Introduction

Methylene blue (MB), a water-soluble polycyclic aromatic hydrocarbon, is commonly used in industrial, pharmaceutical, aquaculture, and skin care areas (Liu et al. 2012). However, causing serious pollution to water due to increased production and wide application, the polluted water with MB has the characteristics of high concentration, high chroma, high pH, and hard to degradation (Saravanan et al. 2013). Researchers have tried many methods to solve the above problem, such as adsorption, photocatalytic technology, and membrane separation technology; photocatalytic technology has been widely concerned owing to its low energy consumption, mild conditions, simple operation, and no secondary pollution.

Over the past few decades, Bi2WO6, a typical narrow band gap semiconductor (Eg ≈ 2.7 eV), has been widely used for photodegradation of organic contaminants due to its unique structure and stable chemical properties (And and Zhu 2010). However, the photoactivity of pure Bi2WO6 is still not satisfactory attributing to the rapid recombination of photogenerated electron-hole pairs, so designing a new photocatalyst to improve its photocatalytic efficiency is imperative. Heterostructure photocatalyst has been proved to be a good choice to enhance the photocatalytic performance in comparation to single-component photocatalyst, with successful cases like Fe3O4/Bi2WO6 (Zhou et al. 2015), Cd/Bi2WO6 (Xu et al. 2015), AgBr/Ag/Bi2WO6 (Zhang et al. 2009), g-C3N4/Bi2WO6 (Lei et al. 2011), Bi2S3/Bi2WO6 (Sangeeta and Do-Heyoung 2018), and CdS/Bi2WO6 (Lei GE, Liu J 2011). These composite materials display an enhanced photocatalytic performance owing to the fast transfer photogenerated electrons-hole pair.

Bi25FeO40, a typical soft bismuth ore material, with a relatively narrow band gap (2.1–2.7 eV) (Ji et al. 2017), has been regarded as a promising photocatalyst attributing to its special structure, a crystal structure with a large number of oxygen vacancies and ion vacancies, which can greatly improve the photocatalytic performance of material (Köferstein et al. 2014). Unfortunately, the photocatalytic performance of pure Bi25FeO40 can only be enhanced by the addition of an electron scavenger such as H2O2, which can surpass the recombination of photogenerated electrons-hole pairs (Wisedsri et al. 2011; Hou et al. 2010). Until now, researchers have been tried many ways to improve its photocatalytic performance, such as Bi25FeO40/RGO (Basith et al. 2018), Bi25FeO40/Bi2Fe4O9 (Wang et al. 2019), and Bi25FeO40/Fe3O4/Fe2O3 (de Góis et al. 2019). In light of this, the emergence of an all-solid-state Z-scheme heterojunction could address the problem to some extent. In this Z-scheme system, the photogenerated electrons in the CB of PSII could recombine with the holes in the VB of PSI; thus, holes with strong oxidizing ability are left in the VB of PSII and photogenerated electrons with strong reducing ability are left in the CB of PSI(Hu et al. 2017; Zhou et al. 2014). Also, conductive medium plays an important role in Z-scheme system, like Au (Bai et al. 2016), Ag (Li et al. 2017), and graphene oxide (GO) (Jiang et al. 2017; Miao et al. 2017; Rakibuddin et al. 2017) or reduced graphene oxide (r-GO) (Li et al. 2014; Jo and Selvam 2017; Chen et al. 2017); among them, r-GO stands out due to its large specific surface area and strong electrical conductivity. Ma et al. (2016) fabricated a Z-scheme catalyst named g-C3N4/r-GO/Bi2WO6; the excellent photocatalytic performance had been verified by degrading TCP. Zhang et al. (2016a, b) constructed a Bi2WO6/MoS2/RGO Z-scheme catalyst to remove Cr6+; the r-GO could effectively improve the photocatalytic performance of Bi2WO6. Wu et al. (2017) synthesized a g-C3N4-r-GO-TiO2 Z-scheme system and further revealed the multi-functional roles of r-GO in enhancing the photodegradation rate.

In summary, a novel Z-scheme photocatalyst Bi2WO6/r-GO/Bi25FeO40 was fabricated, and Bi2WO6 (ECB = + 0.6 V, EVB = + 3.26 V vs. NHE) and Bi25FeO40 (ECB = − 1.3 V, EVB = + 1.13 V vs. NHE) were matched well. Firstly, nanosheet Bi2WO6 was synthesized in the presence of CTAB, then Bi2WO6 was used as a precursor template, and Fe(NO3)3·9H2O and r-GO were added to obtain a ternary material under hydrothermal condition of 180 °C. The photocatalytic performance was evaluated by degrading RhB solution under visible-light irradiation. Ternary material exhibited better photocatalytic performance than binary and single materials. At the same time, a possible photo-degradation mechanism of Bi2WO6/r-GO/Bi25FeO40 has been discussed in detail on the basis of UV–vis and PL results. Such distinctive Z-scheme photocatalyst has greatly overcome the shortcomings of Bi2WO6 and successfully achieved the separation of photogenerated electron-hole pairs.

Experiment

Materials

All reagents used in this experiment were of analytical grade and were not further purified prior to use. All of them were purchased from Chengdu Kelong Co., Ltd.

Preparation of Bi2WO6 nanosheet

The nanosheet Bi2WO6 was constructed through the hydrothermal method. Typically, 1.2125 g of Bi(NO3)3·5H2O and 0.1 g of hexadecyl trimethyl ammonium bromide (CTAB) were firstly dispersed into 50 mL of deionized water under ultrasonic treatment until obtaining a clear solution, subsequently adding 0.4125 g of Na2WO4·2H2O to above solution and keeping vigorous magnetic stirring at room temperature for 30 min. Transferring the mixed solution to a 100-mL Teflon-lined autoclave and keeping at 160 °C for 12 h. The precipitate was washed with ethanol and deionized water for several times after completion of the hydrothermal reaction. Bi2WO6 nanosheet was obtained after drying at 60 °C for 12 h.

Preparation of Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40

At first, graphene oxide (GO) was fabricated through modified Hummer’s method (Shi et al. 2016). Then, 0.5 g of Bi2WO6 was dispersed in deionized water, followed by a certain amount of Fe(NO3)3·9H2O was added to the above solution, adjusting the pH of the solution to 10 using NaOH solution (10 M). In addition, 10 mg of GO and 1 mg of L(+)-ascorbic acid (L-AA) were dissolved in 10 mL of deionized water and ultra-sonicated for 30 min to obtain even solution, mixing the above two solutions and stirring for 1 h, then transferring the mixed solution to a 100-mL Teflon-lined autoclave, and keeping at 180 °C for 12 h. The precipitate was washed with ethanol and deionized water for several times after the hydrothermal reaction was completed. Ternary material was obtained after drying at 100 °C for 10 h. For comparison, the pristine Bi25FeO40 and Bi2WO6/Bi25FeO40 were prepared under identical condition without adding Bi2WO6 or GO.

Characterization

The crystalline phases and purity of catalysts were determined by an X-ray diffractometer (XRD, PANalytical B.V., X Pert PRO MPD) with a Cu-Kα radiation (λ = 1.54060 Å, 40 kV, 40 mA) in the range of 5° ≤ 2θ ≤ 80°, operated at a scanning rate of 1°/min. Raman spectra (Oceanoptics, IDRaman micro IM-52) was used to analyze the molecular structure of materials. The morphology structures, element composition, and distribution of as-prepared samples were observed by scanning electron microscopy (SEM, Carl Zeiss AG, ZEISS EV0 MA15) equipped with an energy dispersive X-ray (EDX) spectroscopy. The nanostructures of catalysts were tested using a transmission electron microscope (TEM, Japan, Jeol 2100F). The UV–vis absorption spectra were recorded by a UV–vis spectrometer (Platinum Elmer Instrument Co., Ltd., PerkinEImerLambda850), choosing BaSO4 as background. The surface chemical composition and chemical states of solid surface were carried out using an X-ray photoelectron spectrometer (XPS, America, Thermo ESCALAB 250XI) with Al Kα source. Photoluminescence (PL) spectra of samples were implemented using a PerKinEImer LS55 spectrofluorometer with an excitation wavelength of 270 nm.

Electrochemical measurements

The electrochemical impedance spectra (EIS) and photocurrent response intensity curve were obtained through CHI604D (Shanghai) electrochemical workstation equipped with a standard three-electrode cell: A Pt counter electrode, a saturated calomel reference electrode, and a FTO glass working electrode in the experiment; the electrolyte is Na2SO4 aqueous solution (0.5 M) and a 200 W LED lamp is used as a light source. Specific steps are as follows: a 5 mg of sample was dissolved in a mixed solution consisting of 50 μL of 5% Nafion solution and 1 mL absolute ethanol, and then ultra-sonicated for 30 min to obtain a uniform solution. The suspension was coated on the FTO glass working electrode to make an effective exposed area of 1 cm2, measuring after ventilation overnight.

Photocatalytic experiment

The photocatalytic performance of all samples was evaluated by degrading MB aqueous solution under visible-light irradiation. There are same steps in each sample: 40 mg of the photocatalyst was dispersed into 50 mL of MB aqueous solution (20 mg/L). Before lighting, the mixed solution was stirred for 30 min in dark to ensure good dispersion and establish an adsorption–desorption equilibrium and then exposed to 200 W of visible light to begin photocatalytic experiment. The solution of aliquots volume was centrifuged to remove photocatalytic powders every 5 min, and then, the concentration of MB molecule at different time was measured by a UV–visible spectrophotometer (UV-1800, Shimadzu, Japan) at λ = 664 nm; deionized water was reference solution. Generally, calculating the degradation rate by the following formula: η =\( \frac{C_0-{C}_t}{C_0}\times 100\% \), where C0 is the initial concentration of MB and Ct is the concentration of MB at time t (min). In addition, cyclic stability of photocatalytic material is a critical performance too, which was recycled four times to ensure its reliability. The photocatalytic powders were washed with deionized water and ethanol for several times at the end of each experiment, and then, put it in fresh MB solution to start next recycle under identical condition.

Results and discussion

Crystal structure and morphology

The XRD patterns of GO, Bi2WO6, Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40 are shown in Fig. 1. The characteristic peak located at 2θ = 9.6° in the GO XRD pattern is assigned to the (001) plane of GO (Zhu et al. 2018). For pure Bi2WO6, the conspicuous peaks located at 2θ = 28°, 34°, 47.5°, 56°, 59°, 69°, 76°, and 79° are assigned to (113), (006), (220), (313), (226), (400), (139), and (420) respectively (Zhang et al. 2016a, b). The sharp peak patterns of Bi25FeO40 are in good assignment with the standard Joint Committee on Powder Diffraction Standards (JCPDS); the peaks at 24°, 27°, 30°, 33°, 39°, 52°, 66°, and 78° are related to (220), (310), (222), (321), (332), (530), (640), and (653) respectively (Zhang et al. 2015). However, the charactiristic peaks of Bi25 FeO40 have been not found in binary composite and ternary composite, only the charactiristic peaks of Bi6 WO6 appeared. Probably because the small dosage of Fe(NO3)3·9H2O (6% by Bi2WO6) is wrapped with Bi2WO6 to make it difficult to detect. Also, the characteristic peak of GO in the XRD pattern of Bi2WO6/r-GO/Bi25FeO40 cannot be appeared, which can be considered as the relatively low amount of GO and the appearance of RGO. No other impurity phases appear, proving the high purity of Bi2WO6/r-GO/Bi25FeO40.

Fig. 1
figure 1

XRD patterns of GO, Bi2WO6, Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40 composites

To discuss the morphology and detailed structure of as-prepared products, SEM images are displayed in Fig. 2a–c. The pure Bi2WO6 is constructed from a large number of regular square nanosheets. GO displays a flake-like structure with a wrinkled surface (Fig. 2b). For Bi2WO6/r-GO/Bi25FeO40 (Fig. 2c), the r-GO and Bi25FeO40 do not cover the square structure of Bi2WO6 due to the small dosage. The Bi25FeO40 particles load on the Bi2WO6 square nanosheet, and they are in close contact with r-GO, which can promote the effective separation of photogenerated electron-hole pairs and further improve the photocatalytic activity of ternary composite. Moreover, the EDS pattern suggests that there are five elements of Bi, W, Fe, O, and C in the ternary composite. In addition, TEM technology is operated to further characterize the morphology of Bi2WO6 and Bi2WO6/r-GO/Bi25FeO40. As shown in Fig. 3a, Bi2WO6 has a square structure. The ternary composite reveals an irregular morphology after the introduction of r-GO and Bi25FeO40, which is in agreement with SEM analysis that Bi25FeO40, Bi2WO6, and r-GO are in good contact with each other.

Fig. 2
figure 2

SEM images of a Bi2WO6, b GO, and c Bi2WO6/r-GO/Bi25FeO40, and the EDS pattern of d Bi2WO6/r-GO/Bi25FeO40

Fig. 3
figure 3

TEM images of a Bi2WO6 and b Bi2WO6/r-GO/Bi25FeO40

Surface structure and properties

Raman spectroscopy was employed to provide further evidence of the local structure of materials. As shown in Fig. 4, the Raman peaks at approximately 304, 712, and 791 cm−1 were appeared in pure Bi2WO6 which were in assignment with the bending vibration of WO6, antisymmetric bridging mode of tungstate chains, and stretching vibration (Ma et al. 2016). For single Bi25FeO40, the peaks mainly concentrate on 273 and 528 cm−1. Two distinct peaks at 1323 and 1528 cm−1 are related to D (the symmetry A1g k-point phonon) and G (the E2g phonon of sp2 carbon atoms) bands of GO (Chen et al. 2017). In general, the intensity ratio (ID/IG) of the D band and the G band can indicate the reduction degree of GO (Shen et al. 2017), the larger the ID/IG, the higher the reduction degree of GO. Similarly, the characteristic peaks of GO and Bi2WO6 in the Raman spectra are clearly visible; however, the peaks of Bi25FeO40 are indistinguishable, indicating that Bi25FeO40 is highly dispersed on the surface of ternary composite. Furthermore, the value of ID/IG (1.48) in Bi2WO6/r-GO/Bi25FeO40 is increased compared with GO (ID/IG = 1.30), indicating successful reduction of GO during the hydrothermal process.

Fig. 4
figure 4

Raman spectra of GO, Bi2WO6, Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40 composites

The chemical states of the ternary composite are studied by means of XPS technique. As the full survey spectra of Bi2WO6/r-GO/Bi25FeO40 are shown in Fig. 5a, the signals of Bi, W, Fe, O, and C elements are found in ternary composite, indicating the presence of Bi2WO6, Bi25FeO40, and r-GO. Figure 5 b exhibits the Bi 4f spectra of ternary composite in which the peaks at 159.4 and 164.7 eV are assigned to Bi 4f7/2 and Bi 4f5/2 respectively, suggesting that the elemental state of bismuth is Bi3+ (Wu et al. 2007) in the catalyst. For the XPS result of Fe 2p, the signal has been divided into Fe 2p1/2 and Fe 2p3/2 attributing to the spin–orbit coupling (Zhang et al. 2016a, b), the peak centers located at 709.8 and 723.4 eV respectively. Figure 5 d shows the XPS result of W; the two typical peaks at 35.78 and 37.9 eV are assigned to W 4f7/2 and W 4f5/2, which can be ascribed to the oxidation state of W6+ (Fu et al. 2005). As shown in Fig. 5e, the characteristic orbital of the O 1s is observed with peak locations at 531.9, 529.8, and 531.0 eV, which are corresponding to the r-GO (Zhu et al. 2018), Fe–O–Bi, and Bi–O–W (Zhang et al. 2016a, b). The binding energy peaks in Fig. 5f at 284.9, 286.8, 287.7, and 288.7 eV are well in accordance with C 1s state in r-GO, which are assigned to C–C, C–O, C=O, and O–C=O (Zhu et al. 2018) respectively. All binding energies are consistent with previous reports.

Fig. 5
figure 5

XPS spectra of Bi2WO6/r-GO/Bi25FeO40 composite, a survey, b Bi 4f, c Fe 2p, d W 4F, e O 1s, and f C 1s

The optical absorption properties of as-prepared samples are investigated via the UV–vis DRS technique. All samples exhibit visible-light absorption in the wavenumber range of 200 to 700 cm−1 in Fig. 6. Ternary composite establishes a wide absorption edge after the introduction of Bi25FeO40 and r-GO; the absorption edges of Bi2WO6, Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40 are 470, 520, 610, and 680 cm−1 respectively, indicating the ternary material can make full use of visible light and further improve the photocatalytic performance. Meanwhile, the band gap energy is calculated according to the following formula:

$$ {\left(\upalpha \mathrm{h}\upnu \right)}^{\mathrm{n}}=A\left(\mathrm{h}\upnu -{E}_{\mathrm{g}}\right) $$

where α, ν, A, and Eg are absorption coefficient, optical frequency, constant, and band gap respectively; the value of n depends on the type of semiconductor; there is 1/2 in direct band gap and 2 in indirect band gap (López and Gómez 2012). The value of n in Bi2WO6 and Bi25FeO40 are both 2 due to they are indirect gap semiconductor. As shown in Fig. 3, the band gap of Bi2WO6 and Bi25FeO40 is 2.66 and 2.42 eV respectively, which are similar to previous reports.

Fig. 6
figure 6

UV–vis DRS spectra of Bi2WO6, Bi25FeO40, and the Bi2WO6/Bi25FeO40 and Bi2WO6/r-GO/Bi25FeO40 composites, and the plot of (αhν)2 vs. hν (inset) for Bi2WO6 and Bi25FeO40

Photoelectrochemical performance

The electrochemical impedance spectroscopy (EIS) test is an effective method for analyzing the separation efficiency of photogenerated electron-hole pairs. Generally, the smaller the arc radius of the EIS diagram, the higher the photogenerated electron-hole pairs separation efficiency (Shi et al. 2016). As shown in Fig. 7a, the arc radius of all samples is in the order Bi25FeO40 > Bi2WO6 > Bi2WO6/Bi25FeO40 > Bi2WO6/r-GO/Bi25FeO40 composites. Obviously, ternary material with the smallest arc radius has the highest photogenerated electron-hole pairs’ separation efficiency, which is consistent with the PL response. The photocurrent response intensity curve of all samples was conducted to the same electrolyte (Na2SO4 solution, 0.5 M) under the condition of alternating light and dark. As seen in Fig. 7b, modified Bi2WO6/r-GO/Bi25FeO40 exhibits a higher photocurrent response intensity than single Bi2WO6; its value can reach 2.38 μA/cm2 which is about 6.4 times than that of pristine Bi2WO6. Such Z-scheme photocatalyst will show excellent potential in water treatment and other fields.

Fig. 7
figure 7

a EIS Nyquist plots and b photocurrent response spectra of Bi2WO6, Bi25FeO40, and the Bi2WO6/Bi25FeO40 and Bi2WO6/r-GO/Bi25FeO40 composites

Photocatalytic degradation of RhB

The photoluminescence (PL) spectra are investigated to evaluate the migration and recombination processes of photogenerated electrons-hole pairs of the photocatalyst. Fluorescence comes from the recombination of free carriers inside the semiconductor, the faster the recombination of photogenerated electron-hole pairs, the higher the fluorescence excitation intensity of the material. As shown in Fig. 8, the emission peaks of all samples focus on 450 nm; Bi2WO6 exhibits a strong PL intensity owing to the rapid recombination of photogenerated electron-hole pairs on its surface. The PL strength of Bi2WO6/Bi25FeO40 is lower than that of Bi2WO6 and Bi25FeO40, indicating the combination of the two can suppress the recombination of photogenerated electron-hole pairs. More importantly, Bi2WO6/r-GO/Bi25FeO40 displays a weaker PL intensity than Bi2WO6/Bi25FeO4, revealing r-GO as an electronic modifier in the Bi2WO6/r-GO/Bi25FeO40 enables higher separation of photogenerated electron-hole pairs, and then, more active substances generated in the photocatalytic process. The result indicates the successful construction of Z-scheme photocatalyst.

Fig. 8
figure 8

Photoluminescence spectra for pure Bi2WO6, Bi25FeO40, Bi2WO6/Bi25FeO40, and Bi2WO6/r-GO/Bi25FeO40 composites

To demonstrate the photocatalytic activity of all samples, the photodegradation of MB aqueous solution (20 mg/L) was operated under visible-light irradiation at room temperature. The dynamic degradation curve of all samples is shown in Fig. 9a. Bi2WO6/r-GO/Bi25FeO40 displays superior photocatalytic performance in 30-min irradiation compared with pure Bi2WO6. Furthermore, the blank experiment without photocatalyst proves that the property of MB is stable, thus the self-photolysis of it can be ignored. The degradation rate of Bi2WO6/r-GO/Bi25FeO40 is close to 98.1% that is about 2.3 times than that of pure Bi2WO6 (42%). Furthermore, the UV–vis absorption spectrum of Bi2WO6/r-GO/Bi25FeO40 is shown in Fig. 9b, which further explains the change in MB concentration during the irradiation time. In addition, the kinetic of the photocatalytic reaction is in line with first-order reaction that can be described as: ln (Ct/C0) = kt. Here, C0 is the initial concentration of MB, Ct is the concentration of MB at irradiation time t (min), and k (min−1) is apparent reaction rate constant. Obviously, Bi2WO6/r-GO/Bi25FeO40 shows a highest reaction rate constant (0.10155 min−1) which is about 9.2 and 11.4 times than that of Bi2WO6 and Bi25FeO40, respectively. Meanwhile, the stability of Bi2WO6 and Bi2WO6/r-GO/Bi25FeO40 was studied via recycling experiment; all conditions are the same as mentioned above. The catalyst was centrifuged and washed with deionized water at the end of each cycle. It is obvious that Bi2WO6/r-GO/Bi25FeO40 exhibits a better stability than Bi2WO6 after four cycles.

Fig. 9
figure 9

a Dynamic degradation curve of MB with different catalysts under visible-light irradiation. b Absorption changes of MB solution in the presence of 6% Bi2WO6/r-GO/Bi25FeO40 composite. c Reaction rate constants (k) of the as-prepared samples. d Cycling performances of photocatalytic degradation of MB over Bi2WO6 and Bi2WO6/r-GO/Bi25FeO40

Photocatalytic mechanism

In order to clearly elaborate the possible degradation mechanism of Bi2WO6/r-GO/Bi25FeO40, the Mott–Schottky plots and UV–vis spectra were carried out to estimate the CB and VB potentials of Bi2WO6 and Bi25FeO40. As shown in Fig. 10, the flat band potential (Ufb) values of Bi2WO6 and Bi25FeO40 are + 0.56 and − 1.34 V (vs. SCE). Then, the values of CB are 0.36 and − 1.54 V (vs. SCE, CB ≈ Ufb − 0.2 V), and they are 0.6 and − 1.3 V vs. normal hydrogen electrode (NHE, NHE = SCE + 0.24 V) (Xu et al. 2018). In addition, the VB potentials of Bi2WO6 and Bi25FeO40 could calculate through EVB = ECB + Eg (Lin et al. 2017), and the Eg values for Bi2WO6 and Bi25FeO40 are 2.66 and 2.43 V according to UV–vis results. Therefore, the VB potentials of Bi2WO6 and Bi25FeO40 are determined to be 3.26 and 1.13 V. The band structures of them are shown in Fig. 11.

Fig. 10
figure 10

Mott–Schottky plots of Bi2WO6 (a) and Bi25FeO40 (b)

Fig. 11
figure 11

Energy band structure of Bi2WO6 and Bi25FeO40

Zhang et al. (2016a, b) had analyzed the double transfer theory of Bi2WO6/Bi25FeO40. The photogenerated electrons in the CB of Bi25FeO40 transfer to the VB of Bi2WO6, while the holes deposited on VB of Bi25FeO40 and electrons were left at CB of Bi2WO6. The photogenerated electrons and holes are effectively separated to improve the photocatalytic performance of the material. It is obvious that the double transfer theory cannot explain the mechanism of Bi2WO6/r-GO/Bi25FeO40 in MB dye degradation under visible light. So, we propose the Z-scheme mechanism of Bi2WO6/r-GO/Bi25FeO40 followed the previous reports and illustrated in Fig. 12. The band gap of Bi2WO6 and Bi25FeO40 is 2.66 and 2.43 eV respectively, while the CB and VB potentials of Bi2WO6 are estimated to be + 0.6 and + 3.26 eV, and for Bi25FeO40, they are − 1.3 and 1.13 eV. As shown in Fig. 12, after irradiated with visible light, the photogenerated electrons in the CB of Bi2WO6 could rapidly transfer to the VB of Bi25FeO40 through r-GO and recombine with holes, and the remaining photogenerated electrons and holes are accumulated on the CB of Bi25FeO40 and the VB of Bi2WO6, respectively. Notably, the CB potential (− 1.3 eV) on Bi25FeO40 is more negative than that of O2/ ̇O2 (− 0.46 eV) (Wu et al. 1998), which can reduce O2 to ̇O2 radicals. Meanwhile, the VB potential (3.26 eV) on Bi2WO6 is more positive than that of H2O/ ̇OH (2.85 eV) (Li et al. 2015), thus the H2O could be oxidized to ̇OH, and the holes with excellent oxidation ability on VB can directly degrade the dye in the photocatalytic process. r-GO plays two important roles in this photocatalytic system, on the one hand, increasing the contact area and reaction space between catalyst and contaminant, and, on the other hand, accelerating the transfer of electrons between Bi2WO6 and Bi25FeO40, effectively suppressing the recombination of photogenerated electron-hole pairs. In summary, the photocatalytic activity of Bi2WO6/r-GO/Bi25FeO40 is greatly improved. In addition, the photocatalytic process of Bi2WO6/r-GO/Bi25FeO40 Z-scheme can be described as follows:

$$ {\displaystyle \begin{array}{l}{\mathrm{Bi}}_2{\mathrm{WO}}_6+\mathrm{h}\upnu \to {\mathrm{Bi}}_2{\mathrm{WO}}_6\left({\mathrm{e}}^{-}+{\mathrm{h}}^{+}\right)\\ {}{\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}+\mathrm{h}\upnu \to {\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}\left({\mathrm{e}}^{-}+{\mathrm{h}}^{+}\right)\\ {}{\mathrm{Bi}}_2{\mathrm{WO}}_6\left({\mathrm{e}}^{-}+{\mathrm{h}}^{+}\right)+{\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}\left({\mathrm{e}}^{-}+{\mathrm{h}}^{+}\right)\to {\mathrm{Bi}}_2{\mathrm{WO}}_6\left({\mathrm{h}}^{+}\right)+{\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}\left({\mathrm{e}}^{-}\right)\\ {}{\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}\left({\mathrm{e}}^{-}\right)+{\mathrm{O}}_2\to {\mathrm{Bi}}_{25}{\mathrm{FeO}}_{40}+\dot{\mkern6mu}{{\mathrm{O}}_2}^{-}\\ {}{\mathrm{Bi}}_2{\mathrm{WO}}_6\left({\mathrm{h}}^{+}\right)+{\mathrm{H}}_2\mathrm{O}\to \dot{\mkern6mu}\mathrm{OH}+{\mathrm{Bi}}_2{\mathrm{WO}}_6\\ {}{\mathrm{Bi}}_2{\mathrm{WO}}_6\left({\mathrm{h}}^{+}\right)+\dot{\mkern6mu}{{\mathrm{O}}_2}^{-}+\dot{\mkern6mu}\mathrm{OH}+\mathrm{MB}\to {\mathrm{Bi}}_2{\mathrm{WO}}_6+\mathrm{Degradation}\ \mathrm{products}\end{array}} $$
Fig. 12
figure 12

Z-Scheme charge transfer process in the Bi2WO6/r-GO/Bi25FeO40 composites

Conclusion

In summary, the Z-scheme Bi2WO6/r-GO/Bi25FeO40 photocatalyst is successfully prepared by the hydrothermal method, which exhibits high photocatalytic activity in the degradation of MB solution under visible light compared with single Bi2WO6 and Bi25FeO40. Furthermore, Bi2WO6/r-GO/Bi25FeO40 still maintains a high degradation rate for MB solution after four consecutive reuses. In this photocatalytic process, the r-GO is regarded as an electron transport bridge that could rapidly accumulate h+ in VB of Bi2WO6 and e in CB of Bi25FeO40, both of them are the main active substances in dye degradation; therefore, the photocatalytic activity of the ternary composite is improved. We believe this work may provide new insights and perspectives for the preparation of new Z-scheme photocatalysts and be successfully applied in the environmental field.