electrons-holes would recombine in pure TiO2, which limited its application in the photodegradation. By preventing electron/hole recombination and semiconductors aggregation, multi-component support can improve the photoreactivity of the obtained composite29.

In this work, a novel ternary photocatalyst: NH2-PDI/TiO2/MoS2 was synthesized by hydrothermal method (Scheme 1). The CB/VB (conduction band/valence band) of TiO2, NH2-PDI, and MoS2 are evaluated to be − 1.26/1.72 eV, − 1.56/0.8 eV, and -0.75/1.00 eV, respectively. These matched potentials are suitable for rapid charge separation between these semiconductors. The composite showed enhanced activity in the photodegradation of Methylene Blue (MB) and amoxicillin (AMX).

Scheme 1
scheme 1

Molecular structure of 1-aminoperylene diimide (NH2-PDI).

Experimental

Materials and methods

All other chemicals were analytically pure, purchased from commercial sources and used without further purification. Distilled water was used throughout the experiment. FT-IR spectrum has been studied by a Bruker Tensor-27 spectrophotometer. The 1H NMR and mass spectrum of NH2-PDI were measured on a Bruker Advance 400 spectrometer and a Bruker Maxis UHR-TOF mass spectrometer, respectively. Scanning electron microscopy (SEM) images were obtained on a Sigma, Zeiss microscope equipped with an energy dispersive X-ray (EDX) spectrometer. The specific surface area of the catalyst has been analyzed using nitrogen adsorption at 77 K applying the Brunauer–Emmett–Teller (BET) method using a micrometrics ASAP 2020 V3.00 H. X-ray diffraction (XRD) measurements have been performed using a Rigaku R-AXIS RAPID X-ray diffractometer. Crystallinity and phase composition of the as-synthesized composite have been confirmed from the Lab Ram HR800, Jobin Yvon micro-Raman measurement. The ultraviolet–visible diffuse reflectance absorption spectra (DRS) have been measured on a Shimadzu model UV–Vis diffuse reflectance spectrometer. X-ray photoelectron spectroscopy (XPS) has been performed on a KRATOS model XSAM800 instrument. Fluorescence measurements were performed at a Hitachi FL-4500 spectrometer with 290 nm excitation wavelength at room temperature. The Mott-Schottky plots, photocurrent, and electrochemical impedance spectroscopy (EIS) were measured on a CHI760E electrochemical workstation using a three-electrode system. Platinum wire as counter electrode, ITO deposited by photocatalyst was used as working electrodes (50 mg of photocatalyst was dispersed in 2 mL of ethanol and grounded for 20 min, then it was uniformly dropped onto the FTO glass and stay still overnight to dry off the ethanol.), saturated calomel as reference electrode, and 0.1 M Na2SO4 aqueous solution was used as electrolyte. The EIS was performed at an open circuit potential at a frequency of 0–10,000 Hz, and the photoelectric response of the sample was measured at 0.0 V. Photoelectrochemical properties were measured with a 300 W xenon lamp as the light source.

Synthesis and characterization of NH2-PDI

The compound NH2-PDI was synthesized according to literature method30. The synthetic route along with the characterization data of NH2-PDI (Fig. S-1) are reported in the Supplementary Methods.

Preparation of TiO2 and NH2-PDI/TiO2

The NH2-PDI/TiO2 composite was prepared by the hydrothermal synthesis method. First, tetrabutyl titanate (10.0 mL) was dissolved in anhydrous ethanol (20.0 mL). Then, 5.0 mL of distilled water was slowly added under vigorous stirring. Later, NH2-PDI (0.01 g) was dissolved in dichloromethane (5.0 mL) and added to the solution under sonication. The resulting mixture was stirred for 12 h and then transfer to the hydrothermal kettle for 3 h at 200 °C. The precipitate was filtered, washed thoroughly with distilled water, and dried in an air oven at 100 °C for 4 h. This catalyst contained 0.1 wt% of NH2-PDI. Pure TiO2 was prepared by the same method.

Preparation of MoS2

MoS2 was synthesized by hydrothermal route using Na2MoO4·2H2O and NH2CSNH2 as Mo and S sources, respectively. Briefly, Na2MoO4·2H2O (2.0 g) and NH2CSNH2 (8.0 g) was dissolved in 60 mL of water under vigorous stirring for 30 min. The mixture was kept in a hydrothermal kettle at 200 °C for 24 h. The resulting black precipitate was washed several times with ethanol and water to further remove impurities and contaminants followed by a centrifugation and drying process at 60 °C for 6 h to obtain MoS2 nanoarchitectures.

Preparation of MoS2/TiO2 and NH2-PDI/TiO2/MoS2

The NH2-PDI/TiO2/MoS2 composite was fabricated by loading MoS2 onto the preformed NH2-PDI/TiO2 solid solutions. Hydrothermal synthesis of NH2-PDI/TiO2/MoS2 composite was prepared as follows: 3.0 g NH2-PDI/TiO2 and 0.06 g of MoS2 were dispersed in 5 mL water. The resulting mixture was stirred for 24 h and then transferred to a hydrothermal kettle at 200 °C for 4 h. The prepared composite material was filtered, washed thoroughly with distilled water, and dried in an air oven at 100 °C for 4 h. NH2-PDI/TiO2/MoS2 was formed as a gray powder. This catalyst contained 0.1 wt% NH2-PDI and 2.0 wt% MoS2. MoS2/TiO2, NH2-PDI/TiO2/1%MoS2, and NH2-PDI/TiO2/3%MoS2 were prepared using the same procedure.

Photocatalytic activity tests

The photocatalytic experiments were carried out in a photochemical reactor (PhchemIII, Beijing China NBeT) consisting of a 500 W xenon lamp (XE-JY500). The reaction chamber is equipped with 50 mL capacity reaction glass tubes and a magnetic stirrer. The specially designed reflector was made of highly polished aluminum and a built-in cooling fan. The light exposure length is 230 mm.

Methylene Blue (MB) and amoxicillin (AMX) were selected to test the photodegradation activity of the prepared photocatalysts. First, 50 mg of synthesized sample was suspended in 50 ml of 10 mg/L MB aqueous solution. The visible light source was obtained by a 500 W xenon lamp. Before irradiation, the suspension solutions were stirred magnetically for 120 min in the dark to achieve adsorption–desorption equilibrium. The samples were withdrawn at given time intervals and the photocatalyst was removed by centrifugation. A UV–vis spectrophotometer at 664 nm and 198 nm was used to measure the absorbance of MB and AMX solutions, respectivly.

Results and discussion

Figure 1 shows SEM images of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2. Figure 1(a) shows that TiO2 forms uniform elongated spherical particles with a length of about 300 nm and a width of about 200 nm. As can be seen from Fig. 1(b), NH2-PDI forms nanorods structure. The average width is 1 μm, and the length is in the range of 1–3 μm. Figure 1(c) depicts that NH2-PDI/TiO2 particles formed an agglomerated spherical structure. As shown in Fig. 1d, the average diameter of MoS2 nanoflowers is 0.5–1 μm. These nanoflowers have massive petals on the surfaces, which are free to gather closely. Figure 1(e) and (f) shows the morphology of the synthesized MoS2/TiO2 heterostructure characterized by SEM. It can be seen that MoS2 and TiO2 mixed together. In Fig. 1(h) and (i), a peculiar morphology of NH2-PDI/TiO2/MoS2 is shown. The NH2-PDI and MoS2 mixed with TiO2, and they can provide more active sites and mass charge transport pathways in the catalytic system. The color of MoS2/TiO2 and NH2-PDI/TiO2/MoS2 are shown in Fig. 1(g) and (j), respectively. MoS2/TiO2 was light grey while NH2-PDI/TiO2/MoS2 sample showed a red grey color.

Figure 1
figure 1

SEM images of TiO2 (a), NH2-PDI (b), NH2-PDI/TiO2 (c), MoS2 (d), MoS2/TiO2 (e,f), NH2-PDI/TiO2/MoS2 (h,i), and the appearance of MoS2/TiO2 (g), NH2-PDI/TiO2/MoS2 (j).

The composition of the NH2-PDI/TiO2/MoS2 was further investigated by an EDX attached to SEM (Fig. 2). Figure 2(b)–(h) shows the element mappings in a selected area of the composite (Fig. 2a). The homogeneous distributions of the elements Mo, S, C, N, Ti, and O can be clearly seen from the graphs. As shown in Fig. 2(i) and Fig. S-2, EDX analysis reveals that the composite contains Ti and O (in the case of TiO2), Mo and S (in the case of MoS2), or C and N (in the case of NH2-PDI). The atomic ratio of Ti, Mo and C equals to 25.2: 1: 1.3, meaning that the molar ratio of TiO2, MoS2, and NH2-PDI is about 95.3: 1.89: 0.68. The high proportion of C is due to the elemental carbon and carbonate species adsorbed on the TiO2 surface.

Figure 2
figure 2

(a) SEM image of NH2-PDI/TiO2/MoS2 photocatalyst; Element mappings of Mo, S, C, N, Ti, and O elements (b), C (c), N (d), O (e), Mo (f), S (g), and Ti (h) in the NH2-PDI/TiO2/MoS2 photocatalyst; (i) EDX spectrum of NH2-PDI/TiO2/MoS2 photocatalyst.

XRD measurements were performed to determine the crystalline structures of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 composites (Fig. 3). The crystal phase of TiO2 was consistent with anatase phase TiO2 (JCPDS NO. 71–1167) (Fig. 3a). Diffraction peaks appeared at 2θ = 25.3°, 37.7°, 48.0°, 53.8°, 55.0°, 62.6°, 58.6°, 70.6°, and 75.2°, corresponding to (101), (311), (200), (105), (211), (204), (116), (220), and (241) planes of TiO2, respectively31. The NH2-PDI appeared in the corresponding d-spacing as 10.35 Å, 9.22 Å, 5.24 Å, and 3.43 Å (Fig. 3b). The peak at 2θ = 25.74° (d spacing 3.43 Å) can be attributed to the π-π stacking of the adjacent NH2-PDI, since the distance of π-π stacking between the perylene nuclei was about 3.5 Å32. It was well known that PDIs exhibit a typical sharp diffraction peak at about 8–9°, which was centered at 9.57° (d spacing 9.22 Å)33. Additionally, X-ray powder diffraction measurements on NH2-PDI showed that the first diffraction peak appeared at 2θ = 8.53° (d spacing 10.35 Å) and the second-order diffraction peak appeared at 2θ = 16.85° (d spacing 5.24 Å). The first diffraction peak was assigned as (001) and the second peak was assigned as (002) α-form crystal diffraction, which has been reported by Miyata34. The multiple orders of reflection indicated that the self-assembled structures of NH2-PDI are well-ordered and layered microstructures. For the pure 3D flower-like spherical MoS2 nanostructure, the peaks correspond to (002), (100), (103), and (110) diffraction planes of 2H-MoS2 (JCPDS: 37-1492), which is consistent with previous studies (Fig. 3d)35. For MoS2/TiO2, the addition of MoS2 did not change the diffraction peak positions of TiO2 obviously, indicating that MoS2 was not incorporated into the TiO2 lattice. Obviously, both NH2-PDI/TiO2 and NH2-PDI/TiO2/MoS2 formed ordered structures. Comparison of the XRD patterns in Fig. 3a,c,e,f confirms that the XRD patterns of the NH2-PDI/TiO2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 nanoparticles have good consistency with the XRD data of anatase phase TiO2 (JCPDS NO. 71–1167) at 2θ degrees of 25.3°, 37.7°, 48.0°, 53.8°, 55.0°, 62.6°, 58.6°, 70.6°, and 75.2°36. Compare with pure TiO2, the crystallization properties of the two composites were slightly weakened, which should be due to the binding of NH2-PDI (Fig. 3c,f). Since the NH2-PDI/TiO2/MoS2 composite was loaded with very little NH2-PDI and MoS2, the characteristic peaks of NH2-PDI and MoS2 did not appear in its XRD patterns (Fig. 3f). Additionally, compared with pure TiO2, NH2-PDI/TiO2/MoS2 exhibited a widened peak width. This may be caused by the formation of heterojunction between TiO2, NH2-PDI and MoS2, and the heterojunction lead to lattice distortion of TiO2. The average particle sizes of TiO2, NH2-PDI/TiO2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 calculated by Scherrer’s equation were 10.2 nm, 12.5 nm, 11.3 nm, and 13.6 nm, respectively37.

Figure 3
figure 3

XRD patterns of TiO2 (a), NH2-PDI (b), NH2-PDI/TiO2 (c), MoS2 (d), MoS2/TiO2 (e), NH2-PDI/TiO2/MoS2 (f).

Raman spectroscopy was applied to further check the phase and formation of the NH2-PDI/TiO2/MoS2 nanocomposite. As shown in Fig. S-3, the pristine TiO2 nanoparticles exhibit five peaks at 145.1 ± 0.2, 196.6 ± 0.1, 396.1 ± 0.2, 513.7 ± 0.2, and 638.1 ± 0.1 cm−1, which belong to the Eg, Eg, B1g, A1g, and Eg anatase tetragonal vibration modes of TiO2, respectively38. The Eg, B1g, and A1g peaks correspond to symmetric stretching, symmetric bending, and antisymmetric bending vibrations of O-Ti–O, respectively. The results demonstrate the existence of typical anatase TiO2 phase, and it is consistent with the XRD results. Characteristic peaks of MoS2 (A1g and E12g modes) were observed in the Raman spectra of NH2-PDI/TiO2/MoS2 photocatalyst, indicating that MoS2 was loaded on the surface of TiO239. Compared with pure TiO2, NH2-PDI/TiO2/MoS2 photocatalyst exhibited red shifts of Eg, Eg, A1g, and Eg modes (145.6 ± 0.3, 197.8 ± 0.1, 514.1 ± 0.1, and 638.4 ± 0.1 cm−1). This indicates that there are strong intimate interactions between TiO2, MoS2, and NH2-PDI.

FT-IR measurements revealed the connection between TiO2, NH2-PDI and MoS2 in NH2-PDI/TiO2/MoS2 nanocomposite (Fig. S-4). As shown in Fig. S-4a, for pure TiO2, the vibration modes observed at 617 cm−1 and 749 cm−1 are due to the O–Ti–O bending and Ti–O stretching vibrations, respectively. For MoS2, the vibration mode at 1389 cm−1 is assigned to Mo–S vibration, while the vibration mode at 1550 cm−1 may be S–O asymmetric stretching (Fig. S-4d)40. In MoS2/TiO2 and NH2-PDI/TiO2/MoS2, all characteristic vibration modes were present but with very low Mo–S and S–O vibration modes, which indicates that the MoS2/TiO2 and NH2-PDI/TiO2/MoS2 contain very little amount of MoS2 (Fig. S-4e, f). Remarkably, the bending vibration of C-H at 2926–2817 cm−1 observed in NH2-PDI shifted to 2968–2875 cm−1 in NH2-PDI/TiO2 and to 2992–2889 cm−1 in NH2-PDI/TiO2/MoS2, which may be an indicative of the binding between NH2-PDI and TiO2 or MoS2. In addition, the stretching vibration of C = C in NH2-PDI at 1645 cm−1 shifted to 1639 cm−1 in NH2-PDI/TiO2 and to 1627 cm−1 in NH2-PDI/TiO2/MoS2 nanocomposite. Compared with NH2-PDI, the C=C absorption peak of NH2-PDI/TiO2/MoS2 was blue shifted, indicating that the π-π interaction within NH2-PDI decreased. The absorption peak around 3300 cm−1 was caused by the stretching vibration of O–H bond, which was related to the atmospheric water adsorbed on the catalyst surface. As shown in Fig. 4, the Ti…O bond and S…H bond can be formed when NH2-PDI meets with TiO2 and MoS2. After the addition of deionized water, NH2-PDI self-assembled through intermolecular π–π stacking interactions. For NH2-PDI, dichloromethane is a “good” solvent, while water is a “poor” solvent. When the “poor” solvent was added, the solubility of originally dissolved NH2-PDI was limited, and NH2-PDI self-assembled through non-covalent interactions and precipitated into solid.

Figure 4
figure 4

Graphical representation of the investigated adsorption mode of NH2-PDI onto the TiO2 and MoS2 surface.

The chemical composition and valence state of NH2-PDI/TiO2/MoS2 sample were studied by XPS (Fig. 5). According to Fig. 5a, NH2-PDI/TiO2/MoS2 composite contains Ti, O, C, Mo, and S elements. For NH2-PDI/TiO2/MoS2 nanocomposite, the positions of Ti2p1/2 and Ti2p3/2 were observed at 464.9 eV and 459.1 eV, respectively. While for pure TiO2 nanoparticles, Ti2p1/2 and Ti2p3/2 peaks were observed at 464.3 eV and 458.5 eV, respectively (Fig. 5b)41. The O1s peak positions of NH2-PDI/TiO2/MoS2 and pure TiO2 were observed at 530.2 eV and 529.8 eV, respectively (Fig. 5c). These results indicate that after adding NH2-PDI and MoS2 to TiO2 nanoparticles, Ti2p peaks move to a higher energy by 0.6 eV than pure TiO2. Also, the peak position of O1s moved 0.4 eV towards high energy. The shifted spectrum implies the presence of more defects or adsorbed hydroxyl groups on the surface of NH2-PDI/TiO2/MoS242. These defect states may serve as shallow donors to enhance charge transfer at the multiple interfaces and thus improve the overall degradation efficiency towards dyes. In NH2-PDI/TiO2/MoS2 sample, The C1s spectra (Fig. 5d) shows the corresponding peak at 284.8 eV, which can be identified as C–C/C=C/C–H functional groups. The positions of the Mo3d5/2 and Mo3d3/2 peaks were observed at 232.4 eV and 228.5 eV, respectively. Similarly, the position of S2p was observed at 161.9 eV43.

Figure 5
figure 5

XPS spectra of (a) survey spectra in NH2-PDI/TiO2/MoS2, (b) Ti2p in pure TiO2 and NH2-PDI/TiO2/MoS2, (c) O1s in pure TiO2 and NH2-PDI/TiO2/MoS2, (d) C1s in NH2-PDI/TiO2/MoS2, (e) Mo3d in NH2-PDI/TiO2/MoS2, and (f) S2p in NH2-PDI/TiO2/MoS2.

The surface areas of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 were determined by nitrogen adsorption method. The BET surface area of NH2-PDI/TiO2/MoS2 (39 m2/g) is higher than that of TiO2 (21 m2/g), NH2-PDI/TiO2 (37 m2/g), MoS2 (23 m2/g), or MoS2/TiO2 (26 m2/g). Therefore, it can be speculated that the large surface area of NH2-PDI/TiO2/MoS2 could promote the photocatalytic activity.

The UV–vis diffuse reflection spectra (DRS) and the corresponding Tauc plots of samples are shown in Fig. 6. Pure TiO2 absorbs only UV light and exhibits absorption below 380 nm. Adding NH2-PDI and MoS2 nanoflakes to TiO2 nanoparticles can result in more light-harvesting in visible region (Fig. 6a). The absorption intensity of NH2-PDI/TiO2/MoS2 in visible light region was obviously enhanced, which could be attributed to the visible light absorption characteristics of NH2-PDI and MoS2 nanoflakes. The band gap energy of the samples can be estimated by the formula Eg (eV) = 1240/λAE (nm) using the position of the absorption edge (λAE)44. The downward slopes of the absorption curves are extrapolated to cross the X-axis and the λAE-values of TiO2, MoS2/TiO2, NH2-PDI/TiO2, and NH2-PDI/TiO2/MoS2 samples were estimated to be at 389 nm, 423 nm, 459 nm, and 467 nm, respectively (Fig. S-5). The Eg values of TiO2, MoS2/TiO2, NH2-PDI/TiO2, and NH2-PDI/TiO2/MoS2 samples were estimated to be at 3.18, 2.93, 2.70, and 2.66 eV, respectively.

Figure 6
figure 6

(a) UV–vis diffuse reflection spectra (DRS) and (b) The corresponding Tauc plot of samples.

The band gaps (Eg, eV) of the samples are also calculated by the Tauc equation and Kubelka–Munk function45:

$${[F(R)h\upsilon]^0.5} =A(h\upsilon-Eg)$$
(1)
$$F(R)={(1-R)^{2}}/2R$$
(2)

where, R is the calibrated reflection of samples with BaSO4 reflection, and F(R) is proportional to the absorption constant. h and ν represent the Planck constant and frequency, while A and Eg represent constant and band gap energy, respectively. The [F(R)hν]0.5 versus hν is shown in Fig. 6b. Extrapolation of the linear portion at [F(R)hν]0.5 = 0 provides Eg value of the samples. The band gaps were found to be 2.98, 2.75, 2.47, and 2.27 eV for TiO2, MoS2/TiO2, NH2-PDI/TiO2, and NH2-PDI/TiO2/MoS2, respectively. The increase of UV–visible light absorption and decrease of band gap energy enhanced the photodegradation efficiency of NH2-PDI/TiO2/MoS2 towards MB.

The activity diagram of MB degradation by different catalysts in visible light and the first order kinetics curve fitting of MB degradation by different catalysts are shown in Fig. 7. Prior to the illumination, each catalyst was dispersed in the dye solution. After 120 min of vigorous magnetic stirring under dark condition, 51.3 ± 1.5% of MB was absorbed by NH2-PDI/TiO2/MoS2, whereas pure TiO2 only absorbed 8.2 ± 0.7% of MB. It can be seen that the spongy NH2-PDI clusters and multi-layer MoS2 nanoflowers decorated TiO2 exhibited extremely high adsorption efficiency under dark condition. Both MoS2 and TiO2/MoS2 have strong adsorption capacity, which may be due to the flower-like MoS2 with negative surface charge promoted the adsorption of cationic dye MB. It can be noted that after 120 min of irradiation, the degradation rate of dye was 99.0 ± 1.0% for NH2-PDI/TiO2/MoS2, while the corresponding data was 4.0 ± 0.3%, 31.3 ± 1.5%, 71.7 ± 1.6%, 78.7 ± 1.3%, and 87.6 ± 1.5% for NH2-PDI, TiO2, NH2-PDI/TiO2, MoS2, and MoS2/TiO2, respectively (Fig. 7A). Though TiO2 has no visible light absorption, it has been found to be solar active. This may be due to the dye sensitization mechanism of MB to TiO2. For better comparison, NH2-PDI/TiO2/1%MoS2 or NH2-PDI/TiO2/3%MoS2 photocatalysts was used for dye degradation under identical conditions. As a result, 97.3 ± 0.6% or 94.0 ± 1.1% of dye were degraded respectively under visible light irradiation in 120 min. The results show that 0.1%NH2-PDI/TiO2/2%MoS2 had higher efficiency in MB degradation than other catalysts. The kinetics plot shows that ln (C0/C) has linear relationships with reaction time, indicating that the photodegradation of MB follows first-order kinetics (Fig. 7B). The apparent rate constants (k) were determined as 0.00616, 0.00352, 0.00738, 0.00517, 0.00752, and 0.00806 min−1 for TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2, respectively. These results confirmed the high photocatalytic activity of NH2-PDI/TiO2/MoS2. Fig. S-6 shows the decolorization of MB dye (0.01 g/L) within 90 min in the presence of NH2-PDI/TiO2/MoS2 catalyst. The decrease of absorption spectrum indicates decolorization of the dye under the applied conditions. There were no additional peaks in the UV–Vis spectrum, indicating that the dye was completely degraded. It is reported that the products of the decolorization process are H2O, CO2, NO2, and SO246, and the mechanism of MB degradation is shown in Fig. S-7.

Figure 7
figure 7

(A) The activity diagram of MB degradation by different catalysts in visible light, (B) The first order kinetics curve fitting of MB degradation by different catalysts. (a) TiO2, (b) NH2-PDI, (c) NH2-PDI/TiO2, (d) MoS2, (e) MoS2/TiO2, (f) NH2-PDI/TiO2/2%MoS2, (g) NH2-PDI/TiO2/1%MoS2, and (h) NH2-PDI/TiO2/3%MoS2. [MB] = 0.01 g/L, pH = 7, catalyst suspended = 1 g/L.

Chemical oxygen demand (COD) is an indicator that reflecting the required oxygen for the oxidation of organic matters present in solution47. A smaller COD value means a lower pollution of the sample. The mineralization degree of MB was calculated by COD result, and the kinetic results shown that the k value of NH2-PDI/TiO2/MoS2 was 0.0076 min−1 (Fig. S-8). The value correspond to t1/2 values of 85.9 and 91.2 min (based on t1/2 = 0.693/k) for the degradation and mineralization of MB molecules, respectively48. These confirmed that the photodegradation of MB was 1.06 times faster than its mineralization. The results confirmed a relatively smaller mineralization extent for MB molecules than its photodegradation extent.

The photonic efficiency of MB dye degradation by NH2-PDI/TiO2/MoS2 under optimum condition was calculated using a reported method49. The quantum yield of reaction (Ø) is defined as follows (Eq. (1)):

$$\O =\frac{\mathrm{Number \,of\, molecules\, decomposed }}{\mathrm{Number\, of\, photons\, of\, light\, absorbed}}$$

The photo-degradation rate constants (k’) of MB dyes under monochromatic light source can also be used to calculate the reaction quantum yield (Eq. (2)).

$$\O =\frac{k^{\prime}}{303I\varepsilon l}$$

where Ø is the reaction of quantum yield. I is the light intensity range in 400–800 nm (2.312 × 10–3 Einstein). ɛ is the molar absorptivity of MB at 664 nm (25,440 cm–1 M–1). l is the path length of reaction tube, which is 0.23 m for 50 mL solution. The degradation quantum yields obtained with TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 were 0.000152, 0.000869, 0.000182, 0.000128 and 0.000186, respectively. These results indicate that the quantum yield of NH2-PDI/TiO2/MoS2 catalyzes process was higher than that of other processes.

During the process of MB degradation, a large amount of MB was adsorbed on the surface of catalysts. The visible light driven catalysis of the NH2-PDI/TiO2/MoS2 may be enhanced due to the effect of dye sensitization. Therefore, amoxicillin (AMX, a colorless medical antibiotic) was selected to evaluate the photocatalytic activity of NH2-PDI/TiO2/MoS2 nanoparticles under the same conditions. The photocatalytic degradation results of AMX are shown in Fig. S-9. NH2-PDI/TiO2/MoS2 exhibited higher visible photocatalytic activity than NH2-PDI/TiO2 or MoS2/TiO2. When AMX solution was mixed with NH2-PDI/TiO2/MoS2 in the absence of light, the removal rate of AMX was 48%. When NH2-PDI/TiO2 or MoS2/TiO2 was used under visible light irradiation for 120 min, the degradation rates of AMX were 56.9% or 71.2%, respectively. Comparatively, NH2-PDI/TiO2/MoS2 exhibited superior photodegradation performance (84.8%). The degradation processes conforms to pseudo-first-order kinetics. The photocatalytic activity of NH2-PDI/TiO2/MoS2 reached a maximum value of 0.00986 min−1, which was about 1.67 times and 1.79 times higher than NH2-PDI/TiO2 and MoS2/TiO2 (0.00590 min−1 and 0.00551 min−1), respectively. The result indicates that NH2-PDI/TiO2/MoS2 has strong visible-light driven photocatalytic activity.

The reusability and stability of a composite photocatalyst are important factor affecting its application (Fig. S-10). The NH2-PDI/TiO2/MoS2 was used for the degradation of different solutions of MB for six consecutive experiments under the same reaction conditions. After each experiment, the catalyst was recovered by centrifugation and dried in an air oven. The degradation efficiency of NH2-PDI/TiO2/MoS2 reduced from 99 ± 0.9% to 89 ± 2.3% after six cycles of reuse, which was probably due to a small loss of catalyst during recycle (Fig. S-10a). NH2-PDI/TiO2/MoS2 was prepared by hydrothermal synthesis method, which made the contact between three monomers more compacted and reduced the dissolution of MoS2 or NH2-PDI molecule. Therefore, the composite catalyst exhibits high stability in the degradation process. In addition, the composite material packaged TiO2 by NH2-PDI and MoS2. The contact area between the three semiconductors increases due to the point-to-surface contact. Due to the close contact interface, the interactions between the three semiconductors would also be more intense. The π-conjugated electrons in NH2-PDI can be quickly transferred from the interior of the semiconductor to the surface, and then to the surface of TiO2 or MoS2. This reduces electrons and holes recombination and improves the photocatalytic activity. Fig. S-10b shows the FT-IR spectra of the NH2-PDI/TiO2/MoS2 photocatalyst before and after six cycles of photocatalytic degradation of MB. It can be seen that the spectrum of the regenerated photocatalyst was basically the same as that of the fresh photocatalyst. There is no spectrum of MB, which proves that MB has been completely degraded rather than adsorbed on the catalyst surface during the illumination process. The result of UV–vis spectra shows that the concentration of organic matter in the washing solution of catalyst was very low. Few intermediate species and NH2-PDI were detected, indicating that the dissolution of NH2-PDI in the composite structure was very few and MB was mineralized to CO2. These results show that the NH2-PDI/TiO2/MoS2 catalyst is stable and reusable.

The photoluminescence (PL) spectrum was used to analyze the recombination of electron hole pairs. In general, a higher peak indicates a more rapid charge recombination. As shown in Fig. S-11, the peak at 399 nm is attributed to a direct transition from the conduction band to the valence band. Moreover, the luminescence peaks at 440, 451, and 469 nm are caused by inter-band transitions, and the peaks at 483 and 492 nm are caused by intra-band transitions within the energy level traps or surface states50. The peak intensity was arranged as follow: TiO2 > NH2-PDI/TiO2 > MoS2/TiO2 > NH2-PDI/TiO2/MoS2. Especially, the peak intensity of NH2-PDI/TiO2/MoS2 composite was much weaker than that of the pure TiO2. The –C=O…H hydrogen bond between the -OH of TiO2 and –C=O of NH2-PDI, and the S…H bond between the -S of MoS2 and -NH2 of NH2-PDI acted as short and fast channels for migrating the photogenerated charge carriers from NH2-PDI to TiO2 and MoS2. Based on the above results, the introduction of NH2-PDI and MoS2 can greatly accelerate the charge transfer process.

The mechanism of enhanced photocatalytic activity of NH2-PDI/TiO2/MoS2 composite was further studied by photoelectrochemistry method. The separation of photogenerated electrons and holes plays an important role in the photocatalytic decomposition of organic pollutants, which can be evaluated by transient photocurrent responses (I-t) and electrochemical impedance spectra (EIS) (Fig. 8). Figure 8a shows the I-t curves of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 under several intermittent visible light irradiation cycles. It can be seen that the photocurrent of all samples exhibited good repeatability when the light was turned on and off. The light current response of pure TiO2 and NH2-PDI were very weak (only 0.119 and 0.035 μA cm−2, respectively). NH2-PDI/TiO2/MoS2 composite had the largest photocurrent response, and the stable photocurrent was about 1.007 μA cm-2, which was higher than that of MoS2 (0.372 μA cm−2), MoS2/TiO2 (0.522 μA cm−2), and NH2-PDI/TiO2 (0.846 μA cm−2). The enhanced photocurrent response indicates that the separation efficiency of photogenerated carriers and the photocatalytic performance were improved. Figure 8b displays the EIS Nyquist plots of as-prepared samples. The radius of the arc on the EIS spectrum reflects the surface charge transfer resistance and the solid interface delamination resistance. The smaller the semicircle arc is, the easier the charge transfer proceeds51. Well discussion on EIS spectra and the resulted Bodes' plots has been presented in previous work52. Here, the following trend was obtained for the arc radius of the Nyquist plots, confirming that the NH2-PDI/TiO2/MoS2 composite have better charge transfer capability.

Figure 8
figure 8

(a) Photocurrent transient responses and (b) EIS spectra of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 at a constant potential of 0.4 V with simulated solar light at 300 mW/cm2. frequency: 0–10,000 Hz, electrolyte: 0.1 M Na2SO4, amplitude: 10 mV.

NH2-PDI > TiO2 > MoS2 > NH2-PDI/TiO2 > MoS2/TiO2 > NH2-PDI/TiO2/MoS2.

The Mott-Schottky plots were used to estimate the conduction band (CB) of the TiO2, NH2-PDI, and MoS2, which were − 1.26 eV, − 1.61 eV, and − 0.75 eV, respectively (Fig. S-12). Combining with their bandgaps (TiO2 at 2.98 eV [Fig. 5b], NH2-PDI at 2.36 eV30, and MoS2 at 1.75 eV26), the CB/VB (conduction band/valence band) of TiO2, NH2-PDI, and MoS2 were evaluated to be − 1.26/1.72 eV, − 1.56/0.8 eV, and − 0.75/1.00 eV, respectively. It can be found that the CB position of singlet oxygen (·O2-) (− 0.18 eV) is shallower than that of NH2-PDI, TiO2, and MoS2. This result indicates that ·O2- may be the main active substance in NH2-PDI/TiO2/MoS2 composite for photocatalytic degradation of MB. The VB positions of MoS2 was shallower than that of hydroxyl radical (·OH) (2.40 eV)53, so it is possible that the ·OH cannot be generated by NH2-PDI/TiO2/MoS2.

The active species in the photocatalytic process was studied by free radical capture experiment53. As can be seen from Fig. S-13, with the addition of p-BQ scavenger, the degradation rate decreased from 99.9% to 33%, indicating that ·O2 is one of the main active species in the photocatalytic degradation process. In the presence of EDTA-2Na, the degradation efficiency was significantly reduced, indicating that h+ is also one of the major active radicals. After the addition of IPA, the degradation activity of the catalyst did not decrease, indicating that ·OH is not the main active species in MB degradation.

Based on the results above, two mechanisms including ‘type II- heterojunction’ and ‘Z-scheme’ were suggested to illustrate the photodegradation pathways occurred in this work (Fig. 9)54. In ‘type II- heterojunction’ mechanism (Fig. 9a), both the NH2-PDI and MoS2 semiconductors can be excited by the arrived photons under visible light irradiation, resulting in production of e-/h+ pairs in both semiconductors. As the CB and VB of NH2-PDI are located at higher energy levels than those of TiO2 and MoS2, the photogenerated electrons in the CB of NH2-PDI can transfer to the CB of TiO2, and then to the CB of MoS2, while holes travel in the opposite direction in the VB. These charge carriers’ transfer processes result in the accumulation of the photogenerated electron in the CB level of MoS2 and the photogenerated holes in the VB level of NH2-PDI. While in the ‘Z-scheme’ mechanism (Fig. 9b), the photogenerated electrons at the CB level of MoS2 can migrate to the VB level of NH2-PDI based on the standard potentials, and thus, the ‘Z-scheme’ pathway accumulates the photogenerated electrons at the CB level of NH2-PDI and the photogenerated holes at the VB level of MoS2. The accumulated electrons at the CB level of NH2-PDI have the potential of − 1.61 eV, which is more negative than the potential of MoS2 (-0.75 eV). Therefore, the accumulated electrons at the CB level of NH2-PDI can effectively reduce dissolved oxygen and generate enough ·O2 as an efficient reactive species for the degradation of MB. These holes are also stronger oxidizing agents than the accumulated holes in the VB level of NH2-PDI (accumulated in type II mechanism) that can directly oxidize MB molecules. Therefore, the ‘Z-scheme’ mechanism is a better illustration for the enhanced photocatalytic activity of NH2-PDI/TiO2/MoS2 composite in MB degradation than the ‘type II-heterojunction’ mechanism. Nevertheless, in both mechanisms, TiO2 facilitates the electron transfer between the conduction bands, reduces photoinduced charge carrier recombination and increases photocatalytic activity.

Figure 9
figure 9

Schematic illustration of the photocatalytic mechanism over the NH2-PDI/TiO2/MoS2 heterostructure. (a) Type II-Heterojunction and (b) Z-scheme pathways.

Conclusions

In conclusion, NH2-PDI/TiO2/MoS2 ternary composite was successfully synthesized by hydrothermal synthesis method. Morphological and chemical features of TiO2, NH2-PDI, NH2-PDI/TiO2, MoS2, MoS2/TiO2, and NH2-PDI/TiO2/MoS2 composites were examined by SEM, XRD, FTIR, XPS, and DRS. SEM results demonstrate that NH2-PDI nanorods and MoS2 nanoflowers mixed with TiO2, which can provide more active sites and mass charge transport pathways in the catalytic system. The photoelectrochemical measurement results show that the photocurrent performance of the ternary composite catalyst was superior to that of binary composites or pure NH2-PDI, MoS2, or TiO2. The NH2-PDI/TiO2/MoS2 exhibits excellent photocatalytic activity and high stability. The ‘Z-scheme’ accumulates more electrons at the CB level of NH2-PDI and photogenerated holes at the VB level of MoS2. The accumulated electrons can effectively reduce dissolved oxygen and generate enough ·O2 as an efficient reactive species for the degradation of MB. These holes are also strong oxidizing agents that can directly oxidize MB molecules. TiO2 could reduce photoinduced charge carrier recombination and increases photocatalytic activity. This study highlights the potential application of organic and inorganic photocatalysts, and provides a feasible strategy for the photodegradation of dyes.