Catalytic applications of amorphous alloys in wastewater treatment: A review on mechanisms, recent trends, challenges and future directions

Yulong Liu Haoran Lu Tong Yang Peng Cheng Xu Han Wenyan Liang

Citation:  Yulong Liu, Haoran Lu, Tong Yang, Peng Cheng, Xu Han, Wenyan Liang. Catalytic applications of amorphous alloys in wastewater treatment: A review on mechanisms, recent trends, challenges and future directions[J]. Chinese Chemical Letters, 2024, 35(10): 109492. doi: 10.1016/j.cclet.2024.109492 shu

Catalytic applications of amorphous alloys in wastewater treatment: A review on mechanisms, recent trends, challenges and future directions

English

  • Water pollution has become a serious environmental problem along with the rapid development of modern industries. With the improvement of environmental awareness, wastewater treatment techniques have been vigorously developed [1]. Advanced oxidation processes (AOPs), including photocatalysis [2], ozone [3], Fenton and Fenton-like oxidation [4,5], and electrochemical catalysis [6], have been extensively studied as promising techniques due to their superior degradation and mineralization efficiency on pollutants in wastewater. Highly reactive and transitory species are the primary oxidants employed in AOPs and can be activated by catalysts [7]. For instance, the reported single-atom catalysts and high-valent metal species in recent years have been applied in the field of environmental remediation because of their good selectivity [8-10]. However, zero-valent metal single atoms are extremely unstable, requiring solid host materials to bear ligating atoms to form chemical bonds with metal single atoms [11]. While amorphous alloys can overcome such problems, because the zero-valent metal atoms are embedded in the alloy, such as Fe0 in Fe-based amorphous alloys.

    Amorphous alloys, also called metallic glasses, with short-range ordered and long-range disordered atomic structures, have recently attracted increasing attention as advanced functional materials [12]. As shown in Table 1 [13-21], many unique properties, such as high strength and elasticity, superior anti-corrosion behavior, excellent wear resistance and attractive soft/hard magnetic performance, compared to crystalline counterparts or pure metals, have been discovered for amorphous alloys. Moreover, these investigations could provide new insights into commercial applications and fundamental studies. The first known successful preparation of amorphous alloy was gold-silicon alloy, prepared in 1960, which was observed with highly disordered arrangements of the atoms, similar to that of the liquid state [22]. After that, amorphous alloys with the unique atomic packing structure as metastable materials have been carried out by plenty of researchers and widely used in space exploration [23,24], resonators, electrical components and biomedical equipment [25-27].

    Table 1

    Table 1.  Properties of common amorphous alloys.
    DownLoad: CSV

    Recent reports demonstrate that amorphous alloys have both favorable catalytic capacity and unique selectivity properties and have become a new promising catalyst for wastewater treatment [28]. Many research outcomes have exhibited the fast-catalytic efficiency, excellent sustainability, and high stability of amorphous alloys [29]. For instance, concerning water treatment, the Fe76B12Si9Y3 powder showed 1000 times higher reactivity than crystalline Fe powder in treating methyl orange [30]. Moreover, the Fe-B amorphous alloy delivered 1.8 times faster treatment than the crystalline counterpart in treating direct blue 6 [31]. In addition, the catalyst regeneration and recyclability improvement are also important to realize practical applications using the metal-based catalysts [32]. Amorphous alloy ribbons with excellent stability have potential prospects in the field of practical water treatment. At present, there already have been some reviews on the processing method, properties, catalytic performance and applications of amorphous alloys in wastewater treatment [27,29,33]. However, as more and more attentions and developments have been paid to amorphous alloys, the study of their catalytic applications has also increased rapidly. Therefore, it is essential to provide a comprehensive summary of recent achievements in water or wastewater treatment using amorphous alloys as catalysts.

    In this paper, the preparation of amorphous alloys, particularly as catalysts, is initially introduced to demonstrate the variety of morphology. Then, the developments in the effects of their physical characteristics, such as atomic components, electronic structure and atomic configuration, on catalytic performance are described. Additionally, the present review shows several benefits of using amorphous alloys in catalytic degradation processes, including experimental parameter effects, stability and reusability in environmental applications. Existing catalytic mechanisms, such as oxidative and reductive degradation, are also discussed in detail. The last section summarizes a few remaining technical challenges, presenting future directions in this field.

    Amorphous alloys are materials whose atomic structure differs from traditional crystalline alloys [34], which can be determined by characterization (Note S1 and Fig. S1 in Supporting information). The atomic arrangement of amorphous alloys shows the characteristics of short-range order and long-range disorder (Fig. S2a in Supporting information). It implies that the atoms in amorphous alloys are arranged in space without periodicity and translational symmetry. In contrast, the bonding (parameters such as coordination number, atomic spacing, bond angle and bond length) between the nearest or next nearest neighbor atoms has certain regularity [35]. In general, the short-range order can be divided into chemical and geometrical short-range order. The chemical short-range order is used to describe the chaotic state of the various elements in the alloy. It is mainly used to measure the degree that the chemical composition around each atom deviates from the average composition of the entire alloy. The geometrical short-range order describes the structure of the alloy from the geometric arrangement, ignoring the differences in element types and regards each atom as an abstract geometric point. The atomic bonding mode of amorphous alloys is mainly based on metal bonds, which have no directionality. Therefore, it has some characteristics of metals and alloys [36,37]. Currently, the technology for preparing these alloys is well-developed (Note S2 in Supporting information). The amorphous formation is based on the transition between the metastable liquid phase and metastable phase (Fig. S2b in Supporting information). The preparation methods of amorphous alloys are briefly summarized in Table S1 (Supporting information) and the schematic diagrams in Figs. S2c, S3, and S4 (Supporting information).

    Only a few specific elements and alloys have been applied as catalytic materials, although many combinations of elements in the periodic table could form amorphous structures [38]. Generally, there are mainly two types of amorphous alloys used as catalysts. One type is transition metals of Ⅷ group in the periodic table and metalloid combination, such as Fe-B, Ni-B, Ni-P and Co-Si-B. Another type is a metal and metal combination, such as Cu-Zr and Ni-Zr [39]. Easy manipulation of atomic composition is a superior advantage for amorphous alloys in the catalytic field [40]. Adding metal elements could not only improve the stability of these structures but also adjust the electronic structure [41]. Zhang et al. compared the effects of Co, Cr and Ni addition on the Fe-Si-B in the degradation of acid orange Ⅱ [42]. They discovered that Co could reduce the activation energy without losing catalytic efficiency. Simultaneously, Cr eradicated the dye degradation ability across a wide temperature range, and Ni changed the decolorization mechanism from dye degradation to dye adsorption. For the addition of metalloids (Fig. 1), adding B facilitated hybrid structures of Fe-based amorphous alloys [43]. Moreover, the generation of three galvanic cells improved degradation efficiency of methyl orange.

    Figure 1

    Figure 1.  Schematic illustration of the different degradation mechanisms for B element facilitated Fe-based amorphous alloy catalysts. Reprinted with permission [43]. Copyright 2019, Wiley‐VCH Verlag GmbH & Co. KGaA, Weinheim.

    It is well known that the electronic structure of a catalyst could play a vital influence on the catalytic performance. Moreover, an electron could be an efficient catalyst to promote different radical cascade reactions [44]. Unlike the crystalline catalysts with a highly ordered atomic structure, the amorphous alloys with a disordered atomic structure could possess plentiful possibilities to tailor their electronic structures in the catalytic application. Hu et al. found that the Ni40Fe40P20 possesses the metallic Ni-Ni and Fe-Fe bonds and the Ni-P and Fe-P bonds due to partial Ni or Fe atoms coordinated with P atoms [45]. The formation of Ni-P and Fe-P bonds improved the electron transport at a moderate distance, facilitating the optimization of their electronic structure and thereby promoting the corresponding catalytic efficiency.

    Compared with the well-defined atomic configuration in crystalline catalysts, the amorphous alloy catalysts could provide more active sites in catalytic reactivity. Hu et al. found that the Pd-based amorphous alloys could possess many active sites with various local atomic packing structures [46]. In addition, the atomic configuration of these Pd-based alloys could also significantly contribute to the catalytic performance of the material. Yang et al. prepared the three-dimensional hierarchical porous architectures introducing Cu as a reductant into amorphous alloys via laser 3D printing technique [47]. The materials exhibited exceptional catalytic efficiency in degrading Rhodamine B. Due to the crystallization of the amorphous phase during the 3D printing process, a certain proportion of nanocrystalline Fe-based metal phosphides was formed in the catalysts, exhibiting a positive effect on the catalytic activity. The theoretical calculations using density functional theory (DFT) obtained the atomic configuration of Fe3P and amorphous alloy components (Fig. 2). It was indicated that amorphous alloy and metal phosphides were the main active species in the catalysts for activation of H2O2. Such outstanding synergetic atomic configurations are almost unachievable in crystalline catalysts with a specific atomic packing structure, further proving the potential applicability of these alloys in catalysis.

    Figure 2

    Figure 2.  Atomic configuration of Fe3P and amorphous alloy components calculated by DFT. Reprinted with permission [47]. Copyright 2021, American Chemical Society.

    Following the discovery of the unique atomic packing structures of long-range disorder and short-range order, there have been many studies on the underlying structural evolution of amorphous alloys [48]. In general, relaxation phenomena are normally undesirable in structural applications because they result in brittleness. Recently, Chang et al. found that the fast relaxation in amorphous alloys could continue the dynamics of high-temperature liquids [49]. Its carriers could rapidly diffuse, making liquid-like atoms inherited from high-temperature liquids. Molecular dynamics simulations revealed that this fast relaxation arose from string-like diffusion of the liquid-like atoms in these alloys. Importantly, the motions of liquid-like atoms could form string-like structures, suggesting that the fast diffusion of these atoms in amorphous alloys is not random but cooperative. These findings could clarify the mechanism of fast relaxation and reveal deeper insights into the nature of glasses and the glass transition. However, rare work has been done on the effect of relaxation in applied catalysis. Therefore, the effects of the atomic rearrangement during the relaxation on catalysis of amorphous alloy still require more attention.

    Amorphous alloys have been proven to exhibit excellent removal properties for heavy metals in wastewater. Yan et al. used Fe78Si9B13 amorphous ribbons to remove copper ions from wastewater by a spontaneous redox reaction and the displacement reaction of copper ions [50]. The report demonstrated that Fe78Si9B13 could help the electrons rapidly transfer from iron atoms in the active atomic layer to copper ions owing to its unique surface mobility. Abundant galvanic cells are produced due to iron and copper, which is beneficial to the reaction process. In contrast, the surface mobility of the crystalline iron ribbons was much weaker than the alloys due to the stable thermodynamic state. Consequently, amorphous alloys showed higher removal efficiency on heavy metals, lower reaction activation energy and higher corrosion current density than crystalline iron. Liang et al. employed Fe78Si9B13 amorphous alloys in real industrial contaminated water treatment by investigating effective separation of arsenic and reduction of nitrate [51]. They confirmed Fe-based amorphous alloys demonstrates attractively high removal rate of arsenic in 30 min, which is ascribed to synergistic effect of reduction/adsorption by amorphous alloys, precipitation of arsenic sulfide and adsorption of generated iron sulfide. On the other hand, 20 reused times of Fe-based amorphous alloys for nitrate reduction suggests remarkable sustainability.

    For much industrial wastewater, such as the petrochemical industry, pesticides and insecticides, phenolic compounds pose a great threat to the aquatic system due to their toxicity and poor biodegradability [52]. Amorphous alloys have been proposed for the degradation of phenolic compounds [53]. Wang et al. used Fe78Si9B13 as heterogeneous Fenton catalysts in the phenol degradation process [54]. More than 99% of phenol was completely removed within 10 min for a solution containing 1000 mg/L of phenol. They also demonstrated that phenol was degraded by OH decomposed by H2O2 on the surface of Fe78Si9B13. The results suggested that the H2O2 was first adsorbed onto the surface of Fe78Si9B13 to produce OH. The OH reacted with phenol in the ortho and para positions to form catechol and hydroquinone. Further oxidation of the dihydroxybenzenes occurs to produce benzoquinones, and the phenolic ring opens to form lower molecular weight organic compounds that are ultimately oxidized to CO2 and H2O.

    Wan et al. fabricated TiO2 nanowires decorated with a trace amount of amorphous Fe-Si-B microspheres, forming hybrid films by the simple painting method [55]. They demonstrated excellent photo degradability of antibiotic contaminants by these hybrid films. It was found that the films containing 1 wt% Fe-Si-B exhibited great tetracycline degradation at general pH conditions from 2 to 12. The degradation efficiencies of tetracycline reached 98.2% at pH 2 and 85.5% at pH 12 within 4 h. Moreover, the enhanced reusability was attributed to the corrosion resistance of amorphous microspheres Fe-Si-B wound around the TiO2 nanowires.

    Oily wastewater, usually made up of oil and grease, contains high chemical oxygen demand (COD) due to the presence of multiplicate aromatic hydrocarbons and phenolic compounds. Xu et al. prepared Fe78Si9B13 amorphous ribbons to treat simulated petroleum wastewater by a Fenton-like process [56]. The COD and oil & grease removal rates could reach up to 79.6% and 84.6%. It is worth mentioning that the amorphous ribbons exhibited good structural stability, presenting an excellent catalytic performance for petroleum wastewater treatment.

    Numerous reports indicated the degradation of dyeing wastewater or simulated solution using amorphous alloys. The destruction of the chromophore could be considered the main reason for decolorizing dyes [57]. In general, two oxidation processes are considered functional for dyes degradation by amorphous alloys. On the one hand, direct oxidation of dye molecules by electron transfer reaction. Conversely, indirect oxidation generates reactive radicals by activating peroxides [58]. The decomposition of dyes in forming aromatic intermediates is further oxidized to ring opening products and final products of inorganic substances.

    Coking wastewater from steel-making industries generally contains phenols, ammonia, cyanides, thiocyanate, polycyclic aromatic compounds, and other contaminants. Its complex composition and high concentration result in heavy chroma and high COD [59]. Qin et al. applied two types of coking wastewater before anaerobic treatment and after aerobic treatment to examine the catalytic performance of Fe78Si8B14 in a Fenton-like system [60]. Results showed that Fe78Si8B14 amorphous alloys exhibited a superior catalytic capability and the COD removal efficiency of coking wastewater before anaerobic treatment. After the aerobic treatment, 71% within 30 min and 89% within 60 min were achieved under the optimal condition, respectively.

    The chemical bonds of the organic compounds in complex and high concentrations are usually difficult to break, which leads to the ineffectiveness of traditional oxidation materials. Yang et al. investigated the treatment of high-concentration organic wastewater using Fe78Si9B13 amorphous alloys in the Fenton-like system [61]. The high-concentration organics in wastewater mainly consisted of methylbenzene and methanol, with COD being 43,639 mg/L. Results showed that the ultimate removal by amorphous alloys was as high as 85%. It has also been proved that the Fe78Si9B13 played an excellent catalytic activity in the oxidative degradation of organic compounds. Its microstructure transformation during the corrosion process promoted the degradation efficiency of the wastewater treatment process. In fact, industrial wastewater is a coexistence system with many kinds of pollutants. The effect of the interwoven system on the degradation performance needs to be thoroughly considered. Li et al. found that the degradation of carbamazepine in their system was not or little affected by the water matrices [62], including organic matter, various inorganic anions (Cl, SO42−, NO3) and cations (Na+, K+, Ca2+, Mg2+) (Fig. S5 in Supporting information). H2PO4 was found to suppress the reaction significantly because it prevented the formation of active species.

    Several types of wastewater with high organic concentrations have been successfully degraded using amorphous alloy catalysts. Recently, amorphous alloys containing transitional metals were reported as advanced reductive agents owing to efficient chemical reactivity and strong surface stability [31,63]. The reductive chemical reactivity could be strongly based on the electron transfer ability of the catalysts. Due to the disordered atomic packing structure, it would be easier to activate amorphous alloys with weak atomic bonding than to activate their crystalline counterparts. This could provide better electron mobility in the solution to further enhance the organic pollutants' ability to gain electrons [64]. The degradation of pollutants using amorphous alloys also could be combined with AOPs. The catalytic efficiency could be highly dependent on several operational parameters based on the specific processes [65]. In addition, to investigate degradation efficiency, the pseudo-first-order kinetic model is often employed (Eq. 1). Table 2 [12,41,42,64,66-79] summarizes the kinetic rate (kobs) of wastewater degradation by various compositions of amorphous alloys.

    (1)

    Table 2

    Table 2.  The kobs of several amorphous alloys in wastewater degradation.
    DownLoad: CSV

    where C0 and Ct are wastewater concentrations at the initial time and time t, respectively.

    4.1.1   Combined with Fenton-like process

    Amorphous alloys were first used to degrade dye wastewater in 2010. It was found that the Fe-Mo-Si-B achieved a 4 times faster degradation rate than the corresponding crystalline alloy in treating direct blue 2B [73]. Zuo et al. used Fe80Si10B10 and Fe83Si5B8P4 ribbons to degrade methyl blue and rhodamine B by Fenton-like oxidation and found that the color removal reached nearly 100% within 11 min for both the dyes [80]. Wang et al. revealed that Fe83Si5B8P4 enhanced the efficiency of phenol and rhodamine B degradation in combination with H2O2 [54]. Zhao et al. degraded acid orange Ⅱ using Cu47.5Zr46Al6.5 amorphous ribbons [81]. The catalytic mechanism they proposed can be illustrated in Fig. S6 (Supporting information), which followed two pathways: corrosion reaction and catalytic degradation process. The interaction between H+ and Cl will invade and attack the surface of the ribbons because of the pitting reaction, bringing about the copper nanoparticles to be exposed naturally in solution. The exposed copper nanoparticles and the produced Cu2O with strong reducing ability can directly reduce the dye molecules in the solution. In addition, they can both destroy O2 molecules to develop OH under acidic conditions and oxidize azo dye molecules. Moreover, a mere handful of Al elements also have a certain contribution in the catalytic degradation experiments. The first part can be interpreted as the direct reduction of zero-valent Al. Secondly, new reduced hydrogen [H] can act on the reduction of azo bonds under the action of surface catalysis.

    4.1.2   Combined with ultrasonic vibration process

    As a kind of physical field, the ultrasonic treatment approach could have satisfactory catalytic oxidation characteristics in the chemical reaction process, which can produce OH to induce and promote a chain reaction of free radicals and ultimately achieve the purpose of degradation and treatment of organic wastewater [82]. Lv et al. found that ultrasonic vibration could achieve an effective approach that could dramatically improve the degradation of methylene blue using industrial Fe78Si9B13 amorphous powders [83]. Fig. S7 (Supporting information) illustrates the pathway of the degradation process. Metallic iron from Fe78Si9B13 amorphous powders reacts with H2O2 in the acidic methylene blue solution and generates the strong oxidizing OH. Then, methylene blue molecules undergo redox reaction and are decomposed into small molecules including H2O and CO2 by cleavage of interlinkage. The work indicates that the micro-channels on the surface of or inside Fe78Si9B13 particles could offer a shortcut for fast mass transfer and supply plenty of reactive sites with low density and high energy due to the structural regeneration after ultrasonic vibration. In addition, amorphous structures also undergo a severe plastic deformation during ultrasonic vibration, which results in the large residual stress inside each particle. All of these factors contribute to the highly effective reaction of metallic iron with H2O2 that occurs on the surface of the loose-packed high-energy powders.

    4.1.3   Combined with electricity process

    A strong electron transfer process could mainly characterize an electrochemical degradation. Qin et al. compared Fe78Si9B13 ribbons with dimensionally stable anode and metal-like boron-doped diamond electrodes in azo dye degradation and found that the Fe78Si9B13 was superior in saving degradation time and energy consumption and improving degradation efficiency [84]. Deng et al. used the Cu55Zr45 amorphous ribbons as an electrode to degrade acid orange Ⅱ dyes and reached 95.6% within 40 min for degradation [85]. The degradation can be divided into the following three ways such as direct electron transfer on the surface of the ribbons, the effect of nano-copper on ribbons surface, and the action of active chlorine (Fig. S8 in Supporting information). The primary way of degrading dye molecules is to produce strongly oxidizing activated radicals by copper nanoparticles and oxidized copper at the ribbon surface under acidic conditions. The process is when Cu0/Cu+ loses electrons and converts to Cu+/Cu2+, and it can react with O2 and H2O to generate OH and O2 under acidic conditions. The report demonstrated that the outstanding properties were derived from the unique atomic structure of amorphous alloys and the application of electrochemical technologies. The as-developed method combining amorphous alloys with electrochemical technology could significantly affect the applicability of this material. In addition, amorphous alloys could also be combined with electro-Fenton process. Compared to the traditional Fenton processes, the electro-Fenton process could present the advantages of reducing H2O2 and power consumption, electrode mass loss and electrolysis time [86]. During the electro-Fenton treatment of rhodamine B using Fe83Si5B8P4, the in-situ production of H2O2 by 2e reduction of dissolved oxygen at the cathode produced hydroxyl radicals with Fe2+ improved the degradation efficiency [80]. Moreover, the Fe2+ was regenerated by electroreduction of Fe3+ on the surface of the cathode, deserving or even avoiding the production of iron sludge [87]. These results suggest that amorphous alloys combined with electricity process could effectively improve the degradation performance in nearly actual wastewater.

    4.1.4   Combined with light process

    Amorphous alloys were proven to be very effective catalysts in the photodegradation process for wastewater treatment [12,68]. Jia et al. reported that Fe-based amorphous alloys exhibited the as-contained advanced catalytic capability when degrading methyl blue and methyl orange in a photo-Fenton oxidation [67]. The production rate of OH was 5–10 times faster than other Fe-based catalysts. The as-employed UV–vis light energy during the dye degradation significantly enhanced the activation of the electrons on 4s2 orbital of the amorphous Fe atoms. In addition, the light energy was also advantageous to Fe3+ to convert back to Fe2+ when providing more catalyst dosage. The inclusion of Si and B atoms in the amorphous alloys enhances the surface stability thereby improving the reusability [66]. Zuo et al. successfully synthesized a series of Fe-B-C-Ti amorphous ribbons to degrade methylene blue under light irradiation [88]. The Fe-B-C-Ti could be activated under visible light, contributing significantly to methylene blue degradation from acidic to neutral pH values. Complete degradation of methylene blue was achieved within 60 min at pH 5 under simulated solar irradiation (Fig. S9 in Supporting information). They also investigated comparing methylene blue degradation with and without adding H2O2. It was found that the appropriate addition of H2O2 can effectively improve the photocatalytic degradation rate. At the same time, the Fe75B10C10Ti5 ribbons were able to degrade the methylene blue by 100% within 6 min at pH 3, even without adding H2O2. Therefore, the Fe75B10C10Ti5 ribbons were effective in pollutant degradation without any addition of oxidant in the acidic environment, which is an excellent property of this amorphous alloy catalyst.

    4.1.5   Mechanism

    According to the above processes of wastewater treatment using amorphous alloys, it is accepted that the oxidative degradation of pollutants in wastewater is mostly achieved by activating peroxides or persulfate using amorphous alloys to produce reactive radicals, such as OH and SO4•‒. Numerous publications dealt with the reactive species generated by amorphous alloys and their roles in wastewater degradation. Recent reports have demonstrated that amorphous alloys could have an excellent ability to activate peroxides. Liang et al. manufactured Fe78Si9B13 amorphous alloy ribbons and compared their activation behavior on three peroxides, namely H2O2, PS and PMS [89]. The results showed that Fe78Si9B13 amorphous alloys had an exceptionally high capability for activating these peroxides to produce OH and/or SO4•‒ (Fig. 3). The reactive radicals of H2O2 were demonstrated as OH, while the PS and PMS activation mainly generated SO4•‒. As shown in Eq. 2, the transitional metals and the generated metal ions could act as electron donors to activate the peroxides using amorphous alloys. (3), (4), (5) present that during the H2O2 activation, the OH as the main oxidative agent, is finally produced to degrade organics. The as-proposed reactions of PS and PMS by amorphous alloys suggested that both OH and SO4•‒ could be generated from PS (Eqs. 6 and 7) and PMS ((8), (9), (10), (11)). The order of predominant radical generation rate by Fe78Si9B13 activation under UV–vis irradiation was PS > H2O2 > PMS.

    (2)

    (3)

    (4)

    (5)

    (6)

    (7)

    (8)

    (9)

    (10)

    (11)

    Figure 3

    Figure 3.  The generation of radicals from H2O2, PS and PMS by amorphous alloys. Reprinted with permission [89]. Copyright 2017, Elsevier B.V.

    Surface stability and catalytic reusability of amorphous alloys have always been important topics in practical applications. It was reported that Co65Mo15B20 exhibited excellent reusability in the treatment of blue 6. The degradation efficiency achieved 97% after 20 times of use due to the synergistic coordination of Co and Mo bimetals and the self-exfoliation effect of the corrosion products [90]. Similar results, such as 35 times of Fe83Si2B11P3C1 [41] in degradation of rhodamine B, 20 times and 19 times reusability of Fe81Cu1Si2B10P6 [79] and Fe80P13C7 [72] in methylene blue degradation, were also achieved. Jia et al. developed Fe78Si9B13 alloys that could be reused up to 30 times while maintaining an acceptable methylene blue degradation rate [68]. This was attributed to the production of SiO2 layer to protect the buried Fe (Fig. 4).

    Figure 4

    Figure 4.  SEM micrographs of Fe78Si9B13 amorphous alloys of (a) initial use; (b) 5th use; (c) 10th use; (d) 20th use; (e) 30th use. Reprinted with permission [68]. Copyright 2016, The Author(s).

    The comparable results of degradation capability versus reusability for various ion states, as well as for amorphous and crystalline Fe-based catalysts, are summarized in Fig. 5. The ion-state Fe-based catalysts with restricted reusability and the produced Fe sludge secondary pollution have become an impediment to their rapid development. Comparatively, the zero-valence irons and Fe-based oxides with the superiorities of low cost, high efficiency, and large surface area. However, the stability and efficiency of these crystalline catalysts is constrained due to the limitations of their structural defects. The reported reusability of crystalline Fe-based Fenton catalysts is within 10 times. Compared with the ion-state and crystalline alternatives, the recent Fe-based amorphous alloy catalysts present higher essential treating ability and more enhanced stability when degrading organic pollutants. Jia et al. reported the Fe83Si2B11P3C1 amorphous ribbon belongs to the greatest performance group, with both ultrahigh essential treating ability and a reusability of 35 times without efficiency decay [41].

    Figure 5

    Figure 5.  Comparison of degradation capability versus reusability for various amorphous and crystalline catalysts. Reprinted with permission [41]. Copyright 2019, Wiley‐VCH Verlag GmbH & Co. KGaA, Weinheim.

    Compared to alloy-based catalysts with a partially crystalline phase, one of the most important properties of amorphous alloys is that the chemical activity could be significantly improved [91]. As it is easier to activate the electrons around the randomly disordered atoms with a weak atomic bonding structure, the high electron mobility in the catalysts will greatly promote the efficiency of wastewater treatment. Table 3 [30,31,63,64,66,71,75,78,90] summarizes the current development of different wastewater using amorphous alloys as catalysts and the advantages compared to the crystalline counterparts or other forms of catalysts, demonstrating the superiorities of using amorphous alloy catalysts in practical applications.

    Table 3

    Table 3.  Catalytic advantages compared to the crystalline counterparts.
    DownLoad: CSV

    It is worth mentioning that examining the reaction activation energy (Ea) of these alloys in different reaction temperatures could provide important insights into explaining catalytic performance in wastewater treatment. Eq. 12 is the Arrhenius equation employed to demonstrate the Ea for wastewater treatment using amorphous alloys.

    (12)

    where k is the kinetic rate at different reaction temperatures (T), Ea is the activation energy, R is the gas constant, and A is a constant.

    The calculated Ea value could show a significant effect on revealing the potential advantages of amorphous alloy catalysts. Fig. 6 and Table 4 [30,41,42,63,66,70,74,76-78,92-98] summarize the as-calculated Ea values of using amorphous alloys with various atomic components compared to the ordinary crystalline catalysts. One of the most important advantages of amorphous alloys is that they present a lower Ea value during wastewater treatment than other catalysts. It was found that the calculated Ea value of using amorphous alloys as catalysts presents a much low value, normally less than 60 kJ/mol, presenting a much lower value than the thermal reactions using ordinary crystalline catalysts (60–250 kJ/mol) [99]. The low value of Ea indicates that less energy is required to pass the reaction barrier, thus making the electrons easier to activate in regard to wastewater treatment.

    Figure 6

    Figure 6.  Comparable results of Ea between the amorphous alloy catalysts and crystalline catalysts. Reprinted with permission [33]. Copyright 2019, Elsevier Ltd.

    Table 4

    Table 4.  Comparison of Ea between amorphous alloys and ordinary crystalline catalysts in wastewater treatment.
    DownLoad: CSV

    The present review has summarized the recent trends in catalytic applications of amorphous alloys as catalysts in wastewater treatment. Amorphous alloys could exhibit several enhancements in oxidative or reductive processes for removing organic pollutants from industrial wastewater. This could be assigned to the unique intrinsic characteristics in contrast with the crystalline catalysts, such as the disordered atomic structure, various compositions, unique electronic structure and atomic configuration. As such, amorphous alloys have tremendous potential in wastewater treatment applications.

    The development of novel amorphous alloys with favorable catalytic activity and stability could be an emerging issue. First, the fine-tuning of electronic structure in amorphous alloys could make the atoms randomly coordinate with each other in short range, providing more unsaturated active sites in catalysis. However, a challenge for optimizing further amorphous alloy catalysts is how to improve the electron transfer ability. Second, elemental distribution on amorphous alloys surface could be significant indicator of material stability and reusability. Including specific elements to improve amorphous alloy surface stability was achievable. Therefore, the element introduction and composition regulation of amorphous alloys are important for equilibrium activity and stability. Moreover, the crystallization behavior of amorphous alloys for the wastewater treatment is recently also attractive to investigate their catalytic application. Some methods, such as electrochemical etching, can construct amorphous-nanocrystalline structures and maintain the catalytic activity.

    Currently, as a new type of environmental remediation catalyst, the interaction mechanism between amorphous alloys and pollutants is still not clear. It is important to verify the catalytic mechanism of amorphous alloys in different systems and in different pollutant reaction processes. In addition, the application of amorphous alloys in practical water treatment is an important proposition. Due to the special characteristics of amorphous alloys, such as low cost, high surface stability, simple operation process and easy modification of atomic composition, it is reasonable to believe that amorphous alloy catalysts are expected to be valuable in practical application in the future. However, to promote the development of amorphous alloy catalysts, several technical restrictions ranging from the composition's design to reaction optimization of amorphous alloys also need to be overcome.

    (1) Reducing the cost and enhancing the reaction efficiency by optimizing the process parameters and improving the catalyst composition with/or surface modification. Moreover, it should focus on establishing the relationship between microstructure and catalytic parameters to reveal the mechanism of performance of these alloys.

    (2) Establishing the theoretical relationship between the electronic structure and catalytic performance to improve the electron transport in the moderate distance and provide the possibility for improving the catalytic performance by regulating the electronic structure.

    (3) Establishing the relationship between the corrosion failure mechanism and surface atomic structure for improving the sustainability of amorphous alloy catalysts, searching for the new atomic composition of amorphous alloys and constructing a suitable atomic structure for catalytic applications.

    (4) The difficulty of composition design of amorphous alloys comes from the uncertainty of amorphous structure. Therefore, adopting the machine learning algorithm to predict the structure and catalytic properties of amorphous alloys and to realize the preparation of amorphous alloys with excellent catalytic properties.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    The authors are grateful to National Natural Science Foundation of China (No. 51672028), National Water Project of China (No. 2018ZX07105–001) for financial support. We also thank editors and reviewers for their constructive suggestions and comments on this paper.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109492.


    1. [1]

      F.L. Dong, Z. Pang, S.Y. Yang, et al., ACS Nano 16 (2022) 3449–3475. doi: 10.1021/acsnano.1c10755

    2. [2]

      Y.W. Fei, N. Han, M.H. Zhang, et al., Chemosphere 307 (2022) 135718. doi: 10.1016/j.chemosphere.2022.135718

    3. [3]

      S. Lim, J.L. Shi, U. von Gunten, D.L. McCurry, Water Res. 213 (2022) 118053. doi: 10.1016/j.watres.2022.118053

    4. [4]

      K.X. Yin, R.X. Wu, Y.N. Shang, et al., Appl. Catal. B Environ. 329 (2023) 122558. doi: 10.1016/j.apcatb.2023.122558

    5. [5]

      Z.C. Shangguan, X.Z. Yuan, L.B. Jiang, et al., Chin. Chem. Lett. 33 (2022) 4719–4731. doi: 10.1016/j.cclet.2022.01.001

    6. [6]

      Y.Y. Zhou, M.C. Cui, Y.M. Ren, et al., Chemosphere 306 (2022) 135547. doi: 10.1016/j.chemosphere.2022.135547

    7. [7]

      K. GracePavithra, P. Senthil Kumar, V. Jaikumar, P. SundarRajan, Rev. Environ. Sci. Biotechnol. 19 (2020) 873–896. doi: 10.1007/s11157-020-09550-0

    8. [8]

      M.X. Yang, Z.X. Hou, X. Zhang, et al., Environ. Sci. Technol. 56 (2022) 11635–11645. doi: 10.1021/acs.est.2c01261

    9. [9]

      R.D. Su, N. Li, Z. Liu, et al., Environ. Sci. Technol. 57 (2023) 1882–1893. doi: 10.1021/acs.est.2c04903

    10. [10]

      Y.N. Shang, X.N. Liu, Y.W. Li, et al., Chem. Eng. J. 446 (2022) 137120. doi: 10.1016/j.cej.2022.137120

    11. [11]

      X. Liang, N.H. Fu, S.C. Yao, Z. Li, Y.D. Li, J. Am. Chem. Soc. 144 (2022) 18155–18174. doi: 10.1021/jacs.1c12642

    12. [12]

      S. Liang, Z. Jia, W.C. Zhang, W.M. Wang, L.C. Zhang, Mater. Des. 119 (2017) 244–253. doi: 10.1016/j.matdes.2017.01.039

    13. [13]

      D.S. Song, J.H. Kim, E. Fleury, W.T. Kim, D.H. Kim, J. Alloys Compd. 389 (2005) 159–164. doi: 10.1016/j.jallcom.2004.08.014

    14. [14]

      J.L. Wu, Z. Peng, Appl. Phys. A 124 (2018) 632. doi: 10.1007/s00339-018-2057-8

    15. [15]

      X.L. Zhang, J.L. Sun, J. Luo, B.B. Wang, J.L. Cheng, Mater. Sci. Technol. 33 (2017) 1186–1191. doi: 10.1080/02670836.2016.1271934

    16. [16]

      H.B. Lu, L.C. Zhang, A. Gebert, L. Schultz, J. Alloys Compd. 462 (2008) 60–67. doi: 10.1016/j.jallcom.2007.08.023

    17. [17]

      J. Henao, A. Concustell, I.G. Cano, et al., Mater. Des. 94 (2016) 253–261. doi: 10.1016/j.matdes.2016.01.040

    18. [18]

      B. Zhang, R.J. Wang, D.Q. Zhao, M.X. Pan, W.H. Wang, Phys. Rev. B 70 (2004) 224208. doi: 10.1103/PhysRevB.70.224208

    19. [19]

      J.L. Zhang, Y.M. Wang, C.H. Shek, J. Appl. Phys. 110 (2011) 083919. doi: 10.1063/1.3650244

    20. [20]

      L. Xia, M.B. Tang, H. Xu, et al., J. Mater. Res. 19 (2004) 1307–1310. doi: 10.1557/JMR.2004.0172

    21. [21]

      S.T. Rajan, A.K. Nandakumar, T. Hanawa, B. Subramanian, J. Non Cryst. Solids 461 (2017) 104–112. doi: 10.1016/j.jnoncrysol.2017.01.008

    22. [22]

      W. Klement, R.H. Willens, P.O.L. Duwez, Nature 187 (1960) 869–870. doi: 10.1038/187869b0

    23. [23]

      Y.C. Hu, Y.W. Li, Y. Yang, et al., Proc. Natl. Acad. Sci. U. S. A. 115 (2018) 6375–6380. doi: 10.1073/pnas.1802300115

    24. [24]

      D.C. Hofmann, S.N. Roberts, NPJ Microgravity 1 (2015) 15003. doi: 10.1038/npjmgrav.2015.3

    25. [25]

      M. Fan, A. Nawano, J. Schroers, M.D. Shattuck, C.S. O'Hern, J. Chem. Phys. 151 (2019) 144506. doi: 10.1063/1.5116895

    26. [26]

      S.T. Rajan, A. Arockiarajan, J. Alloys Compd. 876 (2021) 159939. doi: 10.1016/j.jallcom.2021.159939

    27. [27]

      Q. Halim, N.A.N. Mohamed, M.R.T. Rejab, W.N.W.A. Naim, Q. Ma, Int. J. Adv. Manuf. Technol. 112 (2021) 1231–1258. doi: 10.1007/s00170-020-06515-z

    28. [28]

      S.X. Liang, L.C. Zhang, Mater. Sci. Forum 960 (2019) 200–206. doi: 10.4028/www.scientific.net/msf.960.200

    29. [29]

      A. Molnar, Appl. Surf. Sci. 257 (2011) 8151–8164. doi: 10.1016/j.apsusc.2010.12.046

    30. [30]

      S.H. Xie, P. Huang, J.J. Kruzic, X.R. Zeng, H.X. Qian, Sci. Rep. 6 (2016) 21947. doi: 10.1038/srep21947

    31. [31]

      J.Q. Wang, Y.H. Liu, M.W. Chen, et al., Adv. Funct. Mater. 22 (2012) 2567–2570. doi: 10.1002/adfm.201103015

    32. [32]

      Y.N. Shang, Y.J. Kan, X. Xu, Chin. Chem. Lett. 34 (2023) 108278. doi: 10.1016/j.cclet.2023.108278

    33. [33]

      L.C. Zhang, Z. Jia, F.C. Lyu, S.X. Liang, J. Lu, Prog. Mater Sci. 105 (2019) 100576. doi: 10.1016/j.pmatsci.2019.100576

    34. [34]

      S.X. Liang, Z. Jia, Y.J. Liu, et al., Adv. Mater. 30 (2018) 1802764. doi: 10.1002/adma.201802764

    35. [35]

      X.Y. Li, W.Z. Cai, D.S. Li, et al., Nano Res. 16 (2021) 4277–4288. doi: 10.1039/d1se00919b

    36. [36]

      X. Han, G. Wu, Y.Y. Ge, et al., Adv. Mater. 34 (2022) 2206994. doi: 10.1002/adma.202206994

    37. [37]

      J.Z. Xie, S. Ewing, J.N. Boyn, et al., Nature 611 (2022) 479–484. doi: 10.1038/s41586-022-05261-4

    38. [38]

      Y. Pei, G.B. Zhou, L. Nguyen, et al., Chem. Soc. Rev. 41 (2012) 8140–8162. doi: 10.1039/c2cs35182j

    39. [39]

      Á. Molnáar, G.V. Smith, M. Bartók, Adv. Catal. 36 (1989) 329–383.

    40. [40]

      C.J. Byrne, M. Eldrup, Science 321 (2008) 502–503. doi: 10.1126/science.1158864

    41. [41]

      Z. Jia, Q. Wang, L.G. Sun, et al., Adv. Funct. Mater. 29 (2019) 1807857. doi: 10.1002/adfm.201807857

    42. [42]

      C.Q. Zhang, Z.W. Zhu, H.F. Zhang, J. Phys. Chem. Solids 110 (2017) 152–160. doi: 10.1016/j.jpcs.2017.06.010

    43. [43]

      S.H. Xie, Y.L. Xie, J.J. Kruzic, et al., ChemCatChem 12 (2020) 750–761. doi: 10.1002/cctc.201901623

    44. [44]

      A. Studer, D.P. Curran, Nat. Chem. 6 (2014) 765–773. doi: 10.1038/nchem.2031

    45. [45]

      F. Hu, S.L. Zhu, S.M. Chen, et al., Adv. Mater. 29 (2017) 1606570. doi: 10.1002/adma.201606570

    46. [46]

      Y.C. Hu, Y.Z. Wang, R. Su, et al., Adv. Mater. 28 (2016) 10293–10297. doi: 10.1002/adma.201603880

    47. [47]

      C. Yang, C. Zhang, Z.J. Chen, et al., ACS Appl. Mater. Interfaces 13 (2021) 7227–7237. doi: 10.1021/acsami.0c20832

    48. [48]

      W.H. Wang, J. Appl. Phys. 110 (2011) 053521. doi: 10.1063/1.3632972

    49. [49]

      C. Chang, H.P. Zhang, R. Zhao, et al., Nat. Mater. 21 (2022) 1240–1245. doi: 10.1038/s41563-022-01327-w

    50. [50]

      Y.Q. Yan, X. Liang, J. Ma, J. Shen, Intermetallics 124 (2020) 106849. doi: 10.1016/j.intermet.2020.106849

    51. [51]

      S.X. Liang, W.C. Zhang, L.N. Zhang, W.M. Wang, L.C. Zhang, Sustain. Mater. Technol. 22 (2019) e00126.

    52. [52]

      G.F. Lv, D.C. Wu, R.W. Fu, J. Hazard. Mater. 165 (2009) 961–966. doi: 10.1016/j.jhazmat.2008.10.090

    53. [53]

      Z. Deng, C. Zhang, L. Liu, Intermetallics 52 (2014) 9–14. doi: 10.1016/j.intermet.2014.03.007

    54. [54]

      P. Wang, X.F. Bian, Y.X. Li, Chin. Sci. Bull. 57 (2012) 33–40. doi: 10.1007/s11434-011-4876-2

    55. [55]

      L. Wan, C. Li, G. Long, et al., Colloids Surf. A Physicochem. Eng. Asp. 645 (2022) 128924. doi: 10.1016/j.colsurfa.2022.128924

    56. [56]

      X.C. Xu, Z.L. Shi, K.Q. Qiu, Mater. Res. Express 7 (2020) 045205. doi: 10.1088/2053-1591/ab89df

    57. [57]

      H.Z. Zhao, Y. Sun, L.N. Xu, J.R. Ni, Chemosphere 78 (2010) 46–51. doi: 10.1016/j.chemosphere.2009.10.034

    58. [58]

      N.R. Neti, R. Misra, Chem. Eng. J. 184 (2012) 23–32. doi: 10.1016/j.cej.2011.12.014

    59. [59]

      P. Pal, R. Kumar, Sep. Purif. Rev. 43 (2014) 89–123. doi: 10.1080/15422119.2012.717161

    60. [60]

      X.D. Qin, Z.K. Li, Z.W. Zhu, D.W. Fang, H.F. Zhang, J. Phys. Chem. Solids 133 (2019) 85–91. doi: 10.1016/j.jpcs.2019.05.017

    61. [61]

      J.F. Yang, X.F. Bian, Y.W. Bai, X.Q. Lv, P. Wang, J. Non Cryst. Solids 358 (2012) 2571–2574. doi: 10.1016/j.jnoncrysol.2012.06.002

    62. [62]

      X.Y. Li, R.L. Lv, W.M. Zhang, et al., Water Res. 228 (2023) 119363. doi: 10.1016/j.watres.2022.119363

    63. [63]

      J.Q. Wang, Y.H. Liu, M.W. Chen, et al., Sci. Rep. 2 (2012) 418. doi: 10.1038/srep00418

    64. [64]

      Z. Jia, X.G. Duan, P. Qin, et al., Adv. Funct. Mater. 27 (2017) 1702258. doi: 10.1002/adfm.201702258

    65. [65]

      B.W. Zhao, Y.L. Liu, H. Zhang, et al., Sustain. Mater. Technol. 35 (2023) e00539.

    66. [66]

      Y. Tang, Y. Shao, N. Chen, K.F. Yao, RSC Adv. 5 (2015) 6215–6221. doi: 10.1039/C4RA10000J

    67. [67]

      Z. Jia, J. Kang, W.C. Zhang, et al., Appl. Catal. B Environ. 204 (2017) 537–547. doi: 10.1016/j.apcatb.2016.12.001

    68. [68]

      Z. Jia, X.G. Duan, W.C. Zhang, et al., Sci. Rep. 6 (2016) 38520. doi: 10.1038/srep38520

    69. [69]

      Z. Jia, S.X. Liang, W.C. Zhang, et al., J. Taiwan Inst. Chem. Eng. 71 (2017) 128–136. doi: 10.1016/j.jtice.2016.11.021

    70. [70]

      X.F. Wang, Y. Pan, Z.R. Zhu, J.L. Wu, Chemosphere 117 (2014) 638–643. doi: 10.1016/j.chemosphere.2014.09.055

    71. [71]

      C.Q. Zhang, Z.W. Zhu, H.F. Zhang, Z.Q. Hu, J. Environ. Sci. 24 (2012) 1021–1026. doi: 10.1016/S1001-0742(11)60894-2

    72. [72]

      Q.Q. Wang, M.X. Chen, P.H. Lin, et al., J. Mater. Chem. A 6 (2018) 10686–10699. doi: 10.1039/c8ta01534a

    73. [73]

      C.Q. Zhang, H.F. Zhang, M.Q. Lv, Z.Q. Hu, J. Non Cryst. Solids 356 (2010) 1703–1706. doi: 10.1016/j.jnoncrysol.2010.06.019

    74. [74]

      C.Q. Zhang, Z.W. Zhu, H.F. Zhang, Z.Q. Hu, J. Non Cryst. Solids 358 (2012) 61–64. doi: 10.1016/j.jnoncrysol.2011.08.023

    75. [75]

      C.Q. Zhang, Q.L. Sun, J. Non Cryst. Solids 470 (2017) 93–98. doi: 10.1016/j.jnoncrysol.2017.05.009

    76. [76]

      N. Weng, F. Wang, F.X. Qin, W.Y. Tang, Z.H. Dan, Materials 10 (2017) 1001. doi: 10.3390/ma10091001

    77. [77]

      J.C. Wang, Z. Jia, S.X. Liang, et al., Mater. Des. 140 (2018) 73–84. doi: 10.14257/ijfgcn.2018.11.5.07

    78. [78]

      S.Q. Chen, G.N. Yang, S.T. Luo, et al., J. Mater. Chem. A 5 (2017) 14230–14240. doi: 10.1039/C7TA01206C

    79. [79]

      Q.Q. Wang, L. Yun, M.X. Chen, et al., ACS Appl. Nano Mater. 2 (2019) 214–227. doi: 10.1021/acsanm.8b01669

    80. [80]

      M.Q. Zuo, S. Yi, J. Choi, J. Environ. Sci. 105 (2021) 116–127. doi: 10.1016/j.jes.2020.12.032

    81. [81]

      B.W. Zhao, Z.W. Zhu, X.D. Qin, Z.K. Li, H.F. Zhang, J. Mater. Sci. Technol. 46 (2020) 88–97. doi: 10.1016/j.jmst.2019.12.012

    82. [82]

      Y.Y. Cai, J.Y. Li, G.F. Qu, et al., Water Environ. Res. 93 (2021) 1243–1253. doi: 10.1002/wer.1533

    83. [83]

      Z.W. Lv, Y.Q. Yan, C.C. Yuan, et al., Mater. Des. 194 (2020) 108876. doi: 10.1016/j.matdes.2020.108876

    84. [84]

      X.D. Qin, Z.K. Li, Z.W. Zhu, et al., J. Mater. Sci. Technol. 34 (2018) 2290–2296. doi: 10.1016/j.jmst.2018.04.012

    85. [85]

      Z.W. Deng, B.W. Zhao, S.T. Li, et al., J. Environ. Sci. 136 (2022) 537–546. doi: 10.1109/tec.2021.3092786

    86. [86]

      K.S. Aneeshkumar, J.S. Tian, J. Shen, Chin. Chem. Lett. 33 (2022) 2327–2344. doi: 10.1016/j.cclet.2021.12.013

    87. [87]

      V. Poza-Nogueiras, E. Rosales, M. Pazos, M.A. Sanroman, Chemosphere 201 (2018) 399–416. doi: 10.1016/j.chemosphere.2018.03.002

    88. [88]

      M.Q. Zuo, M. Moztahida, D.S. Lee, S. Yi, Chemosphere 287 (2022) 132175. doi: 10.1016/j.chemosphere.2021.132175

    89. [89]

      S.X. Liang, Z. Jia, W.C. Zhang, et al., Appl. Catal. B: Environ. 221 (2018) 108–118. doi: 10.1016/j.apcatb.2017.09.007

    90. [90]

      M.F. Tang, L.M. Lai, D.Y. Ding, et al., J. Non-Cryst. Solids 576 (2022) 121282. doi: 10.1016/j.jnoncrysol.2021.121282

    91. [91]

      Q. Chen, Z.G. Qi, Y. Feng, et al., J. Mol. Liq. 364 (2022) 120058. doi: 10.1016/j.molliq.2022.120058

    92. [92]

      F. Miao, Q.Q. Wang, Q.S. Zeng, et al., J. Mater. Sci. Technol. 38 (2020) 107–118. doi: 10.1016/j.jmst.2019.07.050

    93. [93]

      X.Q. Wang, Q.Y. Zhang, S.X. Liang, et al., Catalysts 10 (2020) 48. doi: 10.3390/catal10010048

    94. [94]

      E. Saputra, S. Muhammad, H. Sun, et al., J. Colloid Interface Sci. 407 (2013) 467–473. doi: 10.1016/j.jcis.2013.06.061

    95. [95]

      P. Shukla, H. Sun, S. Wang, H.M. Ang, M.O. Tadé, Sep. Purif. Technol. 77 (2011) 230–236. doi: 10.1016/j.seppur.2010.12.011

    96. [96]

      P. Shukla, S. Wang, K. Singh, H.M. Ang, M.O. Tadé, Appl. Catal. B: Environ. 99 (2010) 163–169. doi: 10.1016/j.apcatb.2010.06.013

    97. [97]

      P. Shukla, H. Sun, S. Wang, H.M. Ang, M.O. Tadé, Catal. Today 175 (2011) 380–385. doi: 10.1016/j.cattod.2011.03.005

    98. [98]

      Y. Yang, H. Zhang, Y. Yan, R. Soc. Open Sci. 5 (2018) 171731. doi: 10.1098/rsos.171731

    99. [99]

      J.X. Chen, L.Z. Zhu, Catal. Today 126 (2007) 463–470. doi: 10.1016/j.cattod.2007.06.022

  • Figure 1  Schematic illustration of the different degradation mechanisms for B element facilitated Fe-based amorphous alloy catalysts. Reprinted with permission [43]. Copyright 2019, Wiley‐VCH Verlag GmbH & Co. KGaA, Weinheim.

    Figure 2  Atomic configuration of Fe3P and amorphous alloy components calculated by DFT. Reprinted with permission [47]. Copyright 2021, American Chemical Society.

    Figure 3  The generation of radicals from H2O2, PS and PMS by amorphous alloys. Reprinted with permission [89]. Copyright 2017, Elsevier B.V.

    Figure 4  SEM micrographs of Fe78Si9B13 amorphous alloys of (a) initial use; (b) 5th use; (c) 10th use; (d) 20th use; (e) 30th use. Reprinted with permission [68]. Copyright 2016, The Author(s).

    Figure 5  Comparison of degradation capability versus reusability for various amorphous and crystalline catalysts. Reprinted with permission [41]. Copyright 2019, Wiley‐VCH Verlag GmbH & Co. KGaA, Weinheim.

    Figure 6  Comparable results of Ea between the amorphous alloy catalysts and crystalline catalysts. Reprinted with permission [33]. Copyright 2019, Elsevier Ltd.

    Table 1.  Properties of common amorphous alloys.

    下载: 导出CSV

    Table 2.  The kobs of several amorphous alloys in wastewater degradation.

    下载: 导出CSV

    Table 3.  Catalytic advantages compared to the crystalline counterparts.

    下载: 导出CSV

    Table 4.  Comparison of Ea between amorphous alloys and ordinary crystalline catalysts in wastewater treatment.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  54
  • HTML全文浏览量:  5
文章相关
  • 发布日期:  2024-10-15
  • 收稿日期:  2023-09-24
  • 接受日期:  2024-01-02
  • 修回日期:  2023-12-09
  • 网络出版日期:  2024-01-05
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章