Next Article in Journal
Effect of Hot Rolling on the Microstructure and Properties of Nanostructured 15Cr ODS Alloys with Al and Zr Addition
Previous Article in Journal
Applications of Bionano Sensor for Extracellular Vesicles Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High MB Solution Degradation Efficiency of FeSiBZr Amorphous Ribbon with Surface Tunnels

1
Key Laboratory for Liquid-Solid Structural Evolution and Processing of Materials (Ministry of Education), School of Materials Science and Engineering, Shandong University, Jinan 250061, China
2
School of Engineering, Edith Cowan University, 270 Joondalup Drive, Joondalup, Perth WA6027, Australia
3
School of Mechanical, Electrical and Vehicle Engineering, Weifang Institute, Weifang 261061, China
4
National Engineering Laboratory for Reducing Emissions from Coal Combustion, Engineering Research Center of Environmental Thermal Technology of Ministry of Education, Shandong Key Laboratory of Energy Carbon Reduction and Resource Utilization, School of Energy and Power Engineering, Shandong University, Jinan 250061, China
*
Author to whom correspondence should be addressed.
Materials 2020, 13(17), 3694; https://doi.org/10.3390/ma13173694
Submission received: 24 July 2020 / Revised: 12 August 2020 / Accepted: 13 August 2020 / Published: 21 August 2020

Abstract

:
The as spun amorphous (Fe78Si9B13)99.5Zr0.5 (Zr0.5) and (Fe78Si9B13)99Zr1 (Zr1) ribbons having a Fenton-like reaction are proved to bear a good degradation performance in organic dye wastewater treatment for the first time by evaluating their degradation efficiency in methylene blue (MB) solution. Compared to the widely studied (Fe78Si9B13)100Zr0 (Zr0) amorphous ribbon for degradation, with increasing cZr (Zr atomic content), the as-spun Zr0, Zr0.5 and Zr1 amorphous ribbons have gradually increased degradation rate of MB solution. According to δ c (characteristic distance) of as-spun Zr0, Zr0.5 and Zr1 ribbons, the free volume in Zr1 ribbon is higher Zr0 and Zr0.5 ribbons. In the reaction process, the Zr1 ribbon surface formed the 3D nano-porous structure with specific surface area higher than the cotton floc structure formed by Zr0 ribbon and coarse porous structure formed by Zr0.5 ribbon. The Zr1 ribbon’s high free volume and high specific surface area make its degradation rate of MB solution higher than that of Zr0 and Zr0.5 ribbons. This work not only provides a new method to remedying the organic dyes wastewater with high efficiency and low-cost, but also improves an application prospect of Fe-based glassy alloys.

Graphical Abstract

1. Introduction

Nowadays, organic dyes wastewater is commonly produced in the industrial production of dyestuff, textiles, paper and plastics. This organic dyes wastewater contains carcinogenic, teratogenic and biological toxic substances, which will cause serious pollution to the environment [1,2,3]. Thus, increasing attention has been paid to the harmful pollution of industrial organic dyes to our water bodies. In the past several decades, numerous research works have been done to reduce their hazardous effects, including the physical adsorption of activated carbon and clay [4,5], biodegradation [6], chemical degradation by advanced oxidation process [7,8,9,10] and degradation of specific alloys [11,12,13,14]. However, these methods have obvious disadvantages such as low-efficiency, high-cost and short service life. Thus, we should actively explore advanced materials to better degrade organic dyes in polluted water [15].
Amorphous alloys (metallic glass alloys) have excellent properties such as high strength and corrosion resistance due to their long-range disordered atomic structures, and the importance of fundamental scientific and engineering application potential is attracting rising attention [16,17,18]. At present, the amorphous alloy ribbons including Mg-based ribbons [19,20,21,22,23], Al-based ribbons [24,25,26], Co-based ribbons [27,28,29] and Fe-based ribbons [30,31,32,33,34,35,36] have been proved to have good degradation properties to organic dyes wastewater. It is well-believed that the excellent degradation ability of amorphous alloys is due to three factors: (1) their high residual stress, (2) thermodynamic instabilityand (3) lots of unsaturated sites on the surface.
Among Fe-, Cu-, Al- and Mg-based amorphous alloys used to degrade organic dyes wastewater, Fe-based alloys have attracted the most attention because of their high degradation efficiency, low cost and good reusability. The frequently used Fe-based alloys for degradation are FeSiB systems, and FeSiB alloys with other elements. Jia et al., using the prepared Fe78Si9B13 and Fe73.5Si13.5B9Cu1Nb3 amorphous ribbons, showed higher degradation efficiency when degrading cibacron brilliant red 3B-A, methyl blue and methyl orange dyes by Fenton-like reaction (Fe0 + H2O2 → Fe2+ + 2OH, Fe2+ + H2O2 → Fe3+ + •OH + OH, •OH + organics → products), which proved that Fe-based amorphous ribbons have better degradation performance of organic dyes than other kinds of amorphous alloys [37,38]. Xie et al. used amorphous Fe76Si9B12Y3 powder to degrade methyl orange dye, which was 1000 times more reactive than industrial iron powder [39].
In this paper, using (Fe78Si9B13)99.5Zr0.5 (Zr0.5) and (Fe78Si9B13)99Zr1 (Zr1) amorphous ribbons, the methylene blue (MB) dyes degradation by Fenton-like reactions is reported for the first time, and is compared with Fe78Si9B13 (Zr0) amorphous ribbon. The addition of trace Zr element effectively adjusts the interatomic force of Zr0 amorphous ribbon, which makes the atoms on the surface of amorphous ribbon participate in Fenton-like reaction more easily, and with the increase of Zr element content, it is easier to form developed 3D nano-porous acicular structure on the ribbon surface, thus increasing the specific surface area of reaction and improving the degradation rate. The effects of initial pH and H2O2 concentration on the degradation efficiency of MB using the Zr0, Zr0.5 and Zr1 amorphous ribbons during Fenton-like reactions are investigated. This work not only provides a new routine for the remediation of organic dyes wastewater, but also extends the application range of Fe-based amorphous/glassy alloys.

2. Experimental

2.1. Materials and Reagent

Alloy ingots with a nominal composition of (Fe78Si9B13)100Zr0 (Zr0, at.%), (Fe78Si9B13)99.5Zr0.5 (Zr0.5, at.%) and (Fe78Si9B13)99Zr1 (Zr1, at.%) were prepared by arc melting of pre-alloyed Fe78Si9B13 ingots and high-purity Zr (99.99 wt.%) in an arc melting furnace (MAM-1 Edmund Buhler, Berlin, Germany), which was vacuumed to 5 × 10−3 Pa first and then filled with purified argon (99.999%). The ribbons with thickness of ~30 μm and width of ~2.5 mm were prepared by single copper roller melt-spinning (i.e., planar flow casting) system. The roller/wheel speed was 42 m·s−1. The amorphous ribbons were cut into 5 cm long strips for degradation tests. Commercially available methylene blue (MB, C16H18ClN3S, AR grade, Tianjin Beichen Fangzheng Reagent Factory, Tianjin, China), Hydrochloric acid (HCl, AR grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), Sodium hydroxide (NaOH, AR grade, Tianjin Hengxing Chemical Reagent Manufacturing Co., Ltd., Tianjin, China) and Hydrogen peroxide (H2O2, AR grade, Tianjin Kemeo Chemical Reagent Co., Ltd. Tianjin, China) are used in the experiment.

2.2. Characterization

The amorphous structure of the as-spun ribbons (in a stripe-like shape) was investigated by X-ray diffraction (XRD, Bruker D8 Discover, Brooke (Beijing) Technology Co., Ltd., Beijing, China) with Cu-Kα radiation and transmission electron microscopy (TEM, FEI Tecnai G2 F20, American FEI Company, Portland, OR, America). The amorphous character of the ribbon samples was also confirmed by differential scanning calorimetry (DSC, Netzsch-404, Netzsch, Bavaria, Germany) at a heating rate of 20 K/min. The surface morphology of the as-spun and reacted ribbons was observed by scanning electron microscope (SEM, JSM-7800F, Japan Electronics Co., Ltd., Beijing, China) equipped with an energy dispersive X-ray spectrometer (EDS).

2.3. Degradation Tests

Preparing MB solution in 500 mL volumetric flask with deionized water (DW), then pour 100 mL MB solution (100 mg L−1) in 250 mL beakers. A certain mass of ribbons (0.5 g L−1) and H2O2 (1 mM) was placed to the MB solution, stirring at a fixed speed (300 r min−1) during the degradation process, and the constant temperature (298 K) of the MB solution was maintained with a water bath. The pH (pH = 3) of the MB solution was adjusted using 12 mol L−1 HCl, as well as 1 M NaOH. At time intervals, a 3 mL solution was extracted with a syringe and filtered with a 0.45 μm membrane, and the concentration of MB solution was monitored in real-time with UV-Vis spectrophotometer (UV-4802) to obtain the absorbance spectrum of the solution.

2.4. Electrochemical Tests

The polarization curves and electrochemical impedance spectra (EIS) were measured using an electrochemical measuring instrument (CHI 660E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) in the 20 mL DW or MB solutions (pH = 3, T = 298 K, CH2O2 = 1 mM and CMB = 100 mg·L−1). The three-electrode cell was used for measurement, the saturated calomel electrode (SCE) was used as reference electrode, platinum was used as counter electrode and as-spun ribbon was used as working electrode. When the open-circuit potential stabilizes, the polarization curve was recorded at the potential scanning speed of 1 mV s−1. EIS was performed in static states, with scanning frequencies from 100 kHz to 0.01 Hz and the amplitude of ± 10 mV.

3. Results

3.1. XRD, DSC and TEM Analysis

Figure 1a shows the x-ray diffraction XRD curves of as-spun (Fe78Si9B13)100Zr0 (Zr0), (Fe78Si9B13)99.5Zr0.5 (Zr0.5) and (Fe78Si9B13)99Zr1 (Zr1) ribbons. The XRD curves of as-spun Zr0, Zr0.5 and Zr1 ribbons have only a typical diffuse scattering peak at 2θmax = 44.45°, 44.32° and 44.14° respectively, indicating that the as-spun Zr0, Zr0.5 and Zr1 ribbons owns a fully amorphous structure. With the increase of Zr content, the diffuse scattering peak moves to a low degree. The mean neighboring atomic distance increases with Zr content due to the large atom size of Zr element [40]. Figure 1b shows the DSC curves of as-spun Zr0, Zr0.5 and Zr1 ribbons. In the DSC curves of as-spun Zr0, Zr0.5 and Zr1 ribbons, there are two crystallization peak temperatures TP1 and TP2, a melting peak temperature TP3 in the heating scan, respectively. The TP2-TP1 (crystallization peak temperature range) of as-spun Zr0, Zr0.5 and Zr1 ribbons are 18, 41 and 68 K, respectively. Not only are the two crystallization peak temperatures TP1 and TP2 of as-spun Zr0.5 and Zr1 ribbons higher than that of as-spun Zr0 ribbon, but also their crystallization peak temperature range of TP2-TP1 higher than that of as-spun Zr0 ribbon, indicating that the as-spun Zr0.5 and Zr1 ribbons have better amorphous stability and formability. The melting peak temperature TP3 of as-spun Zr0.5 and Zr1 ribbons was lower than that of as-spun Zr0 ribbon, indicating that the addition of Zr element can reduce the bonding force between atoms in amorphous ribbons, thus have a lower melting peak temperature.
In order to further characterize the microstructure of the as-spun Zr0, Zr0.5 and Zr1 ribbons, we conducted TEM investigations, which are shown in Figure 2. There is mainly maze shape pattern without crystallites in the high-resolution bright field images of the as-spun Zr0, Zr0.5 and Zr1 ribbons (Figure 2a–c), and the corresponding SAED (Selected area electron diffraction) patterns have two typical diffraction halos (Figure 2d–f), confirming that the as-spun Zr0, Zr0.5 and Zr1 ribbons own a fully amorphous structure. Thus, the results of TEM are agreeing with the XRD curves and DSC curves (Figure 1a,b).

3.2. Degradation Performance

Figure 3a–c exhibits the ultraviolet–visible (UV–vis) absorption spectra of filtered methylene blue (MB) solution in a series of time intervals after adding the as-spun Zr0, Zr0.5 and Zr1 ribbons in the reaction batch, respectively. The UV–vis absorption spectra of MB solution have two major absorption peaks at about 610 nm and 654 nm, which represent the auxochrome and chromophore groups, respectively [38]. The normalized concentration of the MB solution obtains its peak value at 654 nm, which represents the chromogenic species, as shown in Figure 3d. With increasing tr the absorption peak at 654 nm gradually decayed, indicating that the chromophore groups of MB disappeared gradually. In the first 9 min, the auxochrome groups react more quickly with Fe-based ribbons than the chromophore groups. The degradation kinetics is usually described by the pseudo-first-order equation as follows [41]:
Ct = C0 exp(−ktr),
where k is the reaction rate constant (min−1), C0 is the initial concentration of MB solution (mg L−1), tr is the reaction time (min), and Ct is the instant concentration of MB solution (mg L−1) at tr. In this work, the ln (C0/Ct) − tr curves are shown in the inset of Figure 3d. The deduced k of as-spun Zr0.5 and Zr1 ribbons are 0.22 min–1 and 0.24 min–1, which is larger than 0.19 min–1 for as-spun Zr0 ribbon. Here, the fit goodness values R2 of Zr0, Zr0.5 and Zr1 ribbons are 0.97, 0.98 and 0.97, respectively. Thus, the as-spun Zr0.5 and Zr1 ribbons bear a higher degradation performance for MB solution compared with as-spun Zr0 ribbon.

3.3. Surface Morphology

In order to deeply understand the MB solution degradation mechanism with the Zr0, Zr0.5 and Zr1 amorphous ribbons, we test to study the structural evolution of the ribbon surfaces during the Fenton-like reaction process. The SEM images on the surface of the as-spun and reacted Zr0, Zr0.5 and Zr1 ribbons are displayed in Figure 4 and the EDS results are listed in Table 1. The as-spun Zr0, Zr0.5 and Zr1 ribbons have a typical smooth amorphous surface, as shown in Figure 4a–c, respectively. There are cotton floc structures and some corrosion pit on the reacted Zr0 ribbon (Figure 4d). The reacted Zr0.5 ribbon surface appears in coarse porous structure with the ligament width of 100 nm (Figure 4e). The reacted Zr1 ribbon surface has a developed 3D nano-porous structure with some acicular matters on the ligaments (Figure 4f). The porous surface structures should have enhancing effects on the degradation process because they can provide mass transfer channels. Compared with Zr0 and Zr0.5 ribbons, the higher pore density of Zr1 ribbon may be the reason for its higher degradation performance in Fenton-like reaction.
The XRD patterns of Zr0, Zr0.5 and Zr1 ribbons (upper right inset in Figure 4d–f) after reaction still have only a typical diffuse scattering peak at 2θmax = 44.56°, 44.47° and 44.39° respectively, indicating that they still own a fully amorphous structure. Here, the different peak positions of the reacted ribbons are higher than the as-spun ribbons. Moreover, the 2θmax deviation Δ2θmax values of Zr0, Zr0.5 and Zr1 ribbons are 0.11°, 0.15° and 0.25°, respectively, showing a high tendency with increasing cZr. After reacting with MB solution for 15 min, the Zr0, Zr0.5 and Zr1 ribbons have a decreased cFe, and the cZr has remained basically unchanged. Moreover, the decrease of Fe element on the surface of Zr1 ribbon is higher than that of Zr0 and Zr0.5 ribbons, indicating that the high degradation rate of Zr1 ribbon is due to a large amount of Fe element participating in the Fenton-like reaction. After degradation, the cO on Zr0, Zr0.5 and Zr1 ribbon surface increases, indicating that the degradation process involves the oxidation of the ribbons.

3.4. Electrochemical Analysis

The polarization curves and electrochemical impedance spectra (EIS) of the as-spun Zr0, Zr0.5 and Zr1 ribbons in DW and MB solution (T = 298 K, pH = 3, CH2O2 = 1 mM and CMB = 100 mg L−1) are shown in Figure 5. In DW, the corrosion potentials (Ecorr) of the Zr0.5 and Zr1 ribbons are −0.75 and −0.71 V (Figure 5a), higher than the Zr0 ribbon (−0.80 V). The corrosion current densities (icorr) of the Zr0.5 and Zr1 ribbons are 5.83 × 10−6 and 3.30 × 10−6 A·cm−2, lower than the Zr0 ribbon (8.23 × 10−6 A cm−2). In MB solution, the Ecorr of the Zr0.5 and Zr1 ribbons are −0.60 and −0.54 V (Figure 5b), higher than the Zr0 ribbon (−0.65 V). The icorr of the Zr0.5 and Zr1 ribbons are 1.54 × 10−4 and 1.20 × 10−4 A cm−2, lower than the Zr0 ribbon (1.82 × 10−4 A cm−2). The above data from polarization curves indicate that the Zr0.5 and Zr1 ribbons have better corrosion resistance than the Zr0 ribbon in DW and MB solution.
In both DW and MB solution (Figure 5c,d), the Nyquist semicircle diameter of the Zr0.5 and Zr1 ribbons is larger than that of Zr0 ribbon. The equivalent circuit composed of R(Q(R(QR))) is used to fit the EIS data. The fitting error (chi square χ2) for the Zr0, Zr0.5 and Zr1 ribbons are 1.47 × 10−5, 1.07 × 10−4 and 6.54 × 10−4 in DW, respectively; while they are 1.05 × 10−5, 6.77 × 10−6 and 7.71 × 10−6 in MB solution, respectively. In the equivalent circuit, the phase element (CPE) Q is defined as [42]:
Q = (jw)−n/Y0,
where Q is the resistance, j is the imaginary unit, w is the frequency, n is the coefficient of CPE and Y0 is the admittance.
The fitting results are summarized in Table 2. The Rs (solution resistance) of the Zr0, Zr0.5 and Zr1 ribbons in MB solution is lower than that in DW, which may be due to the conductive H+ in MB solution. The Rf (resistance of passivation film) and Ra (resistance of electrochemical reaction) of the Zr0, Zr0.5 and Zr1 ribbons in DW are higher than those in MB solution respectively, due to a certain amount of H+ and methylene blue molecules in the MB solution. The Rtotal (total resistance) of the Zr0, Zr0.5 and Zr1 ribbons in DW is higher than that in MB solution, and the Rtotal of Zr0.5 and Zr1 ribbons are higher than that of Zr0 ribbon in either DW or MB solution. Thus, the EIS results are also in good agreement with the results of polarization curves (Figure 5a,b).

3.5. Effect of pH on Ribbon Degradation

The working pH range of the Fenton-like reaction using the as-spun Zr0, Zr0.5 and Zr1 ribbons for the MB solution degradation has been studied, and keep other reaction conditions constant: T = 298 K, CH2O2 = 1 mM, ribbon dosage = 0.5 g L−1 and CMB = 100 mg L−1. The highest degradation rate for as-spun Zr0, Zr0.5 and Zr1 ribbons is achieved at pH =3, as shown in Figure 6a−c, respectively. Surprisingly, as pH = 2, the degradation efficiency (η = (1 − Ct/C0 × 100%, t = 15 min) of as-spun Zr0, Zr0.5 and Zr1 ribbons is lower than that at pH = 3 (Figure 6d). This may be because the iron in the ribbon dissolved to generate hydrogen when the H+ in the solution is too high (H+ + Fe0 → Fe2+ + H2 ↑). This hydrogen evolution reaction generates a large amount of Fe2+, which may consume •OH and lower the oxidation capability of the MB solution (Fe2+ + •OH → OH + Fe3+). As pH > 3, the η of the MB solution decreases with the increasing pH, as there must be enough H+ in the solution to carry out Fenton-like reactions. When the pH increases to 4, this MB solution hardly degrades before 7 min, and then it degrades slowly. As pH > 5 the η of MB solution is nearly zero.

3.6. Effect of H2O2 Concentration on Ribbon Degradation and Surface Morphology

The concentration of H2O2 controls the rate of •OH generation in the Fenton-like reactions. The effect of the concentration on the degradation process of MB solution using as-spun Zr0, Zr0.5 and Zr1 ribbons was investigated, as shown in Figure 7. Various concentrations of H2O2, including 0, 0.5, 1, 5, 10, 30 and 50 mM, are added to MB solution, and other reaction conditions constant: T = 298 K, pH = 3, ribbon dosage = 0.5 g L−1 and CMB = 100 mg L−1. It is proved that H2O2 is necessary to degrade MB solution, because the η is very low without adding H2O2 (Figure 7d). When the CH2O2 is 0.5 mM, the concentration of MB solution remained basically unchanged after 7 min of degradation, which may be due to the fact that H2O2 was completely consumed and •OH could not be produced in Fenton-like reaction. With the CH2O2 increasing from 1 to 10 mM, the η of Fe-based ribbons increases gradually. When the CH2O2 increases to 30 and 50 mM, the η of MB solution began to decrease significantly. The results show that the appropriate addition of H2O2 can effectively accelerate the degradation process. However, due to the well-known hydroxyl radical scavenging effect, excessive H2O2 is not beneficial to the degradation process (H2O2 + •OH → H2O + •HO2). The oxidation potential of generated radical •HO2 is much lower than that of •OH, which slows the degradation rate of MB solution.
Figure 8 shows the SEM images of as-spun Zr1 ribbon reacted with MB solution at different H2O2 concentrations. The surface of the as-spun Zr1 ribbon is relatively smooth (Figure 4c), but fuzzy 3D nano-porous structures appear on the surface of the ribbon as CH2O2 = 0 mM (Figure 8a). With the CH2O2 increasing from 0.5 to 5 mM, the 3D nano-porous structures develop and coarsen (Figure 4f and Figure 8b,c). As CH2O2 = 10 mM, the 3D nano-porous structure begins to transform into an intersecting grid-like structure (Figure 8d). When the CH2O2 reaches 30 and 50 mM, the intersecting grid-like structure gradually becomes thicker and more developed (Figure 8e,f). Table 3 summarizes the EDS results of the Zr1 ribbon reacted with MB solution at different H2O2 concentrations. Comparing the EDS results of the as-spun and reacted Zr1 ribbons (CH2O2 = 1 mM, Table 1), the Fe content decrement on Zr1 ribbon surface increases gradually with increasing CH2O2, the cSi and cB have remained basically unchanged simultaneously. However, the cO rise on the surface of the Zr1 ribbons increases gradually, which is due to H2O2 having strong oxidizability. Excessive H2O2 will oxidize the metal on the ribbon surface, resulting in a large number of oxides.

4. Discussion

With increasing cZr, the 2θmax of Fe-based ribbons decreases gradually. Based on the element date [43], the average atomic radius r ¯ of Zr0, Zr0.5 and Zr1 ribbons is 1.301, 1.302 and 1.303 Å, respectively. According to the equation on the characteristic distance [44,45,46]:
δ c = 2 π Q P ,   with   Q P = ( 4 π   sin θ max ) / λ ,
where δ c is the characteristic distance, Q P is the diffraction vector of the diffusive maximum of the XRD pattern, and the λ is the X-ray wavelength (Cu-Kα = 1.5418 Å). The δ c value of as-spun Zr0, Zr0.5 and Zr1 ribbons is 2.036, 2.042 and 2.050 Å, respectively. Here, the r ¯ Zr 0.5 r ¯ Zr 0 and r ¯ Zr 1 r ¯ Zr 0 equal 1.00077 and 1.00154; meanwhile, δ c , Zr 0.5 δ c , Zr 0 and δ c , Zr 1 δ c , Zr 0 equal 1.00295 and 1.00688, which are much higher than the corresponding r ¯ radius. The high δ c , Zr 1 δ c , Zr 0 suggest that Zr atom can introduce free volume into the glassy matrix [47]. On the other hand, Zr atom can enhance the stability of Fe-Zr and Fe-B bonds in the liquid state and glassy state, according to the lower TP3 and higher TP1, TP2 than those of Zr0 ribbon. It is known that Fe-B can form the network in the Fe-based glasses [48]. Thus, it is expected that the corrosion resistance deduced from potential dynamic scans of the as-spun Zr0.5 and Zr1 ribbons is higher than that of the as-spun Zr0 ribbon (Figure 5). It is understood that the free volume in as-spun Zr0.5 and Zr1 ribbons is higher than the as-spun Zr0 ribbon.
With increasing cZr, the as-spun Zr0, Zr0.5 and Zr1 amorphous ribbons have gradually increased degradation rate of MB solution (Figure 3d), indicating an increased exciting ability in Fenton-like reaction. The as-spun Zr0, Zr0.5 and Zr1 amorphous ribbons formed the cotton floc structure, coarse porous structure and 3D nano-porous structure on the ribbon surfaces during Fenton-like reaction with MB solution (Figure 4d−f), respectively. Obviously, the 3D nano-porous structure formed on the surface of Zr1 ribbon has a larger specific surface area, which can provide more reactive sites, thus improving the degradation performance of Zr1 ribbon to MB solution. According to DSC curves and electrochemical tests, the higher resistance and TP1, TP2 of Zr0.5 and Zr1 ribbon can keep the network, i.e., ligaments of the ribbons, being more stable than Zr0 ribbon. On the other hand, the δ c value of reacted Zr0, Zr0.5 and Zr1 ribbons is 2.031, 2.035 and 2.039 Å, respectively, and the δ c decrement values (Δ δ c ) of Zr0, Zr0.5 and Zr1 ribbons after reaction are 0.005, 0.007 and 0.011 Å, respectively. These data indicate that the free volumes have segregated and fromed the pores on the surface of the ribbons. Hence, it is understood that the pore size increases simultaneously.
Under different CH2O2, the Zr0, Zr0.5 and Zr1 ribbons have different η values in MB solution and have a maximal η as CH2O2 = 10 mM (Figure 7d). Meanwhile, in the range of CH2O2 = 0.5–30 mM, the η of Zr0.5 and Zr1 ribbons is higher than Zr0 ribbon, indicating that the existence of Zr element would slow down the change of η in wide CH2O2 range. As CH2O2 is 1 mM, the amount and size of acicular Fe-based oxides on Zr1 ribbon surface are higher than on the Zr0 surface. Hence, the interface of FexOy/matrix on Zr0 ribbon surface is more compatible than on Zr1 ribbon surface, indicating that the Fe atom below the Zr1 ribbon surface can move to the outer layer more easily than on the Zr0 ribbon surface. In addition, under the help of the porous structure, the specific surface area of the Zr1 ribbon is higher and the η is higher than Zr0 ribbon. As CH2O2 reaches 30 mM, the amount and size of FexOy are very high, and cover the pore mouth and inhibit the inner Fe atom transport to the surface to join Fenton-like reaction. This explains why the η of Fe-based ribbons decreases as CH2O2 increases from 30 mM.
Based on the analysis of the elemental information, surface micro-morphology and micro-structure of Zr0 and Zr1 amorphous ribbons during MB solution (T = 298 K, pH = 3, CH2O2 = 1 mM and CMB = 100 mg L−1) degradation, the pathway of this Fenton-like reaction can be drawn, as shown schematically in Figure 9. According to δ c of as-spun Zr0 and Zr1 ribbons, the free volume in Zr1 ribbon is higher Zr0 ribbon. In the reaction process, the Fe atoms can diffuse under the help of free volume and more to the surface to join the Fenton-like reaction to degrade the MB solution. At the same time, the free volume can segregate to form the pore in the reaction. Meanwhile, TP1 and TP2 of the Zr1 ribbon is higher than that of the Zr0 ribbon, which can make the network in the glassy matrix more stable. Hence, the ligament is more easily formed in the reacted Zr1 ribbon. Moreover, FexOy is not compatible with ligaments and leave the channel active, which helps iron atoms to transport toward the surface from the inner part of the ribbon matrix. Hence, the Zr1 ribbon has a higher η and k than Zr0 ribbon under the same condition. These results not only prove that Zr1 amorphous ribbon has high degradability to organic dyes, but also reveal that Zr element has the role of tunnel construction in Fe-based alloys.

5. Conclusions

In this work, we have prepared Zr0, Zr0.5 and Zr1 amorphous ribbons with melt-spun method and we studied the microstructure, MB solution degradation behavior with several technologies. With increasing cZr, the as-spun Zr0, Zr0.5 and Zr1 amorphous ribbons have gradually increased degradation rate of MB solution. According to δ c of as-spun Zr0, Zr0.5 and Zr1 ribbons, the free volume in Zr1 ribbon is higher for Zr0 and Zr0.5 ribbons. In the reaction process, the 3D nano-porous structure formed on the surface of Zr1 ribbon has a higher specific surface area than the cotton floc structure formed by Zr0 ribbon and coarse porous structure formed by Zr0.5 ribbon. We prove that the high free volume makes it easy to form a pore structure in the reaction process, and the construction of these tunnels is beneficial to the transport of iron atoms inner the ribbon to the surface. Thus, the Zr1 ribbon has high η and k than Zr0 and Zr0.5 ribbons under the same condition. The nonmonotonic influence of pH value and H2O2 concentration on the degradation η and k of Zr1 ribbon is similar to Zr0 and Zr0.5 ribbons. This work not only provides a new method for the remediation of organic dye wastewater, but also extends the application prospect of Fe-based amorphous alloys.

Author Contributions

Conceptualization, Q.C. and Z.Y.; methodology, Q.C. and H.Z.; investigation, Q.C. and L.-C.Z.; data curation, H.M. and W.W. (Wenlong Wang); writing—original draft preparation, Q.C.; writing—review and editing, Q.C. and W.W. (Weimin Wang); supervision, L.-C.Z. and W.W. (Weimin Wang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research Program of China, grant number 2016YFB0300501, National Natural Science Foundation of China, grant number 51471099, 51571132, 51511140291 and 51771103, Key R&D Plan of Shandong Province grant number grant number 2019GGX102016.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lachheb, H.; Puzenat, E.; Houas, A.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV-irradiated titania. Appl. Catal. B 2002, 39, 75–90. [Google Scholar] [CrossRef]
  2. Fu, F.L.; Dionysiou, D.D.; Liu, H. The use of zero-valent iron for groundwater remediation and wastewater treatment: A review. J. Hazard. Mater. 2014, 267, 194–205. [Google Scholar] [CrossRef] [PubMed]
  3. Park, H.; Choi, W. Visible light and Fe(III)-mediated degradation of Acid Orange 7 in the absence of H2O2. J. Photochem. Photobiol. A 2003, 159, 241–247. [Google Scholar] [CrossRef]
  4. Djilani, C.; Zaghdoudi, R.; Djazi, F.; Bouchekima, B.; Lallam, A.; Modarressi, A.; Rogalski, M. Adsorption of dyes on activated carbon prepared from apricot stones and commercial activated carbon. J. Taiwan Inst. Chem. Eng. 2015, 53, 112–121. [Google Scholar] [CrossRef]
  5. Pawar, R.R.; Gupta, P.; Sawant, S.Y.; Shahmoradi, B.; Lee, S.M. Porous synthetic hectorite clay-alginate composite beads for effective adsorption of methylene blue dye from aqueous solution. Int. J. Biol. Macromol. 2018, 114, 1315–1324. [Google Scholar] [CrossRef] [PubMed]
  6. Nigam, P.; Banat, I.M.; Singh, D.; Marchant, R. Microbial process for the decolorization of textile effluent containing azo, diazo and reactive dyes. Process Biochem. 1996, 31, 435–442. [Google Scholar] [CrossRef]
  7. Lucas, M.; Peres, J. Decolorization of the azo dye Reactive Black 5 by Fenton and photo-Fenton oxidation. Dyes Pigm. 2006, 71, 236–244. [Google Scholar] [CrossRef]
  8. Shende, A.G.; Tiwari, C.S.; Bhoyar, T.H.; Vidyasagar, D.; Umare, S.S. BWO nano-octahedron coupled with layered g-C3N4: An efficient visible light active photocatalyst for degradation of cationic/anionic dyes, and N2 reduction. J. Mol. Liq. 2019, 296, 111771. [Google Scholar] [CrossRef]
  9. Tan, W.B.; Wang, L.; Yu, H.X.; Zhang, H.; Zhang, X.H.; Jia, Y.F.; Li, T.T.; Dang, Q.L.; Cui, D.Y.; Xi, B.D. Accelerated microbial reduction of azo dye by using biochar from iron-rich-biomass pyrolysis. Materials 2019, 12, 1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Zhang, Y.Z.; Liu, J.F.; Chen, D.; Qin, Q.D.; Wu, Y.; Huang, F.; Li, W. Preparation of FeOOH/Cu with high catalytic activity for degradation of organic dyes. Materials 2019, 12, 338. [Google Scholar] [CrossRef] [Green Version]
  11. Shen, W.J.; Kang, H.L.; Ai, Z.H. Comparison of aerobic atrazine degradation with zero valent aluminum and zero valent iron. J. Hazard. Mater. 2018, 357, 408–414. [Google Scholar] [CrossRef] [PubMed]
  12. Gupta, N.K.; Ghaffari, Y.; Bae, J.; Kim, K.S. Synthesis of coral-like α-Fe2O3 nanoparticles for dye degradation at neutral pH. J. Mol. Liq. 2020, 301, 112473. [Google Scholar] [CrossRef]
  13. Bhatt, C.S.; Nagaraj, B.; Suresh, A.K. Nanoparticles-shape influenced high-efficient degradation of dyes: Comparative evaluation of nano-cubes vs nano-rods vs nano-spheres. J. Mol. Liq. 2017, 242, 958–965. [Google Scholar] [CrossRef]
  14. Li, X.N.; Li, J.H.; Shi, W.L.; Bao, J.F.; Yang, X.Y. A Fenton-like nanocatalyst based on easily separated magnetic nanorings for oxidation and degradation of dye pollutant. Materials 2020, 13, 332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Tara, N.; Arslan, M.; Hussain, Z.; Iqbal, M.; Khan, Q.M.; Afzal, M. On-site performance of floating treatment wetland macrocosms augmented with dye-degrading bacteria for the remediation of textile industry wastewater. J. Clean. Prod. 2019, 217, 541–548. [Google Scholar] [CrossRef]
  16. Lin, B.; Bian, X.F.; Wang, P.; Luo, G.P. Application of Fe-based metallic glasses in wastewater treatment. Mater. Sci. Eng. B 2012, 177, 92–95. [Google Scholar] [CrossRef]
  17. Zhang, C.Q.; Zhu, Z.W.; Zhang, H.F.; Hu, Z.Q. Rapid reductive degradation of azo dyes by a unique structure of amorphous alloys. Chin. Sci. Bull. 2011, 56, 3988–3992. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, L.C.; Kim, K.B.; Yu, P.; Zhang, W.Y.; Kunz, U.; Eckert, J. Amorphization in mechanically alloyed (Ti, Zr, Nb)-(Cu, Ni)-Al equiatomic alloys. J. Alloys Compd. 2007, 428, 157–163. [Google Scholar] [CrossRef]
  19. Zhao, Y.F.; Si, J.J.; Song, J.G.; Yang, Q.; Hui, X.D. Synthesis of Mg-Zn-Ca metallic glasses by gas-atomization and their excellent capability in degrading azo dyes. Mater. Sci. Eng. B 2014, 181, 46–55. [Google Scholar] [CrossRef]
  20. Wang, J.Q.; Liu, Y.H.; Chen, M.W.; Louzguine-Luzgin, D.V.; Inoue, A.; Perepezko, J.H. Excellent capability in degrading azo dyes by MgZn-based metallic glass powders. Sci. Rep. 2012, 2, 418. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, Q.; Pang, J.; Yan, Z.C.; Hu, Y.H.; Guo, L.Y.; Zhang, H.; Zhang, L.C.; Wang, W.M. MgZn-based amorphous ribbon as a benign decolorizer in methyl blue solution. J. Non Cryst. Solids 2020, 529, 119802. [Google Scholar] [CrossRef]
  22. Chen, Q.; Yan, Z.C.; Guo, L.Y.; Zhang, H.; Zhang, L.C.; Kim, K.; Li, X.Y.; Wang, W.M. Enhancing the acid orange dye degradation efficiency of Mg-based glassy alloys with introducing porous structure and zinc oxide. J. Alloys Compd. 2020, 831, 154817. [Google Scholar] [CrossRef]
  23. Zhang, C.Q.; Zhu, Z.W.; Zhang, H.F. Mg-based amorphous alloys for decolorization of azo dyes. Results Phys. 2017, 7, 2054–2056. [Google Scholar] [CrossRef]
  24. Shaheen, K.; Suo, H.L.; Shah, Z.; Khush, L.; Arshad, T.; Khan, S.B.; Siddique, M.; Ma, L.; Liu, M.; Cui, J.; et al. Ag-Ni and Al-Ni nanoparticles for resistive response of humidity and photocatalytic degradation of methyl orange dye. Mater. Chem. Phys. 2020, 244, 122748. [Google Scholar] [CrossRef]
  25. Das, S.; Garrison, S.; Mukherjee, S. Bi-functional mechanism in degradation of toxic water pollutants by catalytic amorphous metals. Adv. Eng. Mater. 2016, 18, 214–218. [Google Scholar] [CrossRef]
  26. Wang, P.P.; Wang, J.Q.; Li, H.; Yang, H.; Huo, J.T.; Wang, J.G.; Chang, C.T.; Wang, X.M.; Li, R.W.; Wang, G. Fast decolorization of azo dyes in both alkaline and acidic solutions by Al-based metallic glasses. J. Alloys Compd. 2017, 701, 759–767. [Google Scholar] [CrossRef]
  27. Sha, Y.Y.; Mathew, I.; Cui, Q.Z.; Clay, M.; Gao, F.; Zhang, X.Q.J.; Gu, Z.Y. Rapid degradation of azo dye methyl orange using hollow cobalt nanoparticles. Chemosphere 2016, 144, 1530–1535. [Google Scholar] [CrossRef]
  28. Mondal, A.; Adhikary, B.; Mukherjee, D. Room-temperature synthesis of air stable cobalt nanoparticles and their use as catalyst for methyl orange dye degradation. Colloids Surf. 2015, 482, 248–257. [Google Scholar] [CrossRef]
  29. Taneja, P.; Sharma, S.; Umar, A.; Mehta, S.K.; Ibhadon, A.O.; Kansal, S.K. Visible-light driven photocatalytic degradation of brilliant green dye based on cobalt tungstate (CoWO4) nanoparticles. Mater. Chem. Phys. 2018, 211, 335–342. [Google Scholar] [CrossRef]
  30. Wang, J.Q.; Liu, Y.H.; Chen, M.W.; Xie, G.Q.; Louzguine Luzgin, D.V.; Inoue, A.; Perepezko, J.H. Rapid degradation of Azo dye by Fe-based metallic glass powder. Adv. Funct. Mater. 2012, 22, 2567–2570. [Google Scholar] [CrossRef]
  31. Wang, Q.Q.; Chen, M.X.; Lin, P.H.; Cui, Z.Q.; Chu, C.G.; Shen, B.L. Investigation of FePC amorphous alloys with self-renewing behaviour for highly efficient decolorization of methylene blue. J. Mater. Chem. A 2018, 6, 10686–10699. [Google Scholar] [CrossRef]
  32. Liu, P.; Zhang, J.L.; Zha, M.Q.; Shek, C.H. Synthesis of an Fe rich amorphous structure with a catalytic effect to rapidly decolorize Azo dye at room temperature. ACS Appl. Mater. Interfaces 2014, 6, 5500–5505. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, L.C.; Jia, Z.; Lyu, F.; Liang, S.X.; Lu, J. A review of catalytic performance of metallic glasses in wastewater treatment: Recent progress and prospects. Prog. Mater. Sci. 2019, 105, 100576. [Google Scholar] [CrossRef]
  34. Liang, S.X.; Jia, Z.; Liu, Y.J.; Zhang, W.C.; Wang, W.M.; Lu, J. Compelling rejuvenated catalytic performance in metallic glasses. Adv. Mater. 2018, 30, 1802764. [Google Scholar] [CrossRef] [Green Version]
  35. Jia, Z.; Zhang, W.C.; Wang, W.M.; Habibi, D.; Zhang, L.C. Amorphous Fe78Si9B13 alloy: An efficient and reusable photo-enhanced Fenton-like catalyst in degradation of cibacron brilliant red 3B-A dye under UV-vis light. Appl. Catal. B 2016, 192, 46–56. [Google Scholar] [CrossRef]
  36. Liang, S.X.; Wang, X.Q.; Zhang, W.C.; Liu, Y.J.; Wang, W.M.; Zhang, L.C. Selective laser melting manufactured porous Fe-based metallic glass matrix composite with remarkable catalytic activity and reusability. Appl. Mater. Tod. 2020, 19, 100543. [Google Scholar] [CrossRef]
  37. Jia, Z.; Liang, S.X.; Zhang, W.C.; Wang, W.M.; Yang, C.; Zhang, L.C. Heterogeneous photo Fenton-like degradation of cibacron brilliant red 3B-A dye using amorphous Fe78Si9B13 and Fe73.5Si13.5B9Cu1Nb3 alloys: The influence of adsorption. J. Taiwan Inst. Chem. Eng. 2017, 71, 128–136. [Google Scholar] [CrossRef]
  38. Jia, Z.; Kang, J.; Zhang, W.C.; Wang, W.M.; Yang, C.; Sun, H.; Habibi, D.; Zhang, L.C. Surface aging behaviour of Fe-based amorphous alloys as catalysts during heterogeneous photo Fenton-like process for water treatment. Appl. Catal. B 2017, 204, 537–547. [Google Scholar] [CrossRef]
  39. Xie, S.; Huang, P.; Kruzic, J.J.; Zeng, X.; Qian, H. A highly efficient degradation mechanism of methyl orange using Fe-based metallic glass powders. Sci. Rep. 2016, 6, 21947. [Google Scholar] [CrossRef] [Green Version]
  40. Tschopp, M.A.; Horstemeyer, M.F.; Gao, F.; Sun, X.; Khaleel, M. Energetic driving force for preferential binding of self-interstitial atoms to Fe grain boundaries over vacancies. Scr. Mater. 2011, 64, 908–911. [Google Scholar] [CrossRef] [Green Version]
  41. Nam, S.; Tratnyek, P.G. Reduction of azo dyes with zero-valent iron. Water Res. 2000, 34, 1837–1845. [Google Scholar] [CrossRef]
  42. Jia, C.G.; Pang, J.; Pan, S.P.; Zhang, Y.J.; Kim, K.B.; Qin, J.Y.; Wang, W.M. Tailoring the corrosion behavior of Fe-based metallic glasses through inducing Nb-triggered netlike structure. Corros. Sci. 2019, 147, 94–107. [Google Scholar] [CrossRef]
  43. Li, G.H.; Wang, W.M.; Bian, X.F.; Zhang, J.T.; Li, R.; Wang, L. Comparing the dynamic and thermodynamic behaviors of Al86Ni9-La5/(La0.5Ce0.5)5 amorphous alloys. J. Alloys Compd. 2009, 478, 745–749. [Google Scholar] [CrossRef]
  44. Guo, L.Y.; Wang, X.; Shen, K.C.; Kim, K.B.; Lan, S.; Wang, X.L.; Wang, W.M. Structure modification and recovery of amorphous Fe73.5Si13.5B9Nb3Cu1 magnetic ribbons after autoclave treatment: SAXS and thermodynamic analysis. J. Mater. Sci. Technol. 2019, 35, 118–126. [Google Scholar] [CrossRef]
  45. Sokolov, A.P.; Kisliuk, A.; Soltwisch, M.; Quitmann, D. Medium-range order in glasses: Comparison of Raman and diffraction measurements. Phys. Rev. Lett. 1992, 69, 1540–1542. [Google Scholar] [CrossRef]
  46. Wang, W.M.; Zhang, W.X.; Gebert, A.; Roth, S.; Mickel, C.; Schultz, L. Microstructure and Magnetic Properties in Fe61Co9−xZr8Mo5WxB17 (0 ≤ × ≤ 3) Glasses and Glass-Matrix Composites. Metall. Mater. Trans. A 2009, 40, 511–521. [Google Scholar] [CrossRef]
  47. Yavari, A.R.; Moulec, A.L.; Inoue, A.; Nishiyama, N.; Lupu, N.; Matsubara, E.; Botta, W.J.; Vaughan, G.; Michiel, M.D.; Kvick, Å. Excess free volume in metallic glasses measured by X-ray diffraction. Acta Mater. 2005, 53, 1611–1619. [Google Scholar] [CrossRef]
  48. Muneyuki, I.; Shigeo, S.; Hisato, K.; Matsubara, E.; Inoue, A. Crystallization behavior of amorphous Fe90−XNb10BX (X = 10 and 30) alloys. Mater. Trans. 2000, 11, 1526–1529. [Google Scholar]
Figure 1. (a) The X-ray diffraction (XRD) curves of as-spun Zr0, Zr0.5 and Zr1 ribbons, (b) the differential scanning calorimetry (DSC) curves of the as-spun Zr0, Zr0.5 and Zr1 ribbons.
Figure 1. (a) The X-ray diffraction (XRD) curves of as-spun Zr0, Zr0.5 and Zr1 ribbons, (b) the differential scanning calorimetry (DSC) curves of the as-spun Zr0, Zr0.5 and Zr1 ribbons.
Materials 13 03694 g001
Figure 2. The transmission electron microscopy (TEM) images of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons, the SAED (Selected area electron diffraction)patterns of the as-spun (d) Zr0, (e) Zr0.5 and (f) Zr1 ribbons.
Figure 2. The transmission electron microscopy (TEM) images of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons, the SAED (Selected area electron diffraction)patterns of the as-spun (d) Zr0, (e) Zr0.5 and (f) Zr1 ribbons.
Materials 13 03694 g002aMaterials 13 03694 g002b
Figure 3. The ultraviolet–visible (UV–Vis) absorbance spectra of methylene blue (MB) solution during the Fenton-like reactions using as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons and (d) the normalized concentration change of MB solution during the degradation process. The inset in (d): the ln(C0/Ct)-tr curves for as-spun Zr0, Zr0.5 and Zr1 ribbons (T = 298 K, pH = 3, CH2O2 = 1 mM, ribbon dosage = 0.5 g L−1 and CMB = 100 mg L−1). Symbols show the experimental data while solid lines are fitting results.
Figure 3. The ultraviolet–visible (UV–Vis) absorbance spectra of methylene blue (MB) solution during the Fenton-like reactions using as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons and (d) the normalized concentration change of MB solution during the degradation process. The inset in (d): the ln(C0/Ct)-tr curves for as-spun Zr0, Zr0.5 and Zr1 ribbons (T = 298 K, pH = 3, CH2O2 = 1 mM, ribbon dosage = 0.5 g L−1 and CMB = 100 mg L−1). Symbols show the experimental data while solid lines are fitting results.
Materials 13 03694 g003aMaterials 13 03694 g003b
Figure 4. SEM micrographs of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons, and reacted (d) Zr0, (e) Zr0.5 and (f) Zr1 ribbons The insets in (df): the high-magnification images and XRD patterns of reacted Zr0, Zr0.5 and Zr1 ribbons.
Figure 4. SEM micrographs of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons, and reacted (d) Zr0, (e) Zr0.5 and (f) Zr1 ribbons The insets in (df): the high-magnification images and XRD patterns of reacted Zr0, Zr0.5 and Zr1 ribbons.
Materials 13 03694 g004
Figure 5. Polarization curves of the as-spun Zr0, Zr0.5 and Zr1 ribbons in (a) DW and (b) MB solution (T = 298 K, pH = 3, CH2O2 = 1 mM and CMB = 100 mg L−1) and their Nyquist curves in (c) DW and (d) MB solution. The insets in (c,d): the general fitted circuit. Symbols show the experimental data while solid lines are fitting results.
Figure 5. Polarization curves of the as-spun Zr0, Zr0.5 and Zr1 ribbons in (a) DW and (b) MB solution (T = 298 K, pH = 3, CH2O2 = 1 mM and CMB = 100 mg L−1) and their Nyquist curves in (c) DW and (d) MB solution. The insets in (c,d): the general fitted circuit. Symbols show the experimental data while solid lines are fitting results.
Materials 13 03694 g005
Figure 6. The normalized concentration Ct/C0 change of MB solution during the degradation process of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons at different pH values. (d) The degradation efficiency (η = (1 − Ct/C0 × 100%, t = 15 min) of the degradation process vs. pH for Zr0, Zr0.5 and Zr1 ribbons.
Figure 6. The normalized concentration Ct/C0 change of MB solution during the degradation process of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons at different pH values. (d) The degradation efficiency (η = (1 − Ct/C0 × 100%, t = 15 min) of the degradation process vs. pH for Zr0, Zr0.5 and Zr1 ribbons.
Materials 13 03694 g006aMaterials 13 03694 g006b
Figure 7. The normalized concentration Ct/C0 change of MB solution during the degradation process of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons at different H2O2 concentration. (d) The degradation efficiency (η = (1 − Ct/C0 × 100%, t = 15 min) of the degradation process vs. CH2O2 for Zr0, Zr0.5 and Zr1 ribbons.
Figure 7. The normalized concentration Ct/C0 change of MB solution during the degradation process of the as-spun (a) Zr0, (b) Zr0.5 and (c) Zr1 ribbons at different H2O2 concentration. (d) The degradation efficiency (η = (1 − Ct/C0 × 100%, t = 15 min) of the degradation process vs. CH2O2 for Zr0, Zr0.5 and Zr1 ribbons.
Materials 13 03694 g007aMaterials 13 03694 g007b
Figure 8. SEM micrographs of (a) H2O2 = 0 mM, (b) H2O2 = 0.5 mM, (c) H2O2 = 5 mM, (d) H2O2 = 10 mM, (e) H2O2 = 30 mM and (f) H2O2 = 50 mM for the as-spun Zr1 ribbons after reacted with MB solution. The insets in (af): the high-magnification images.
Figure 8. SEM micrographs of (a) H2O2 = 0 mM, (b) H2O2 = 0.5 mM, (c) H2O2 = 5 mM, (d) H2O2 = 10 mM, (e) H2O2 = 30 mM and (f) H2O2 = 50 mM for the as-spun Zr1 ribbons after reacted with MB solution. The insets in (af): the high-magnification images.
Materials 13 03694 g008aMaterials 13 03694 g008b
Figure 9. Schematic diagrams of the pathway of MB solution degradation in Fenton-like reaction using the Zr0 and Zr1 amorphous ribbons.
Figure 9. Schematic diagrams of the pathway of MB solution degradation in Fenton-like reaction using the Zr0 and Zr1 amorphous ribbons.
Materials 13 03694 g009
Table 1. Energy dispersive X-ray spectrometer (EDS) analysis of the Zr0, Zr0.5 and Zr1 ribbons before and after reacted (at.%).
Table 1. Energy dispersive X-ray spectrometer (EDS) analysis of the Zr0, Zr0.5 and Zr1 ribbons before and after reacted (at.%).
AlloyBefore ReactedAfter Reacted
cFecSicBcZrcOcFecSicBcZrcO
Zr077.88.710.9-2.676.38.610.5-4.6
Zr0.575.57.913.00.72.962.17.111.30.818.7
Zr176.19.010.31.43.253.48.59.61.626.9
Table 2. Parameters from EIS measurements: Rs, solution resistance; Qf and Rf, resistance of passivation film; Qa and Ra, resistance of electrochemical reaction; Rtotal, total resistance.
Table 2. Parameters from EIS measurements: Rs, solution resistance; Qf and Rf, resistance of passivation film; Qa and Ra, resistance of electrochemical reaction; Rtotal, total resistance.
SolutionAlloyRs
(Ω·cm2)
QfRf
(Ω·cm2)
QaRa
(Ω·cm2)
Rtotal
(Ω·cm2)
Yf
−1·s−n·cm−2)
NfYa
−1·s−n·cm−2)
Na
DWZr095.35.3 × 10−90.99895.75.2 × 10−90.98511.01502.0
Zr0.5125.73.4 × 10−90.991932.23.5 × 10−80.85593.02650.9
Zr1147.67.8 × 10−90.952842.38.7 × 10−90.99716.43706.3
MOZr018.34.8 × 10−80.99211.23.4 × 10−50.9128.4257.9
Zr0.521.13.9 × 10−80.98250.73.0 × 10−50.9233.7305.5
Zr124.53.8 × 10−80.99274.72.5 × 10−50.9339.3338.5
Table 3. EDS analysis of the Zr1 ribbon reacted with MB solution at different H2O2 concentrations (at.%).
Table 3. EDS analysis of the Zr1 ribbon reacted with MB solution at different H2O2 concentrations (at.%).
ElementCH2O2 (mM)
00.55103050
cFe69.761.147.243.936.429.5
cSi8.38.58.78.18.68.8
cB9.610.89.29.99.79.5
cZr1.51.41.71.61.41.3
cO10.918.233.236.543.950.9

Share and Cite

MDPI and ACS Style

Chen, Q.; Yan, Z.; Zhang, H.; Zhang, L.-C.; Ma, H.; Wang, W.; Wang, W. High MB Solution Degradation Efficiency of FeSiBZr Amorphous Ribbon with Surface Tunnels. Materials 2020, 13, 3694. https://doi.org/10.3390/ma13173694

AMA Style

Chen Q, Yan Z, Zhang H, Zhang L-C, Ma H, Wang W, Wang W. High MB Solution Degradation Efficiency of FeSiBZr Amorphous Ribbon with Surface Tunnels. Materials. 2020; 13(17):3694. https://doi.org/10.3390/ma13173694

Chicago/Turabian Style

Chen, Qi, Zhicheng Yan, Hao Zhang, Lai-Chang Zhang, Haijian Ma, Wenlong Wang, and Weimin Wang. 2020. "High MB Solution Degradation Efficiency of FeSiBZr Amorphous Ribbon with Surface Tunnels" Materials 13, no. 17: 3694. https://doi.org/10.3390/ma13173694

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

Article Metrics

Back to TopTop