Next Article in Journal
The Dual Influence of Silicon Content and Mechanical Stress on Magnetic Barkhausen Noise in Non-Oriented Electrical Steel
Previous Article in Journal
Atomic Revealing of the Dissolution Behavior of Spinel Oxides on the 316L Surface in Alkaline High-Temperature and High-Pressure Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Synthesis of Iron-Based and Aluminum-Based Bimetals: A Systematic Review

by
Jeffrey Ken B. Balangao
1,2,3,*,
Carlito Baltazar Tabelin
1,2,
Theerayut Phengsaart
4,
Joshua B. Zoleta
1,2,
Takahiko Arima
5,
Ilhwan Park
5,
Walubita Mufalo
6,
Mayumi Ito
5,
Richard D. Alorro
7,
Aileen H. Orbecido
8,
Arnel B. Beltran
8,
Michael Angelo B. Promentilla
8,
Sanghee Jeon
9,
Kazutoshi Haga
9 and
Vannie Joy T. Resabal
1,2
1
Resource Processing and Technology Center, RIEIT, Mindanao State University-Iligan Institute of Technology, Iligan City 9200, Philippines
2
Department of Materials and Resources Engineering & Technology, College of Engineering, Mindanao State University-Iligan Institute of Technology, Iligan City 9200, Philippines
3
College of Technology, University of Science and Technology of Southern Philippines, Cagayan de Oro City 9000, Philippines
4
Department of Mining and Petroleum Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
5
Division of Sustainable Resources Engineering, Faculty of Engineering, Hokkaido University, Sapporo 060-8628, Japan
6
Department of Materials Chemistry, National Institute of Technology, Asahikawa College, Asahikawa 071-8142, Japan
7
Western Australian School of Mines: Minerals, Energy and Chemical Engineering, Curtin University, Bentley, WA 6102, Australia
8
Department of Chemical Engineering, De La Salle University, Manila 0922, Philippines
9
Graduate School of International Resource Sciences, Akita University, Akita 010-0865, Japan
*
Author to whom correspondence should be addressed.
Metals 2025, 15(6), 603; https://doi.org/10.3390/met15060603
Submission received: 30 March 2025 / Revised: 19 May 2025 / Accepted: 21 May 2025 / Published: 27 May 2025
(This article belongs to the Section Extractive Metallurgy)

Abstract

:
Bimetals—materials composed of two metal components with dissimilar standard reduction–oxidation (redox) potentials—offer unique electronic, optical, and catalytic properties, surpassing monometallic systems. These materials exhibit not only the combined attributes of their constituent metals but also new and novel properties arising from their synergy. Although many reviews have explored the synthesis, properties, and applications of bimetallic systems, none have focused exclusively on iron (Fe)- and aluminum (Al)-based bimetals. This systematic review addresses this gap by providing a comprehensive overview of conventional and emerging techniques for Fe-based and Al-based bimetal synthesis. Specifically, this work systematically reviewed recent studies from 2014 to 2023 using the Scopus, Web of Science (WoS), and Google Scholar databases, following the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines, and was registered under INPLASY with the registration number INPLASY202540026. Articles were excluded if they were inaccessible, non-English, review articles, conference papers, book chapters, or not directly related to the synthesis of Fe- or Al-based bimetals. Additionally, a bibliometric analysis was performed to evaluate the research trends on the synthesis of Fe-based and Al-based bimetals. Based on the 122 articles analyzed, Fe-based and Al-based bimetal synthesis methods were classified into three types: (i) physical, (ii) chemical, and (iii) biological techniques. Physical methods include mechanical alloying, radiolysis, sonochemical methods, the electrical explosion of metal wires, and magnetic field-assisted laser ablation in liquid (MF-LAL). In comparison, chemical protocols covered reduction, dealloying, supported particle methods, thermogravimetric methods, seed-mediated growth, galvanic replacement, and electrochemical synthesis. Meanwhile, biological techniques utilized plant extracts, chitosan, alginate, and cellulose-based materials as reducing agents and stabilizers during bimetal synthesis. Research works on the synthesis of Fe-based and Al-based bimetals initially declined but increased in 2018, followed by a stable trend, with 50% of the total studies conducted in the last five years. China led in the number of publications (62.3%), followed by Russia, Australia, and India, while Saudi Arabia had the highest number of citations per document (95). RSC Advances was the most active journal, publishing eight papers from 2014 to 2023, while Applied Catalysis B: Environmental had the highest number of citations per document at 203. Among the three synthesis methods, chemical techniques dominated, particularly supported particles, galvanic replacement, and chemical reduction, while biological and physical methods have started gaining interest. Iron–copper (Fe/Cu), iron–aluminum (Fe/Al), and iron–nickel (Fe/Ni) were the most commonly synthesized bimetals in the last 10 years. Finally, this work was funded by DOST-PCIEERD and DOST-ERDT.

1. Introduction

The efficacy of monometallic materials or zero-valent metals (ZVMs) in treating polluted surface water, wastewater, groundwater, and soils, including municipal, industrial, and mining wastes, is well known. Metallic iron (Fe) and zero-valent iron (ZVI), for example, have been effectively used to degrade halogenated organic compounds (HOCs) like trichloroethylene (TCE), tetrabromobisphenol A, pentachlorophenol (PCP), and 2,4-dichlorophenol, trichlorophenol, as well as residual medicine and antibiotics such as florfenicol and diclofenac [1]. Similarly, ZVI and aluminum (Al) or zero-valent aluminum (ZVAl) have been applied for decontaminating lead (Pb)-contaminated soil [2] and Pb- and zinc (Zn)-bearing hydrometallurgical residues from legacy mining and mineral processing operations [3,4]. ZVMs have also been used in the industrial-scale recovery of gold (Au) from pregnant robbing solutions as well as in the recovery of valuable metals like copper (Cu) and cobalt (Co) from electronic wastes [5,6]. Another emerging application of ZVMs is their potential to passivate pyrite, the primary source of acid mine drainage (AMD), via galvanic encapsulation [7,8].
Despite their widespread use, the efficacy of ZVMs is often hampered by issues like excessive corrosion, rapid passivation, and suboptimal utilization [9,10,11,12]. For example, ZVI and ZVAl exhibit slow removal rates and low efficiency at high pollutant concentrations and neutral-to-alkaline pH conditions because of the formation of a metal oxide or carbonate passivation layer [13,14], reducing their reactivity, so these materials work best under highly acidic conditions near pH 3 [15,16]. ZVMs are also prone to agglomerate, reducing their effective surface area, and are challenging to recover and regenerate when used in contaminated solutions and wastewater [10,17]. Finally, ZVMs exhibit low selectivity for target contaminants, posing limitations to their application [18]. One possible solution to address these drawbacks of ZVMs is by combining two metals into bimetals, a configuration that enhances efficiency by taking advantage of the individual monometallic properties and the synergistic effects of the paired metals [19].
In general, bimetals (also referred to as bimetallic materials and bimetallic particles) are materials composed of two metal components with different standard reduction–oxidation (redox) potentials, which are highly attractive for their unique electronic, optical, and catalytic properties, surpassing their monometallic counterparts [19,20,21]. Bimetals have been shown to form galvanic cells and promote either reduction or oxidation reactions depending on the geochemical properties of the system [19,22]. For example, pairing ZVI with more noble metals, such as copper (Cu), cobalt (Co), nickel (Ni), or silver (Ag), has been shown to enhance its reactivity and performance in wastewater treatment by inhibiting surface passivation [23]. For these bimetal systems, the contaminant removal primarily occurred through galvanic couple formation, which influenced iron (Fe) corrosion and facilitated electron transfer to contaminants. However, the relatively high cost of these noble metals poses challenges for scaling this technology up economically. Aluminum is a promising alternative due to its abundance, lower cost, and significantly lower standard redox potential, enabling a strong thermodynamic force for metal removal when coupled with ZVI. Adding ZVAl to other ZVMs, such as ZVI in Al/Fe bimetallic systems, enhances their reactivity by preventing passive layer formation on ZVI [15] and maintaining its reactivity throughout the reaction [24]. In many studies, bimetallic catalysts have improved performance when applied to polluted water and wastewater treatment [25,26,27] and are better than their monometallic counterparts [19]. In an Al/Fe bimetal system, Al acts as the primary reactive metal or anode, while Fe serves as the more noble secondary metal or cathode that ferries electrons to surface reactive compounds [28].
Because of their unique properties and huge potential as sustainable catalysts and adsorbents for circular economy applications, bimetals are gaining a lot of attention in the research community. To date, a number of review papers have been published discussing synthesis protocols, electrochemical properties of products, and applications of bimetal systems [29,30,31,32,33]. For instance, Scaria et al. [30] and Quiton et al. [31] reported physical, chemical, and biological synthesis methods of bimetallic particles and their applications to water and wastewater treatment. Consequently, none of these review articles delved comprehensively and exclusively into Fe- and Al-based bimetals. Hence, this systematic review aims to provide a thorough overview of conventional and emerging synthesis methods for synthesizing Fe-based and Al-based bimetals. A bibliometric analysis was also performed to evaluate research activity related to the syntheses of Fe- and Al-based bimetals.
This review is structured as follows: Section 2 outlines the review methodology; Section 3 discusses the production of Fe- and Al-based bimetals by physical techniques; Section 4 explores the chemical methods for Al- and Fe-based bimetal synthesis; Section 5 examines the use of plant extracts, chitosan, and cellulose-based materials as green reducing agents and stabilizers for bimetallic nanoparticle synthesis; and Section 6 provides a summary and future directions of bimetal material synthesis.

2. Materials and Methods

This systematic review examined the conventional methods and emerging protocols for Fe- and Al-bimetal synthesis from 2014 to 2023, following the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines 2020 [34], and recommendations by Andrews [35]. It was based on the research question: What were the methods employed for synthesizing Fe-based and Al-based bimetals in the last 10 years (2014–2023)? Peer-reviewed journal publications were identified using keywords such as “bimetallic”, “aluminum”, and “iron” using the Web of Science (WoS) and Scopus databases. The initial search resulted in 344 articles—89 from WoS and 255 from Scopus—with 291 remaining after removing duplicates. Articles were screened based on titles, abstracts, and keywords, excluding 76 articles that did not focus on bimetallic materials. Of the remaining 215 papers, further eligibility evaluation of full-text articles excluded 168 papers due to inaccessibility (3), non-English (3), reviews, conference papers, and book chapters (15), and 147 that were unrelated to Fe-based and Al-based bimetallic material synthesis. The final selection categorized the 47 remaining articles based on their synthesis method as physical (3), chemical (42), or biological (2). Physical methods are further subcategorized: mechanical alloying (1), the electrical explosion of metal wires (1), and sonochemical methods (1). On the other hand, the chemical methods are as follows: chemical reduction (5), chemical dealloying (2), supported particles/nanoparticles (12), thermogravimetric methods (7), seed-mediated growth (1), galvanic replacement (12), and electrochemical synthesis (3).
The PRISMA guidelines allow for the addition of supplementary articles from other databases, such as Google Scholar, that were not captured in the original search to achieve a more comprehensive scope [36]. In total, 75 articles were added to the initial 47 articles using this approach, making a total of 122. These additional articles were selected following the same rigorous inclusion criteria used to screen the 47 articles. The final 122 articles were again categorized as physical (15), chemical (95), or biological (12) (Figure 1). Articles on physical methods were further broken down as mechanical alloying (6), the electrical explosion of metal wires (2), sonochemical methods (5), radiolysis (1), or magnetic field-assisted laser ablation in liquid (MF-LAL) (1). The chemical methods were subcategorized as chemical reduction (16), chemical dealloying (2), supported particles/nanoparticles (41), thermogravimetric methods (7), seed-mediated growth (2), galvanic replacement (23), or electrochemical synthesis (4) (Figure 2). The twelve (12) articles on biological methods involved bio-based reducing agents and stabilizers during synthesis. The protocol was registered via the International Platform of Registered Systematic Review and Meta-analysis Protocols (INPLASY) with the registration number INPLASY202540026. Moreover, a bibliometric analysis was conducted to assess research activity on the synthesis of Fe-based and Al-based bimetals, which will be discussed in the following subsections.

2.1. Bimetal Research Publication Trends

The distribution of research works on the synthesis of Fe-based and Al-based bimetals is shown in Figure 3. There was a decline in research interest in the first four years (2014–2017), and there was perhaps a transition toward the development of application-driven and even complex bimetallic materials which would require longer durations of study. Since 2014, interest in Fe- and Al-based bimetals has grown steadily, with China leading global research. During this year, galvanic replacement has emerged as the dominant synthesis method. However, a wide range of techniques—including supported particles, chemical reduction, mechanical alloying, and newer methods like MF-LAL and dealloying synthesis—have been explored to produce environmentally relevant bimetals such as Fe/Ni, Fe/Cu, and Fe/Pt. In 2015, there was rising interest in supported particle approaches for scalable nanomaterials, while biological synthesis emerged as a green alternative. The year 2016 continued the use of scalable, low-waste methods like galvanic replacement and supported particles, along with the rise of electrochemical synthesis for precision and renewable-aligned processes. Commonly studied bimetals (Fe/Al, Fe/Ni, Fe/Cu) and advanced supports (e.g., Al2O3–MCM-41, mesoporous silica) reflected policy-driven interest in sustainable, high-efficiency materials for catalysis, energy, and environmental remediation. In 2017, the research output had a strong focus on Fe/Ni and Fe/Cu bimetals for catalysis, energy storage, and environmental applications. The year also had increased use of supported materials like Al2O3, activated carbon, and biochar, along with eco-friendly synthesis methods such as electrochemical, radiolysis, and sonochemical techniques to enhance material performance and sustainability.
Suddenly, the trend increased in the fifth year (2018). In 2018, the highest peak in research output was manifested, marked by the rising use of biological synthesis, supported particles, and nanostructured bimetals like Fe-Cu, alongside high-energy methods that enhanced material performance and reflected growing sustainability and innovation priorities. Then, a stable trend manifested in the next four years, before it decreased in 2023. In 2019, research manifested an increased use of clay, polymer, and paper-based supports to enhance bimetallic material reusability and stability. Chemical reduction and galvanic replacement remained prominent, while mechanical alloying gained traction as a solvent-free synthesis method. Fe/Ni, Fe/Cu, and Fe/Al bimetals were widely developed for applications in catalysis, water treatment, and energy storage. In 2020, greener synthesis approaches using biological methods and supported particles became more prevalent, with materials like chitosan and zeolite supporting environmental remediation and nanocomposite development for cost-effective pollutant removal.
In 2021, research increasingly focused on multifunctional, green-synthesized bimetallic materials aligned with global goals for sustainability, waste reduction, and clean energy catalysis. Advanced composites like PdFe on N-doped carbon layers and supported systems such as Fe-Ni/AC and Fe-Cu/CAC demonstrated innovation balanced with environmental responsibility. In 2022, publications showed growing emphasis on hybrid and multifunctional composites (e.g., biochar@Fe/Ni, AlOOH/AlFe), combining adsorption, catalysis, and reusability aspects. The adoption of waste-derived supports and energy-efficient synthesis methods like biological approaches and mechanical alloying underscored alignment with circular economy and green chemistry principles. In 2023, China maintained its research leadership, with increased use of bio-based supports like alginate–limestone (Alg–LS) and nitrogen-doped carbon nanotubes (NCNTs), aligning with bioeconomy and sustainability goals to enhance bimetal functionality in catalysis and wastewater treatment. The development of multifunctional composites such as Fe-Ni/Al2O3 and Fe-Ce/NCNT, along with continued synthesis of Fe/Al, Fe/Zn, and Fe/Cu, reflects a global push for efficient, eco-friendly solutions to address industrial and environmental pollution. Finally, it is notable that the research conducted in the last five years would account for 50% of the total research undertaken in the last ten years.

2.2. Active Countries

Table 1 shows a list of the countries most active in publishing documents on the synthesis of Fe-based and Al-based bimetals. China dominated the list (62.3%), followed by Russia (5.7%), Australia (4.9%), and India (4.9%). Meanwhile, Saudi Arabia ranked first in the number of citations per document (C/D) with 95, followed by Iran (64) and China (45.2).

2.3. Active Journals

RSC Advances was the most active journal with 8 (6.6%) documents, followed by Chemical Engineering Journal and Chemosphere with 7 (5.7%) documents each. The most active journals are primarily in the fields of chemical engineering, chemistry, and environmental science. In terms of the number of citations per document (C/D), Applied Catalysis B: Environmental ranked first in the list with 203, followed by Journal of Hazardous Materials (80.7) and Chemical Engineering Journal (70.4). Notably, the active journals in the list were all in the Q1 ranking (Table 2).

2.4. Active Institutions

The list of active institutions involved in the synthesis of Fe-based and Al-based bimetals is illustrated in Table 3. Tongji University in China is the most active institution, with 7 (5.7%) documents, followed by the Institute of Strength Physics and Materials Science (Russia) with 6 (4.9%) documents. China’s Fujian Normal University and Sichuan University ranked third with 5 (4.1%) documents each. It is worth noting that Chinese institutions were the most dominant in developing bimetals, with 7 out of the listed 11 institutions comprising 30 of the total 122 documents.

2.5. Active Authors

Researchers from China dominated the list of active authors, followed by those from Australia (Table 4). However, A. Sharipova, affiliated with two institutions, Technion–Israel Institute of Technology (Israel) and Institute of Strength Physics and Materials Science (Russia), ranked first in the list of active authors for developing Fe-based and Al-based bimetals. The researcher has primarily authored 4 (3.3%) documents. Xin Liu and Jing Wang (China), J.S. Riva (Argentina), and Naeim Ezzatahmadi (Australia) emerged next in the ranking with 3 (2.4%) documents each.

2.6. Most-Used Synthesis Methods

Figure 4 illustrates the most popular synthesis methods for Fe- and Al-based bimetals. It is worth noting that chemical synthesis methods dominated with supported particles ranking first, which were described by researchers in detail in 41 research documents, followed by galvanic replacement and chemical reduction utilized in 23 and 15 studies, respectively. Different biological methods were used in 12 studies. Furthermore, it is notable that the employment of physical processes such as mechanical alloying and sonochemical techniques started to grow with 6 and 5 research studies, respectively.

2.7. Most-Synthesized Iron-Based and Aluminum-Based Bimetals

The bimetals synthesized by the most researchers were Fe/Cu, Fe/Al, and Fe/Ni, with details of their syntheses found in 20, 16, and 16 research studies, respectively (Figure 5). Researchers developed Fe/Ag bimetals in 5 different studies. Moreover, Fe/Pd, Fe/Pt, Fe/Co, and Fe/Rh were reported in 3 research documents each.

3. Physical Methods for Synthesizing Iron-Based and Aluminum-Based Bimetals

Physical synthesis methods are suitable for large-scale production [30] and typically use little to no solvent [37], with solid raw materials as starting materials [38]. However, these methods are costly and require high energy consumption to maintain extreme reaction conditions [39]. Additionally, they often result in lower production rates [40,41]. Among the physical methods reported in this review are mechanical alloying, the electrical explosion of metal wires, radiolysis, sonochemical methods, and magnetic field-assisted laser ablation in liquid (MF-LAL).

3.1. Mechanical Alloying

Mechanical alloying is a material processing method first reported by John S. Benjamin in 1970, a process involving dry high-energy ball milling without surface-active additives to produce coarse, contamination-free composite powder particles. Bimetals are generated by this method through repeated cold welding and flaking of materials until all constituents are finely divided and uniformly distributed within each particle [42]. Ball milling offers several advantages, including particle deformation, lattice defects, and welding, leading to nanometer-scale refinement [43,44]. It allows the synthesis of new compounds, such as from elemental iron, carbon, or sulfur, without requiring solvents or harsh conditions [45,46,47]. Additionally, it provides remarkable stoichiometric control by enabling the formation of distinct products through reaction mixture composition [48]. Table 5 shows studies that employed mechanical alloying for synthesizing Fe-based bimetallic materials.
Hawili et al. [49] utilized ball milling to prepare magnetic Fe/Cu coated on aluminum substrate, during which Fe50Cu50 was made from iron (purity: 99.9%; particle size: 125 µm) and copper (purity: 99.5%; particle size: 100 µm) powders in a planetary ball mill. Specifically, the synthesis involved dry ball milling at a 6:1 ball-to-powder ratio using 10 mm stainless steel balls, operating at 600 rpm for 6 h with periodic breaks and directional reversal, enabling interdiffusion between Fe and Cu particles to form a solid-state Fe/Cu solution. The coating process involved fixing an aluminum collar (0.1 mm thick, 10 × 30 mm) substrate inside the grinding bowls and depositing Fe/Cu through high-speed ball milling after removing the balls. Meanwhile, Vishlaghi and Ataie [50] developed a magnetic carbon nanotube (CNT)-supported Cu-Fe alloy, which was carried out by CNT additions of 2 wt.%, 5 wt.%, and 10 wt.%, mixing Cu80Fe20 with CNTs and milling for 15 h under argon (Ar) in a planetary ball mill. Mechanical alloying was performed with a ball-to-powder weight ratio of 20:1 and a speed of 300 rpm, using a hardened chromium steel vial and balls.
Attrition mills can also be used in the mechanical alloying process, aside from ball milling [51]. In the previous work of Sharipova et al. [52], Fe/Ag nanopowder blends (95:5 and 90:10 vol.%, referred to as Fe5Ag and Fe10Ag) were prepared for biomedical applications by high-energy attrition milling using carbonyl Fe and Ag2O powders. The milling process was conducted in hexane under an Ar atmosphere with a 20:1 ball-to-powder ratio, using stainless steel balls, durations of 4, 8, and 12 h, and drying under vacuum (10−2 Torr). The optimal combination of strength and ductility was achieved for both compositions after annealing at 550 °C. Another study conducted by Sharipova et al. [53] developed Fe5Ag and Fe10Ag (vol%) materials for biodegradable load-bearing scaffolds by high-energy attrition milling of the carbonyl Fe and Ag2O powders, with a milling duration of 8 h. Fe–Ag nanocomposite scaffolds demonstrated high strength in both compression and bending, along with high ductility, while the 70% and 75% macroporous Fe–Ag scaffolds exhibited a compressive strength and permeability comparable to trabecular bone. These authors also synthesized Fe/Ag and Fe/Cu nanocomposite powders using high-energy attrition milling of carbonyl Fe, Ag nanoxide, and nanopowders of Fe and cuprous oxide (Cu2O), with compositions of Fe–10% Ag, Fe–20% Ag, and Fe–25% Cu at a ball-to-powder ratio of 20:1 in hexane. Fe-Ag mixtures were milled for 8 h, while n-Fe-n-Cu2O mixtures were prepared for 6 h. With subsequent cold sintering, the resulting powders with high strength and ductility were intended for biodegradable high-strength implants with slow drug release [54]. Dense samples of Fe/Ag and Fe/Cu powders fabricated through cold sintering were also investigated. Fe–10% Ag, Fe–20% Ag, and Fe–25% Cu nanocomposite powders achieved densities close to theoretical values while maintaining their nanostructure, with samples compressed at 3 GPa exhibiting high plasticity and bending strengths exceeding 1000 MPa, 900 MPa, and 800 MPa, respectively [55].
Table 5. Bimetal synthesis by mechanical alloying.
Table 5. Bimetal synthesis by mechanical alloying.
Bimetal SystemExperimental MaterialsAdvantagesDisadvantagesReferences
Fe-Cu/Al collarFe powder (99.9% pure @125 µm particle size), Cu powder (99.5% @100 µm particles size), aluminum collar substrateSimple synthesis process
Cost-effective
Solid-state diffusion (non-melting process)
Environmentally friendly process
Time-intensive in milling process
Energy-intensive in milling process
[49]
Cu-Fe/CNTPure Fe powder, electrolytic Cu powder, and multi-walled carbon nanotubes (CNTs)CNTs as reinforcement to Cu/Fe bimetals,
application of controlled atmosphere (argon)
Time-intensive in milling process
Energy-intensive in milling process
Multi-step synthesis
[50]
Fe/Ag nanocomposite Carbonyl Fe and Ag2O powdersAnnealing process at 550 °C
Vacuum-drying
Controlled atmosphere (hexane + Ar)
Time intensive milling
Energy-intensive milling
Other energy requirements
[52]
Fe/Ag nanocomposite Carbonyl Fe and Ag2O powdersAttrition milling in hexane under Ar
Vacuum-drying
Heat treatment at 550 °C in H2
Time-intensive milling
High-energy milling
Other energy requirements
[53]
Fe/Ag, Fe/Cu nanocomposites Carbonyl Fe, Ag nanoxide, Fe and cuprous oxide nanopowdersAttrition milling in hexane under Ar
Consolidation process at 400 MPa
Application of cold-sintering
application of hydrogen treatment (450 °C)
Time-intensive milling
High-energy milling
Other energy requirements
Material costs
[54]
Fe/Ag, Fe/Cu nanocompositesCarbonyl Fe, silver oxide, Fe and cuprous oxide nanopowdersAttrition milling in hexane
application of cold-sintering
compression process at 3 GPa
Time-intensive milling
High-energy milling
Other energy requirements
Material costs
[55]

3.2. Electrical Explosion of Metal Wires

An electrical explosion occurs when a metal wire is heated with a current pulse of 106–109 A/cm2 [56], leading to the production of nanopowders of metals and alloys [57,58], the generation of shock waves [59], X-ray radiation [60], and other phenomena. The process is used for obtaining bimetallic and alloy nanoparticles, particularly through the electrical explosion of two intertwined wires (EEIW) made of dissimilar metals [61,62]. Pervikov et al. [63] investigated the heating and explosive destruction conditions of Al/Cu, Fe/Ti, Fe/Cu, and Fe/Pb wires under a current pulse density of 107 A/cm2. Their experiments showed that the energy deposited into wires depends on the thermophysical parameters and specific electric resistivity of the metals. This work also determined that the intertwined wires could explode synchronously or non-synchronously under current pulses. The experimental setup is shown in Figure 6, where MeA and MeB were the intertwined wires of dissimilar metals. Lerner et al. [64] utilized pure Fe and Cu wires (99.9% purity) with diameters of 0.2 and 0.3 mm to produce near-fully dense bimetallic Fe/Cu nanoparticles of different compositions (wt.%), namely, 72 Cu–28 Fe, 53 Cu–47 Fe, and 28 Cu–72 Fe, and as determined by the diameter combinations of the wires: 0.3 mm Cu–0.2 mm Fe, 0.2 mm Cu–0.2 mm Fe, and 0.2 mm Cu–0.3 mm Fe, respectively. The electrical explosion involved passing a high-density current (5 × 107 A/cm2) through braided Fe and Cu wires. The Fe-rich composition (72 Fe–28 Cu) achieved the highest yield strength of 700 MPa in compression and bending strength of 920 MPa, while the Cu-rich composition (28 Fe–72 Cu) demonstrated greater ductility and lower electrical resistivity.
Figure 6. Electrical explosion of intertwined wires of dissimilar metals: MeA and MeB. R and L are the impedance and inductance of the electric circuit of the setup: L = 0.75 μH and R = 0.08 Ω; the capacitance of the capacitor bank is C = 2.0 μF. Reprinted with permission from [63]. 2018, AIP Publishing.
Figure 6. Electrical explosion of intertwined wires of dissimilar metals: MeA and MeB. R and L are the impedance and inductance of the electric circuit of the setup: L = 0.75 μH and R = 0.08 Ω; the capacitance of the capacitor bank is C = 2.0 μF. Reprinted with permission from [63]. 2018, AIP Publishing.
Metals 15 00603 g006

3.3. Radiolysis

Radiolysis occurs when there is irradiation of aqueous solutions of metal salts with laser, X-rays, γ-rays, electrons, or ion beams, which leads to the dissociation of water molecules, generating solvated electrons and hydroxyl radicals that reduce metallic ions to form bimetallic nanoparticles, with their size and structure dependent on the dosage of metal precursors [30]. In the work of Chau et al. [65], bimetallic Fe/Pt nanoparticles were developed without a chemical reducing agent, using femtosecond laser irradiation as shown in Figure 7, where intense optical fields generate highly charged ions and molecules through the optical decomposition of metal precursors. The precursor solution was prepared by dissolving iron (III) chloride and chloroplatinic acid hexahydrate in deionized water, with polyvinylpyrrolidone (PVP) added as a dispersing agent. All experiments were conducted in a darkened room to prevent photosensitive reactions of the precursor molecules. Femtosecond laser irradiation was carried out using a Ti/Sapphire laser system (Newport Corporation, Spectra-Physics, Milpitas, CA, USA) with a 780 nm wavelength, 3 mJ pulse energy, and 100 fs pulse width, focusing the laser beam into the solution with an aspheric lens.
Figure 7. (a) Optical image of experimental set-up for femtosecond laser fabrication of nanoparticles, (b) magnified optical image of femtosecond laser irradiation of aqueous solution. Adapted from [65].
Figure 7. (a) Optical image of experimental set-up for femtosecond laser fabrication of nanoparticles, (b) magnified optical image of femtosecond laser irradiation of aqueous solution. Adapted from [65].
Metals 15 00603 g007

3.4. Sonochemical Method

Ultrasound-assisted synthesis has been widely used to produce nanostructured materials, including transition metals, alloys, carbides, oxides, and colloids. The sonochemical process is driven by acoustic cavitation, involving the formation, growth, and implosive collapse of bubbles in a liquid [66,67,68]. The collapse of cavitation bubbles generates localized hot spots with extreme conditions—temperatures reaching 10,000 K, pressures of about 1000 atm or higher, and cooling rates over 10⁹ K/s—facilitating chemical reactions and physical transformations, enabling the synthesis of nanomaterials with controlled particle size distribution and high catalytic activity. Table 6 presents a summary of studies that utilized sonochemical methods to synthesize various bimetals.
Zhao et al. [69] synthesized Ni/Fe bimetallic nanoparticles in a 500 mL three-necked flask under a nitrogen atmosphere, with ultrasonic elutriation at 20 kHz and 150 W using a rectangular-type ultrasonic apparatus as illustrated in Figure 8. Nanoscale ZVI (nZVI) particles were prepared by gradually adding NaBH4 solution (0.50 mol/L) into a flask containing FeSO4·7H2O solution (0.25 mol/L) under continuous stirring. The reaction was conducted at 25 °C using mechanical stirring with an external cold water circulation system, ensuring proper temperature control. The resulting nanoparticles were rinsed multiple times with deoxygenated deionized water to remove residual reactants and then used to produce Ni/Fe bimetallic nanoparticles by reacting with an aqueous solution of nickel sulfate hexahydrate (8.41 mmol/L) under continuous stirring.
In the work of Yu et al. [70], s-Fe/Cu bimetallic microparticles were prepared, at which sponge iron (s-Fe) particles were pretreated by washing with 1 mol/L hydrochloric acid (HCl) and rinsed with distilled water to remove surface oxides. The pretreated s-Fe particles were immersed in CuSO4 solution and subjected to ultrasonic treatment (200 W, 20 °C), enabling spontaneous Cu deposition via a redox reaction. The copper plating process was completed in 30 min, indicated by the color change of CuSO4 solution from deep blue to pale yellow due to ferric ion (Fe3+) release. The resulting s-Fe/Cu bimetallic particles were washed with distilled water, separated using a magnet, and used fresh for reduction experiments. Additionally, Tabrizian et al. [71] made Fe/Cu-GO (graphene oxide) nanocomposites by first preparing Fe/Cu bimetallic nanoparticles (BNPs) through the reaction of FeSO4·7H2O and CuSO4·5H2O under magnetic stirring, adjusting the pH to 7, and reducing with NaBH4 under nitrogen purging. The Fe/Cu BNPs were magnetically separated, washed with ethanol, dried, and stored at −20 °C. In the fabrication of Fe/Cu-GO nanocomposites, 50 mg of Fe/Cu BNPs was sonicated for 15 min in 10 mL of deionized water, followed by mixing with a 5 mL GO stock solution (0.5 mg/mL in deionized water) and vortexing for 30 s. The resulting nanocomposites (5% GO by weight) were magnetically separated and used fresh.
Figure 8. Schematic diagram of experimental apparatus for sonochemical methods. Reprinted with permission from [69]. 2014, Elsevier.
Figure 8. Schematic diagram of experimental apparatus for sonochemical methods. Reprinted with permission from [69]. 2014, Elsevier.
Metals 15 00603 g008
In the previous work of Li et al. [72], trace bimetallic Fe, manganese (Mn) co-doped N-ketjenblack carbon (Fe-Mn/KB) nanoelectrocatalyst was made using one-pot hydrothermal synthesis, mild calcination, and acid treatment, involving the dispersion of Mn(NO3)2, Fe(NO3)3, melamine, and ketjenblack carbon in deionized water in an ultrasonication bath for 30 min. The mixture was heated in a Teflon-lined container at 120 °C for 12 h, then cooled, filtered, dried, and ground into fine powder, calcined at 650 °C for 2 h under nitrogen. After acid treatment with hydrochloric acid and water at 80 °C for 6 h, the product was filtered, dried, ground, and calcined again at 650 °C for 1 h under nitrogen to obtain the final electrocatalyst. The electrocatalyst exhibited superior oxygen reduction reaction (ORR) activity and performance comparable to commercial Pt/C, maintaining a stable 1.50 V voltage platform for 20 h in Al–air battery tests. Aside from Fe-Mn/KB, composite Fe-Mn@SCAs were prepared through a multi-stage process, beginning with the synthesis of Fe/Mn nanoparticles (Fe-Mn NPs) by dissolving MnSO4, C6H5Na3O7∙2H2O, and K3[Fe(CN)6] in deionized water by ultrasound, followed by stirring for 30 min, centrifugation, and drying at 60 °C for 12 h in a vacuum oven. The Fe/Mn NPs were further corroded using a diluted ammonia water system to form Fe-Mn nanoclusters (Fe/Mn NCs), which were washed and vacuum-dried. In the final stage, Fe-Mn NCs was mixed with corn starch, gelatinized, retrograded at 4 °C, freeze-dried, and combusted at 700 °C to produce a porous carbon adsorbent. The final product, Fe-Mn@SCAs, is a bimetallic-doped starch-based porous carbon material [73].
Table 6. Bimetal synthesis by sonochemical method.
Table 6. Bimetal synthesis by sonochemical method.
Bimetal SystemExperimental MaterialsAdvantagesDisadvantagesReferences
Ni/FeFeSO4·7H2O, NiSO4·6H2O, NaBH4 (reducing agent)Mild synthesis conditions
Sequential reduction process
Versatile synthesis process
Multiple washing steps using deionized water[69]
s-Fe/CuSponge iron (s-Fe) particles, CuSO4·5H2OSimple and fast process
Low energy requirement
Magnetic recoverability
Ultrasound equipment dependency[70]
Fe/Cu-GOFeSO4·7H2O, CuSO4·5H2O, graphene oxide (GO), NaBH4 (reducing agent)Magnetic recoverability
Low-cost process
Graphene oxide as a support
pH-neutral synthesis
Multi-step process[71]
Fe-Mn/KBMn(NO3)2, Fe(NO3)3, melamine, ketjenblack carbonMild calcination conditions
Nitrogen atmosphere calcination
Carbon support integration (ketjenblack)
Energy requirements in calcination, heating ang drying
Long duration of heating and drying requirements
Multiple processing steps
[72]
Fe-Mn@SCAsMnSO4, C6H5Na3O7·2H2O, K3[Fe(CN)6], corn starch Eco-friendly support material
High porosity of the bimetallic material
Time-intensive synthesis process
Energy requirements in the synthesis
Multi-step process
[73]

3.5. Magnetic Field-Assisted Laser Ablation in Liquid (MF-LAL)

The assembly of one-dimensional (1D) magnetic nanoparticle (MNP) chains is highly desirable for advancing new materials and devices, typically requiring separate synthesis and fabrication steps. A novel one-step method, magnetic field-assisted laser ablation in liquid (MF-LAL), integrates NP synthesis and 1D chain fabrication [74], offering a green, efficient, and catalyst-free approach. MF-LAL enables customizable solid targets and solution environments, making it a versatile tool for creating ordered magnetic nanostructures [75]. Liang et al. [76], in their work, used the MF-LAL method (Figure 9) to fabricate one-dimensional (1D) chains of iron-based bimetallic alloy nanoparticles (Fe/Pt, Fe/Co, and Fe/Ni) using a Q-switch Nd/YAG laser and a 9 T magnetic field. The process involved alloying targets submerged in ethanol and irradiated with laser pulses (532 nm wavelength, 10 ns pulse width, 5 Hz frequency, 100 mJ/pulse) for 4 h, with the resulting solution collected for analysis. These 1D chains of alloyed MNPs exhibit ferromagnetism with high-saturation magnetization, low coercivity, and remanent magnetization, making them suitable for applications in magnetotransporters, micromechanical sensors, and magnetic memory materials.
Figure 9. Experimental setup of (a) MF-LAL and (b) a schematic illustration of the formation of 1D chains of MNPs. Reprinted with permission from [76]. 2014, American Chemical Society.
Figure 9. Experimental setup of (a) MF-LAL and (b) a schematic illustration of the formation of 1D chains of MNPs. Reprinted with permission from [76]. 2014, American Chemical Society.
Metals 15 00603 g009

3.6. Influence of Characteristics of Synthesized Bimetals by Physical Methods to Their Properties

Various characteristics and properties of synthesized bimetallic materials by physical methods are summarized in Table 7 below. In the last ten years, mechanical alloying was employed to produce Fe-based bimetallic materials with favorable magnetic properties. For instance, Hawili et al. [49] demonstrated that mechanical alloying of Fe and Cu powders enables solid-state interdiffusion, forming a uniform Fe-Cu alloy coating on an aluminum substrate. The coating achieved a thickness of up to 500 nm, indicating strong adhesion and effective surface coverage. The inclusion of Fe imparts magnetic properties to the coated material, enhancing its potential for magnetic or electromagnetic applications. On the other hand, Vishlaghi and Ataie [50] showed that adding CNTs to Cu-Fe particles via mechanical alloying refines the microstructure, with a higher CNT content (10 wt.%) reducing particle size. This nanoscale refinement improves magnetic properties by increasing surface area and interfacial bonding, facilitating suitability for electromagnetic applications.
A distinctive use of the mechanical alloying method was its effective application in developing Fe-based bimetallic materials intended for biomedical purposes. Sharipova et al. [52] synthesized Fe/Ag nanocomposites with Fe-5Ag and Fe-10Ag (vol.%) compositions. After annealing at 550 °C, these materials exhibited an optimal combination of strength and ductility. Another study by Sharipova et al. [53] produced Fe/Ag materials with 70% and 75% macroporosity, enhancing compressive and bending strength while maintaining high ductility. The porous, biocompatible structure of these materials also enabled controlled biodegradability, making them suitable for use in temporary biomedical implants. Furthermore, Sharipova et al. [54,55] synthesized Fe/Ag and Fe/Cu bimetallic materials with nanocomposite structures using mechanical alloying and compositions of Fe–10% Ag, Fe–20% Ag, and Fe–25% Cu (vol%). The refined microstructure and uniform distribution of alloying elements led to improved mechanical properties, including high strength and ductility. The use of biocompatible and degradable metals supports the biodegradability of these materials, making them suitable for temporary biomedical implants. Specifically, in Sharipova et al. [55], the materials achieved densities close to theoretical values, indicating successful compaction with minimal porosity. These characteristics also enhanced plasticity and bending strength, confirming the materials’ suitability for structural and biomedical applications.
The sonochemical method, another type of physical synthesis technique, was also utilized to produce Fe-based bimetallic materials exhibiting desirable magnetic properties. In the study by Yu et al. [70], s-Fe/Cu bimetallic materials were prepared as irregularly shaped microparticles, where the uneven morphology increased surface roughness and led to a non-uniform distribution of magnetic domains. This microstructural trait improved magnetic responsiveness, supporting the material’s magnetic behavior. The inherent magnetism of iron, combined with the unique particle shape, made the material suitable for magnetic applications. In a separate study by Tabrizian et al. [71], Fe/Cu nanoparticles were incorporated into a graphene oxide (GO) matrix using the sonochemical method to form Fe/Cu-GO nanocomposites. This uniform integration enhanced structural integrity and limited nanoparticle agglomeration, promoting good dispersion. The magnetic properties of Fe allowed for efficient magnetic separation, while the GO matrix contributed to greater surface area and better distribution, improving the composite’s reusability and stability in practical applications.
Another useful application of the sonochemical method is to prepare porous bimetallic materials for energy and catalysis. For example, in the study by Zhao et al. [69], Ni/Fe bimetals were synthesized via said method to form spherically shaped nanoparticles, and that which contributed to their uniform size and large surface area. This morphology significantly enhanced dispersity in solution, reducing particle agglomeration and improving overall stability. The spherical structure also facilitated greater accessibility of the active sites, thereby boosting the material’s adsorptive and catalytic performance. Additionally, the nanoscale configuration allowed for more efficient interaction with target molecules, supporting its application in environmental and catalytic processes. Additionally, in the study by Geng et al. [73], Fe-Mn@SCAs were made, forming Fe-Mn nanoclusters embedded within a porous carbon structure. This composite architecture provided a large number of accessible active sites and channels for mass transport. The integration of nanoclusters into the porous carbon matrix significantly increased the surface area and overall porosity. These characteristics are essential for enhancing catalytic efficiency and adsorption capacity, making the material suitable for energy and environmental applications. Lastly, in the study by Li et al. [72], the Fe-Mn/KB electrocatalyst was synthesized, resulting in uniform nanoparticles with a mean diameter of approximately 41 nm and distinct mesopores. The nanoscale uniformity and mesoporous structure enhanced the electrochemical surface area, facilitating efficient mass transport and active site accessibility. These structural characteristics contributed to superior oxygen reduction reaction (ORR) activity, with a performance comparable to commercial Pt/C electrocatalysts. Furthermore, the catalyst demonstrated a stable 1.50 V voltage platform and maintained durability over 20 h in Al–air battery tests, confirming its long-term operational stability.
Electrical explosion, radiolysis, and MF-LAL have emerged as effective physical methods to synthesize nanostructured bimetallic materials resulting in various properties. In the study by Lerner et al. [64], Fe/Cu bimetallic materials synthesized via the electrical explosion of metal wires exhibited a nanostructured and near-fully dense form across various compositions, including 72 Fe–28 Cu, 47 Fe–53 Cu, and 28 Fe–72 Cu (wt%). The Fe-rich composition (72 Fe–28 Cu) resulted in high mechanical strength, with a yield strength of 700 MPa and bending strength of 920 MPa due to the strength-contributing iron matrix. In contrast, the Cu-rich composition (28 Fe–72 Cu) offered greater ductility and lower electrical resistivity, benefiting from the inherent electrical properties of copper. Additionally, in the study by Chau et al. [65], Fe/Pt bimetallic materials were synthesized using radiolysis, resulting in nanometer-sized fine particles. This nanoscale characteristic contributed to the material’s high surface reactivity. Additionally, the incorporation of polyvinylpyrrolidone (PVP) enhanced particle dispersibility and enabled control over the mean particle size, offering improved stability and tunability for various applications. In the study by Liang et al. [76], Fe/Pt, Fe/Co, and Fe/Ni bimetallic materials were synthesized using MF-LAL, resulting in nano-sized, one-dimensional (1D) chain structures. This unique morphology facilitated the alignment of magnetic domains, contributing to strong ferromagnetic behavior. The 1D nanochains exhibited high saturation magnetization, which enhances magnetic responsiveness under an external field. Additionally, their nanostructured form enabled low coercivity and low remanent magnetization, making them ideal for applications requiring efficient magnetization and demagnetization cycles.

4. Chemical Methods for Synthesizing Iron-Based and Aluminum-Based Bimetals

Chemical methods are cost-effective and achieve high yields [30]. However, they involve toxic precursor chemicals and harmful solvents, which may generate hazardous by-products [39]. Most of the Fe-based and Al-based bimetals are synthesized by chemical methods such as chemical reduction, dealloying, seed-mediated growth, electrochemical synthesis, galvanic replacement, thermogravimetric, and supported particle methods.

4.1. Chemical Reduction

In chemical reduction, metal salts in solution are reduced by a reductant, making it the most common method for preparing metallic nanoparticles [77]. For bimetallic particles, two types of metals are reduced in solution, and their structure can be controlled using co-reduction. Co-reduction leverages the different redox potentials of metals, where the metal with a higher reduction potential forms the core, and the other precipitates onto its surface as a shell [26,77]. This strategy allows for tailoring bimetallic particle structures by adjusting factors such as reducing agent strength, metal ion reduction potentials, reaction temperature, and the properties of ligands and capping agents [78]. Table 8 examines chemical reduction techniques from different studies for developing bimetallic particles.
Raut et al. [79] prepared an Fe/Al bimetallic catalyst by dissolving 36 g of FeSO4·7H2O in 300 mL of deionized water, followed by the addition of 30 g of aluminum chloride, which turned the mixture gray after stirring for 15 min. Concentrated HCl as the reducing agent was then added dropwise, causing a dark-gray color change after 2 h of stirring; the mixture was filtered, washed with deionized water, and dried at 100 °C. In the study of Ou et al. [80], Fe/Al nanoparticles with a 1:1 Fe-to-Al weight ratio were synthesized, where ferric chloride was reduced to nZVI with 1 M sodium borohydride. Aluminum chloride and sodium borohydride were added to deposit Al onto the surface of the Fe particles. Fe/Al polymeric coagulants were recently produced for microplastic and nanoplastic removal. AlCl3 and FeCl3 solutions of 1 mol/L were prepared by dissolving AlCl3·6H2O and FeCl3·6H2O in distilled water, and the monomer coagulants were obtained by mixing these solutions at the desired Al/Fe ratio. Polymeric coagulants were prepared by slowly adding 50 mL of 0.4 mol/L NaOH solution to the monomeric coagulants, achieving a final concentration of 0.1 mol/L with a basicity of 2.0, and stored at 4 °C after measuring the pH [81].
Aside from Fe/Al bimetals, Fe/Cu bimetallic nanoparticles (NPs) with Fe/Cu mass ratios of 0.9:0.1, 0.75:0.25, and 0.5:0.5 were prepared with NaBH4 as the reducing agent, and modeled using molecular dynamics simulations. The solid product was isolated using ultracentrifugation, washed with an ethanol–water solution, degassed to remove residual salts, frozen at −18 °C, and lyophilized. The magnetic behavior of the synthesized NPs showed a decrease in magnetic saturation and coercivity (Hc) with increasing Cu concentration [82]. Ulucan-Altuntas and Kuzu [83] further developed Fe/Cu bimetallic nanoparticles using iron sulfate, sodium borohydride, copper sulfate, and ethanol, with sodium borohydride added dropwise to reduce iron sulfate to nanoscale zero-valent iron (nZVI), followed by the addition of copper sulfate to control the Fe/Cu composition. A 0.03 M iron sulfate solution with 17.25% ethanol and copper sulfate ratios of 2%, 5%, and 10% was used, with sodium borohydride added at a flow rate of 25 mL min−1, resulting in nanoparticles with a particle size of 82 nm. Additionally, another Fe/Cu bimetal was made by Abdel-Aziz et al. [84], this time by mixing FeCl3·6H2O with 0.002 M CuCl2·2H2O, still using NaBH4 solution as the reducing agent.
In the work of Muradova et al. [85], Fe/Cu bimetallic particles were synthesized by reacting 0.01 g CuCl2 with a colloidal solution of nZVI particles under oxygen-free conditions at room temperature. The reaction proceeded for 20 min, forming Fe/Cu particles with an nZVI core and a discontinuous Cu shell, with a copper concentration below the detection limit (<0.1 mg/L). The resulting nanoparticles had an average size of 13.3 nm with a standard deviation of 5 nm. Additionally, Torres-Blancas et al. [86] developed a binary Fe/Cu nanoparticle system by mixing 250 mL of 0.01 M iron sulfate (II) solution and 250 mL of 0.01 M copper sulfate solution at 300 rpm. The solution pH was adjusted with a dropwise addition of 0.5 M NaOH and monitored using a potentiometer. Chemical reduction of the particles was performed by adding an excess (1.1 M, 100 mL) of sodium borohydride solution under a nitrogen atmosphere, producing a black precipitate. The precipitate was stirred for 15 min, vacuum-filtered through a 0.2 mm cellulose acetate filter, and washed with ethanol and acetone to remove excess borohydride.
Naser and Shahwan [87] developed Fe/Ni magnetic nanoparticles by dissolving FeSO4·7H2O and NiCl2·6H2O in a 4:1 ethanol–water solution, followed by the dropwise addition of NaBH4 solution under continuous stirring. The nanoparticles were separated via vacuum filtration, washed with ethanol, dried at 90 °C for 6 h, and stored for future use. An SEM image of the Fe/Ni nanoparticles is shown in Figure 10 indicating a chain-like structure. In the work of Weng et al. [88], an Fe/Ni mixed solution was prepared by dissolving FeCl3·6H2O and NiSO4·6H2O in a 1:4 mixture of distilled water and ethanol, followed by the addition of NaBH4 as a reducing agent. The resulting Fe/Ni particles were collected via vacuum filtration, dried at 333 K under vacuum overnight, and used as a catalyst for the degradation of amoxicillin in aqueous solution. In the study of Zhou et al. [89], Fe/Ni bimetallic nanoparticles were developed using a potassium borohydride liquid-phase chemical reduction method under a nitrogen atmosphere. A solution containing 4.826 g FeCl3·6H2O (1 g Fe) and 0.238 g NiSO4·6H2O (0.05 g Ni) was prepared in 100 mL of ethanol–water solution (4:1 v/v) in a three-necked flask, maintaining an Fe/Ni mass ratio of 20:1. Ferric iron and nickel(II) were reduced to zero-valent Fe/Ni nanoparticles at 30 °C by adding KBH4 solution (4.816 g in 100 mL water) dropwise (~10 mL/min) under vigorous stirring (250 rpm), with the reaction continuing 30 min after hydrogen evolution ceased.
Figure 10. SEM image of Fe-Ni bimetallic NPs. Adapted from [87].
Figure 10. SEM image of Fe-Ni bimetallic NPs. Adapted from [87].
Metals 15 00603 g010
Magnetic Ni/Fe nanoparticles were also developed by Zhou et al. [90]. FeSO4·7H2O was dissolved in an ethanol–water solution and stirred for 10 min in a three-necked flask. KBH4 solution was added dropwise under nitrogen, producing black nZVI particles, which were magnetically separated, washed, and re-dispersed in ethanol. Ni/Fe bimetallic nanoparticles were synthesized by adding NiSO4·7H2O to the nZVI dispersion and stirring for 40 min. The Ni/Fe nanoparticles were centrifuged, washed with ethanol, and vacuum-dried at 70 °C for 8 h. Additionally, Fe/Ni bimetallic nanoparticles were prepared by Mansouriieh et al. [91] by mixing freshly synthesized nZVI with nickel nitrate hexahydrate (5 wt%) in 250 mL distilled, deionized water (DDW) at room temperature for 20 min. The Fe/Ni nanoparticles underwent a separation and washing process similar to that of nZVI particles to ensure purity. The final Fe/Ni bimetallic nanoparticles were dried under a nitrogen atmosphere to prevent oxidation and maintain stability.
Other Fe-based bimetals synthesized by chemical reduction include Fe/Ti, Fe/Co, and Fe/Mn. In the study of Liao et al. [92], an Fe/Ti bimetal was developed by reducing ferrous sulfate with NaBH4 under a nitrogen atmosphere, followed by the addition of titanium sulfate for surface modification and centrifugal separation. The synthesized bimetal was vacuum-dried for 24 h and had a high specific surface area. In Koryam et al. [93], Fe/Co nanoparticles were synthesized using iron and cobalt salts (1:1 mass ratio) in ethanol, with sodium hydroxide and hydrazine hydrate added to promote reduction under stirring. The magnetic nanoparticles were separated, washed with hot distilled water and ethanol, and dried in a vacuum oven at 40 °C for storage. Additionally, a micro-composite adsorbent, Fe/Mn, was synthesized via chemical reduction by preparing an equimolar Fe-Mn solution (100 mL) from iron chloride, manganese chloride, and a sodium tetraborate solution (50 mL). Sodium tetraborate was added dropwise to the Fe-Mn solution under stirring until the formation of dark brown microparticles. The microparticles were filtered, dried at 50 °C for 24 h, and used for selenium removal from water [94].

4.2. Chemical Dealloying

Dealloying is a selective dissolution process that removes specific components from an alloy, creating nanoporous structures [95]. Metal–organic deposition and liquid crystal templates can also produce nanoporous materials [96,97] but are complex and time-intensive. Dealloying methods, such as chemical dealloying, are preferred for their high productivity and precise controllability in fabricating nanoporous metals [98]. This method also allows precise control over the morphology and bimetallic composition of alloy nanostructures [99,100]. Nanoporous structures formed through this method are highly effective for applications in sensing, catalysis, and filtration [95].
In the study of Han and Xu [101], a nanoporous Pd/Fe (NP-Pd/Fe) electrocatalyst, intended for the oxygen reduction reaction, was fabricated through dealloying a Pd/Fe/Al source alloy. The Pd/Fe/Al alloy foils (~50 µm thick) underwent chemical treatment in 0.5 M NaOH for 48 h, followed by washing and air drying, to generate the Pd/Fe electrocatalyst. More recently, a nanoporous (NP)-Pt/Fe alloy was synthesized by Tian et al. [102] via the chemical dealloying of Pt5Fe15Al80 alloy ribbons, which were treated in 2.0 M NaOH at 50 °C for 48 h, washed with ultrapure water, and vacuum-dried for 12 h. The NP-Pt/Fe alloy was used for electrochemical sensor application.

4.3. Seed-Mediated Growth

Seed-mediated growth is a synthetic method for producing bimetallic nanoparticles using pre-synthesized metal crystals as seeds. In this process, atoms of a second metal, which could be obtained through the reduction or thermal decomposition of the corresponding metal salt, are deposited onto the seed metal [78]. The final structure of the bimetallic nanoparticles depends on how the second metal grows on the seed metal [103], forming either core–shell or heterostructures [104]. Core–shell structures are formed when the secondary metal uniformly coats the seed, while heterostructures arise when deposition occurs at specific sites. Adjusting thermodynamic and kinetic parameters during synthesis is crucial for controlling the morphology of bimetallic particles [77].
Wang et al. [105] prepared Ag/Fe bimetallic particles. Zero-valent iron (ZVI) particles were synthesized in an anaerobic chamber and cleaned with 0.4 M HCl to remove surface oxides, followed by rinsing with deoxygenated Milli-Q water. Silver was deposited onto the acid-washed ZVI particles using a dilute Ag2SO4 solution, and the resulting silver-coated particles were freeze-dried, appearing dark gray without visible oxidation. Silver deposition on ZVI particles was achieved at three different concentrations—0.0060, 0.0125, and 0.0228 mol%—assuming complete reduction and precipitation of Ag2SO4.
Meanwhile, Huang et al. [106] synthesized bimetallic Cu/Al particles by mixing a copper ion gel with aluminum metal particles in a fume hood. The copper ion gel was prepared by adding 1.5 g of sodium hydroxide to 30 mL cupric sulfate solution. A total of 5 g of aluminum particles was introduced into the gel, where sodium hydroxide removed surface oxides (e.g., Al2O3) from the aluminum. Cupric ions were rapidly reduced to zero-valent copper (Cu0) by ZVAl and deposited onto the aluminum surface via a redox reaction. Cu/Al particles with copper contents of 10, 20, and 40 wt% were prepared. Figure 11 presents an SEM image of Cu/Al bimetallic particles (20 wt% Cu), where small particles form rod-like aggregates on an aluminum support. EDX analysis confirms the presence of elemental copper and aluminum in the material.
Figure 11. SEM image of bimetallic Cu/Al particles. (a) Magnification: 12,000×; (b) magnification: 90,000×; (c) SEM–EDX spectrum of Cu/Al particles. Reprinted with permission from [106]. Copyright 2015, Elsevier.
Figure 11. SEM image of bimetallic Cu/Al particles. (a) Magnification: 12,000×; (b) magnification: 90,000×; (c) SEM–EDX spectrum of Cu/Al particles. Reprinted with permission from [106]. Copyright 2015, Elsevier.
Metals 15 00603 g011

4.4. Electrochemical Synthesis

Electrochemical deposition is a cost-effective and straightforward nanoparticle synthesis technique [107]. It enables the formation of core–shell configurations and nanostructures that are challenging to achieve using traditional chemical reduction or complexation methods [108]. By reducing metal ions onto selected electrodes from an electrolyte solution, this method allows the controlled growth of nanostructures with varying sizes and shapes on metal [109], semiconductor [110], and polymer surfaces [111], influenced by parameters such as deposition potential, precursor solution composition, and exposure duration [112,113].
Nanostructures combining noble and magnetic metals can enhance performance in catalysis and sensing by precisely tuning their size, shape, composition, and surface properties [114,115,116,117]. Bimetallic nanowires with high aspect ratios are particularly suitable for applications in sensors, catalysts, and electro-catalysts. Hence, Riva et al. [118] synthesized magnetic Fe/Rh nanowires with various nominal compositions (FexRh100-x, x = 15, 25, 54) via AC electrodeposition into alumina hard templates with 20 nm hexagonal nanopores. The electrodeposition was conducted at room temperature using an aqueous bath containing FeSO4·7H2O, RhCl3, 0.75 g/L ascorbic acid (to prevent iron oxidation), and 30 g/L H3BO3, with a pH of 4.0. The process used an AC voltage of 15–20 Vrms at 60–100 Hz for a few minutes, with a two-electrode setup where the aluminum in the template acted as the working electrode and a graphite rod as the auxiliary electrode. In the same conditions, specifically polycrystalline Fe90Rh10 nanowire arrays with a diameter of 20 nm and lengths of 1–3 mm were successfully prepared by Riva et al. [119]. They studied the effects of adding Rh to Fe nanowires in terms of microstructure, magnetic hysteresis, and magnetoresistance properties. In Figure 12, SEM images display side views of (top) an alumina template with 20 nm pore diameter, and (bottom) Fe90Rh10 nanowires after the aluminum support and alumina template have been dissolved. In 2017, another group produced Fe90Rh10 nanowires which were approximately 18 nm in diameter and 1 μm in length, using the same experimental conditions. In this study, the magnetization mechanism at room temperature and the thermal stability of nanowire magnetic configurations were examined by looking into the dependence of the coercive field on the applied field sweeping rate [120].
Moreover, using electrodeposition at a current density of 2.8 A cm−2 for 20 s, Hao et al. [121] prepared micro–nano-dendritic Fe/Zn alloy, and the influence of the electrolyte iron and zinc content on its morphology, composition, and Fenton-like phenol degradation performance was studied. The electrolyte was prepared by dissolving specific amounts of FeSO4·7H2O and ZnSO4·7H2O in ultrapure water, creating varying concentrations of FeSO4 (0.35 to 0.50 mol L−1) and ZnSO4 (0 to 0.15 mol L−1) along with 50 mL L−1 of anhydrous ethanol. Electro-deposited Fe–xZn products, labeled based on the Zn ion content (wt.%) in the electrolyte, were washed and dried under vacuum at 60 °C.
Figure 12. SEM micrographs showing side views of an alumina template of a 20 nm pore diameter (a) and of Fe90Rh10 nanowires, after dissolving the aluminum support and the alumina template (b). Reprinted with permission from [119]. Copyright 2016, Elsevier.
Figure 12. SEM micrographs showing side views of an alumina template of a 20 nm pore diameter (a) and of Fe90Rh10 nanowires, after dissolving the aluminum support and the alumina template (b). Reprinted with permission from [119]. Copyright 2016, Elsevier.
Metals 15 00603 g012

4.5. Galvanic Replacement

According to Gu et al. [103], the galvanic replacement method involves initial metal seeds that act as a reductant for the second metal precursor. When the redox potential of the seeded metal is lower than that of the second metal, the seeded metal atoms dissolve by losing electrons, which are accepted by the second metal ions. The dissolution and deposition sites are influenced by the surface capping agents involved in the reaction. The size and morphology of the final product can be easily controlled by selecting seeded metal of varying sizes and shapes or by adjusting the extent of replacement [122]. Table 9 summarizes the use of the galvanic replacement method from different studies.
In Yuan et al. [123], Fe/Cu bimetallic particles were synthesized by adding ZVI to a CuSO4 aqueous solution. The particles were prepared by depositing copper onto ZVI through a metal displacement reaction, with Cu mass loadings ranging from 0.05 to 1.26 g per gram of Fe, utilizing the high standard reduction potential difference between Cu2+ and ZVI. Mahmoud and Mahmoud [124] developed bimetallic Fe-Cu nanoparticles by adding 1 g of freshly prepared nZVI to a copper sulfate (CuSO4·5H2O) solution at a controlled rate of 0.1 g per 60 s under vigorous stirring. The CuSO4·5H2O solution was prepared by dissolving 0.1 g of CuSO4 in 100 mL of an ethanol–distilled water mixture (1:1) at 60 °C. After mixing, the solution was left to settle for 15 min, indicating copper deposition by the color change of nZVI, then filtered, washed with ethanol, dried at 60 °C for 5 h, and stored under nitrogen to prevent oxidation. Mahmoud et al. [125] also developed Fe/Cu nanoparticles using the same process and precursors.
Furthermore, Lai et al. [126] synthesized Fe/Cu bimetallic particles via displacement plating by mixing iron particles with CuSO4 solution for 10 min, followed by a 5 min precipitation. The particles were rinsed with deionized water and ethanol and then dried under nitrogen protection at 80 °C for 40 min. Copper was deposited onto iron particles via an iron–copper replacement reaction, with the Cu mass loading (0.05–1.81 gCu/gFe) adjusted by varying the CuSO4 concentration to influence the catalytic activity of Fe/Cu bimetallic particles. Another study by Lai et al. [127] prepared Fe/Cu bimetallic particles under varying Cu2+ concentrations (1–12 g/L), with CuSO4 and CuCl2 tested as copper salts. The influence of the planting solution pH (3.0, 4.0, 4.6, and 5.0) and stirring speeds (100–400 rpm) was studied under optimal conditions to refine the preparation process further. The impact of theoretical Cu mass loading (0.03–1.81 gCu/gFe) was also investigated. In the work of Xiong et al. [128], Fe/Cu bimetallic particles were synthesized using the Fe-Cu displacement reaction in an aqueous solution, following the primary preparation procedure described in the work of Lai et al. [129]. However, in this study, the key preparation parameters were optimized, including a theoretical Cu mass loading (TMLCu) of 0.41 g Cu/g Fe, a temperature of 40 °C, a mixing speed of 250 rpm, and a Cu2+ concentration (CuSO4) of 3 g/L in the planting solution.
In the work of Ren and Lai [129], microscale ZVI particles were used as substrates to prepare Fe/Cu bimetallic particles, with copper deposited on the ZVI surface via electroless plating. The plating bath consisted of CuSO4·5H2O (11.25 g/L), sodium hypophosphite (50 g/L), boric acid (H3BO3), NiSO4·7H2O, and complexing agents such as TCD, PSTT, EDTA, En, or TEA. ZVI particles were added to 400 mL of the plating bath, and the slurry was mixed using a mechanical agitator at 300 rpm, with the process performed at 70 ± 1 °C. Additionally, Fe–Cu bimetallic powder was prepared by depositing Cu2+ onto Fe0 through a chemical reaction—Fe0 + Cu2+ → Fe2+ + Cu0—followed by drying the powder in an argon atmosphere (100 mL/min) at room temperature for 24 h. The dried Fe/Cu powder was collected and stored in a refrigerator for further use. The nominal mass ratio of Cu to Fe in the powder was calculated as 0.06742 using the equation WCu/WFe = mCu/(m0 − m1), where mCu is the Cu content, m0 is the initial ZVI mass, and m1 is the reacted ZVI mass [130].
Aside from Fe/Cu bimetals, Fe/Al and Fe/Mg bimetallic particles were also developed. In the study of Fu et al. [131], Fe/Al bimetallic particles were synthesized by reacting Al with Fe2+ ions. The process began with 3 g of ZVAl in distilled water, to which hydrochloric acid and FeSO4 solution were sequentially added. The mixture was stirred at 400 rpm for 15 min, adjusting Fe mass loading via FeSO4 concentration, then rinsed with distilled water and dried in a vacuum freeze desiccator. Cheng et al. [132] developed mesoporous and ferromagnetic Fe/Al particles using the same precursors for As(III) removal. Xiang et al. [133] were also able to produce Fe/Al particles by depositing nZVI nanoparticles onto ZVAl particles, this time via the reaction of Fe2+ and BH4− with the process performed at 80 °C under nitrogen protection. The synthesis involved dissolving ZVAl and FeSO4·7H2O in distilled water, adjusting the pH to 7 with NaOH, gradually adding a NaBH4 solution, and maintaining the reaction at 80 °C for 4 h, followed by centrifugation, washing, and drying in an oxygen-free vacuum freeze desiccator. Additionally, Mg/Fe bimetallic particles were prepared by reducing Fe2+ to Fe0 using Mg0 particles in a deoxygenated FeSO4 solution, followed by rinsing and vacuum freeze-drying at −90 °C for 24 h. Atomic absorption spectroscopy (AAS) analysis confirmed that iron was almost entirely deposited on magnesium particles, enabling the determination of the Mg-to-Fe molar ratio in the bimetallic particles. Mg/Fe bimetallic particles with molar ratios of Mg to Fe at 99:1, 32:1, and 13:1 were successfully prepared for further studies. At a potential of −1.0 V, these bimetallic particles exhibited a higher reduction current density (1.36 mV/cm2) compared to Mg powder (0.95 mV/cm2), indicating their superior oxygen reduction reaction (ORR) activity [134].
Some studies conducted acid-washing pretreatment of the metal precursor prior to bimetallic synthesis. Aghaei et al. [135,136] synthesized Fe/Al bimetallic particles using ZVAl powder and ferric chloride, with acid-washing to remove aluminum oxide from ZVAl using 1 M HCl at 40 °C. Fe3+ was then electrochemically reduced and deposited onto ZVAl by adding Fe3+ solutions and agitating for 30 min. The resulting Fe/Al particles were rinsed with deionized water and dried in a vacuum desiccator. In the study of He et al. [137], Fe/Al bimetallic particles were developed by depositing iron onto zero-valent aluminum. Zero-valent aluminum and ferrous sulfate were used as precursors, with aluminum powder pretreated by washing with hydrochloric acid and deionized water to ensure a fresh aluminum surface. The Fe/Al particles obtained were filtered, washed, and freeze-dried for 24 h under nitrogen to prevent oxidation and then stored in airtight plastic bags to avoid further oxidation. Yeh et al. [138] also synthesized Fe/Al bimetallic particles by acid-washing Al powder, mixing it with FeCl3 solution, and cooling the reaction before rinsing it with deionized water. The study also examined the effectiveness and mechanisms of these Fe/Al particles in inactivating Escherichia coli (E. coli). In the study of Liu et al. [139], granular ZVI (99.94% purity) was acid-washed, neutralized, and reacted with a CuSO4 solution for 15 min to synthesize Fe/Cu powder. The powder was washed, dried under argon at room temperature for 24 h, and stored in a refrigerator.
Park et al. [140] explored a novel recycling route for synthesizing magnetic Fe/Al bimetallic materials from six types of Al alloys (1050, 2024, 3003, 5083, 6061, and 7075). The Al alloy sheets were treated in a 1.6 M NaCl and 0.5 M FeCl2 solution with varying HCl concentrations (0–0.4 M), temperatures (25–50 °C), and durations (15–60 min), and the Fe-cemented samples were analyzed to quantify the Fe deposition. In the study of Lien et al. [141], Fe/Al bimetallic material was synthesized through acid treatment of aluminum followed by the in situ reduction of Fe. The process involved stirring 5.0 g of aluminum scrap in a beaker with 10 mL of deionized water, to which 10 mL of HCl was added, triggering a vigorous reaction. Ferric chloride was added to the mixture to form the Fe/Al, which was then cooled, washed, and dried. Tabelin et al. [142] also utilized aluminum scraps to synthesize magnetic Fe/Al bimetallic materials using a two-stage mechanical–chemical process, which involved polishing these scraps to enhance the surface area and an etching-cementation process to dissolve Al-oxide films and deposit ferric ions (Fe3+) onto the aluminum surface. Etching–cementation experiments were conducted using 0.5 M or 1.0 M FeCl3 solutions in NaCl and HCl, agitated at 50 strokes/min for varying durations, with treated Al-scraps rinsed and dried at 40 °C.
A galvanic couple between two metals causes the oxidation of the metal with the lower reduction potential by the metal ions with the higher reduction potential [143]. In Liu et al. [144], Fe-Al nanopowder was synthesized using granular ZVI and ZVAl, washed with 1% HCl using magnetic stirring, and neutralized by purging with deoxygenated ultrapure water. The neutralized powder was dried at room temperature under an argon atmosphere for 24 h, then collected and stored in a refrigerator for future use. Fan et al. [145] also produced an Fe-Al bimetallic powder employing granular ZVI and ZVAl. The resulting powder was acid-washed and purged with deoxygenated, ultrapure water until neutralized to remove acidic residues. Additionally, in the work of Fan et al. [146], an Fe/Al bimetallic nanopowder was also prepared using granular ZVI and ZVAl without acid-washing or surface cleaning to evaluate the reactivity of iron and aluminum samples with their native oxide layers and the capacity of oxalic acid to induce oxidative reactions.
Table 9. Bimetal synthesis by galvanic replacement.
Table 9. Bimetal synthesis by galvanic replacement.
Bimetal SystemExperimental MaterialsAdvantagesDisadvantagesReference
Fe/CuMetal precursors: CuSO4, ZVISimple synthesis process
Controlled Cu mass loading
High Cu loading can be costly (e.g., 1.26 g Cu/g Fe)[123]
Fe/CuMetal precursors: CuSO4·5H2O, nZVISimple and controllable synthesis
Storage of particles under nitrogen atmosphere
Energy requirements in the overall process
Long processing time
[124,125]
Fe/CuMetal precursors: CuSO4·5H2O, nZVIControlled Cu loading
Efficient Cu deposition
Use of nitrogen atmosphere in the process
Multi-step synthesis method
Drying requirements at 80 °C
[126]
Fe/CuMetal precursors: CuSO4, CuCl2, ZVIWide temperature range studied
Variable Cu2+ concentrations
Controlled Cu loading
pH influence studied
Stirring speed was also varied
Longer coverage of study
Time-intensive study
Material and cost considerations
[127]
Fe/CuMetal precursors: CuSO4·5H2O, nZVISimplicity of the process
Controlled Cu loading and mixing speed
Mild operating temperature (40 °C)
Fixed Cu2+ concentration
Energy requirements in drying (40 °C for 40 min)[128]
Fe/CuMetal precursors: CuSO4, ZVIUse of electroless plating in the synthesis
Fixed copper concentration (11.25 g/L CuSO4·5H2O)
Controlled temperature (70 ± 1 °C) in the process
Controlled agitation
Complex chemical system
involvement of many chemicals
Processing temperature requirements (70 ± 1 °C)
[129]
Fe/CuMetal precursors: CuSO4·5H2O, ZVIVersatile synthesis method
Controlled Cu/Fe mass ratio
Room-temperature drying
Use of argon atmosphere
Long drying time
Refrigeration storage adds complexity
[130]
Fe/AlMetal precursors: FeSO4, ZVAl Simple synthesis method
Rapid synthesis (15 min reaction time)
Controlled Fe mass loading
Use of concentrated HCl[131,132]
Fe/AlMetal precursors: FeSO4·7H2O, ZVAl Simple synthesis method
Controlled pH in the synthesis
Controlled addition of NaBH4 solution
Use of nitrogen atmosphere
Energy-intensive process
Long processing time
[133]
Mg/FeMetal precursors: FeSO4·7H2O, ZVMgSimple and versatile synthesis method
Rapid synthesis process (2 min reaction time)
Controlled Mg/Fe ratios
High energy requirement for freeze-drying
Long drying time
[134]
Fe/Al Metal precursors: FeCl3·6H2O, ZVAl Simple synthesis method
Involving acid-washing of ZVAl particles
Good control over Fe/Al ratio
Mild reaction conditions
Requiring acid (HCl) pretreatment[135,136]
Fe/Al Metal precursors: FeSO4, ZVAl Controlled Fe loading
Al powder is pretreated with HCl and deionized water
Use of nitrogen atmosphere
Use of HCl in acid pretreatment
Long drying time
[137]
Fe/Al Metal precursors: FeCl3, ZVAlSimple synthesis
Al powder is pretreated with HCl
Controlled Fe loading
Processing under ambient conditions
Use of HCl in acid pretreatment[138]
Fe/Cu Metal precursors: CuSO4·5H2O, ZVISimple synthesis
method
Rapid synthesis (15 min reaction time)
Pretreatment of ZVI with dilute HCl
Room-temperature drying
Controlled Cu/Fe ratio
Use of argon atmosphere in the process
Use of HCl in acid pretreatment
Long drying time
[139]
Fe/Al Metal precursors: FeCl2, Al alloys (1050, 2024, 3003, 5083, 6061, and 7075)Simplicity and versatility of the process
Mild temperature conditions (25–50 °C)
Shorter reaction times (15–60 min)
Magnetic recoverability of the particles
Involving different types of Al alloys (adds material cost)
Dependence on acid (HCl) concentration
[140]
Fe/Al Metal precursors: Ferric chloride, Al scrapSimple synthesis process
Efficient Fe deposition
Pretreatment of Al scrap with HCl
Use of Al scrap as ZVAl source
Use of HCl in acid pretreatment[141]
Fe/Al Metal precursors: FeCl3, Al scrapsVersatile and innovative synthesis method
Controlled Fe deposition
Mild drying conditions (40 °C)
Adjustable reaction durations (0.5 to 6 h)
Magnetic recoverability of the particles
Use of Al scraps as ZVAl source
Labor-intensive process
Long processing time (up to 6 h)
[142]
Fe/AlMetal precursors: ZVI (source of Fe2+), ZVAlSimple process
Room-temperature drying
Use of argon atmosphere
Long drying time (24 h)
Use of strong acid (HCl) in the process
Refrigerator storage requirement for the particles
[144,145]
Fe/AlMetal precursors: ZVI (source of Fe2+), ZVAl Simple process
Room-temperature drying
No acid (HCl) treatment involved
Use of argon atmosphere
Long drying time (24 h)
Refrigerator storage requirement for the particles
[146]

4.6. Thermogravimetric Method

The thermogravimetric method is carried out through non-catalytic gas–solid reactions. Non-catalytic gas–solid reactions are integral to many chemical and metallurgical processes [147,148], such as reducing metal oxides. These reactions typically involve the interaction between a gas phase and solid particles aggregated in a pellet. As gas diffuses through the pellet, chemical reactions can occur simultaneously, driving the process forward [147]. This is different from other metallurgical processes where a solid acts as a reducing agent [149,150,151].
Meshkini Far et al. [152] performed the reduction of Fe2O3 and NiO powders. Prior to this, Fe and Ni powders were dissolved in nitric acid, refluxed for 30 min, and then neutralized with ammonium hydroxide at pH 7. The solution was concentrated, the residue evaporated, and the resulting solid calcined at 350 °C for 4 h to yield Fe2O3 and NiO powders. The oxides were reduced to Ni/Fe nanoporous catalysts in a hydrogen–helium gas stream at 300 °C for 4 h.
Tang et al. [153] prepared Nix/Fe catalysts from NixFeLDH precursors. NixFeLDH precursors were synthesized via co-precipitation using nitrate salts of Ni, Mg, Fe, and Al with a cation ratio of [Mg2+]:[Al3+] = 2:1 and a total concentration of 1 mol·L−1, maintaining pH 10.3 ± 0.1. The precursors were aged at 80 °C for 12 h, dried, and calcined at 500 °C for 5 h to yield mixed metal oxides, named NixFe-CLDHs. The calcined precursors were reduced in 5% H2/Ar at 800 °C for 3 h at a heating rate of 5 °C·min−1 to yield the final catalysts, named NixFe-red.
In the study of Shao et al. [154], Ni/Fe-based catalysts were synthesized via the reduction of LDH precursor. NiFe-based catalysts with varying molar ratios of Ni, Fe, Mg, and Al were synthesized via co-precipitation using Mg-Al LDH as the precursor, with nitrate solutions added to a carbonate solution at pH 10 ± 0.5 and 40 °C. The resulting mixture was crystallized at 65 °C, aged for 24 h, filtered, washed, and dried at 80 °C to produce the LDH precursor. The LDH precursor was calcined at 600 °C for 4 h in static air to produce oxides, then reduced at 600 °C for 2 h under a H2/N2 gas flow to obtain metallic species.
Li et al. [155] prepared Ni-Fe alloy nanoparticles from Ni–Mg–Fe–Al hydrotalcite-like compounds (HTlcs). Ni–Mg–Fe–Al HTlcs were calcined at 1073 K for 5 h in a static air atmosphere, then pressed into a disk, crushed, and sieved to particles of 30–60 mesh size (0.3–0.6 mm). The resulting Ni–Fe/Mg/Al catalysts, obtained after calcining HTlcs, have a fixed (Ni + Mg)/(Fe + Al) molar ratio of 3 and varying Fe/Ni ratios from 0.1 to 1.5, with a fixed Ni loading of 12 wt% and Fe content ranging from 1.3 to 18.7 wt%. The catalysts were reduced with a H2/N2 gas flow (30/30 mL/min) at 1073 K for 0.5 h to obtain Ni–Fe alloy nanoparticles.
Zhao et al. [156] developed Co/Fe-alloy and Ni/Fe-alloy products from layered double hydroxides (LDHs) and Co/Fe/Al-LDH and Ni/Fe/Al-LDH nanosheets. Co/Fe/Al-LDH and Ni/Fe/Al-LDH nanosheets were synthesized via a urea-assisted co-precipitation method by dissolving the respective metal nitrates with urea in deionized water, refluxing at 110 °C for 24 h, and collecting the products through centrifugation, washing, and drying under vacuum at 60 °C. The synthesized nanosheets were subjected to reduction in a H2/Ar (10/90 v/v) flow at 650 °C for 5 h, with a heating rate of 5 °C/min. The resulting CoFe-alloy, NiFe-alloy, and NiCo-alloy products were cooled slowly to room temperature under nitrogen flow.
Liu et al. [157] also produced Ni/Fe alloy products from Ni-Fe layered double hydroxides (LDHs). Layered double hydroxides (LDHs) with varying Ni/Fe molar ratios were synthesized via a co-precipitation method using a mixed nitrate solution of Ni2+, Fe3+, Al3+, and Mg2+ at a 1 mol/L total concentration and a ([Ni2+ + Mg2+]/[Fe3+ + Al3+]) molar ratio of 2. The precursors (NixFe1−x)-LDHs were precipitated at pH 9.5 using NaOH/Na2CO3, aged at 60 °C for 12 h, filtered, dried at 80 °C, calcined at 500 °C for 3 h to form mixed metal oxides (NixFe1−x)-LDO, and then reduced at 700 °C in a hydrogen atmosphere for 3 h. The final reduced products were designated as (NixFe1−x)-red.
Lastly, in the study of De Masi et al. [158], bimetallic Fe30/Ni70 nanoparticles (NPs) were synthesized via co-decomposition of {Fe[N(SiMe3)2]2}2 and Ni[iPrNC(CH3)NiPr]2 in the presence of palmitic acid as a stabilizer under 3 bars of H2. It was reported that these nanoparticles demonstrated a high heating capacity and excellent catalytic activity for fully selective CO2 conversion to methane under a low magnetic field.

4.7. Supported Particles

According to Ferrando et al. [159], nanoparticles can be synthesized by supporting them on substrates such as graphite, silicon, or inorganic oxides like silica. The substrate used in developing bimetallic adsorbents significantly influences their physicochemical properties and the extent of metal loading [33], playing a crucial role in wastewater treatment [160]. As a result, the efficiency of various substrates has been extensively evaluated for optimizing bimetallic adsorbent performance [33].

4.7.1. Carbon-Based Materials as Support

In this review, most Fe-based bimetals were synthesized using this method, and with carbon-based materials such as carbon nanotubes (CNTs), activated carbon, and biochar, among others, as shown in Table 10. For instance, Wang et al. [161] prepared Fe/Ce-NCNT by mixing the precursors FeCl3·6H2O, Ce(NO3)3·9H2O, and melamine in ethanol, stirring for 10 h, drying, and ball milling to obtain a yellow powder. A black product was synthesized via two-step pyrolysis: heating the precursor to 350 °C for 30 min and then to 800 °C for 2 h at a rate of 5 °C min−1 in an argon atmosphere, followed by impurity removal using hydrochloric acid and deionized water washing. The final product was named Fe/Ce-NCNT-0.2. In the study of Tian et al. [162], multi-walled carbon nanotubes (MWCNTs) were grown via chemical vapor deposition (CVD) using an Fe-Al precursor at 680 °C under N2 and propane flow, producing materials applicable for high-rate rechargeable Li-ion batteries.
In Wang et al. [163], Fe/Cu bimetallic carbon nanofibers (CNF) were synthesized by dissolving copper (II) acetate monohydrate (Cu(ac)2·H2O), iron (III) acetylacetonate (Fe(acac)3), and 1 g of polyacrylonitrile (PAN) in 20 mL dimethylformamide (DMF) to prepare a homogeneous precursor solution, followed by electrospinning at a flow rate of 1.47 mL/h under an 18 kV applied voltage. The resulting PAN-based nanofibers were stabilized at 240 °C in air for 3 h and then carbonized at 900 °C under nitrogen in a tube furnace using a heating rate of 5 °C/min. for 2 h to produce FeCu/CNF. Additionally, Nam et al. [164] synthesized a noble-metal-free Cu/Fe alloy encapsulated in a graphitic carbon shell, showing high efficiency and durability as an electrocatalyst for the oxygen reduction reaction (ORR) in alkaline solutions. The catalyst precursor, prepared by drying an ethanol–water mixture of chlorophyllin and iron (II) acetylacetonate, was heat-treated at 800 °C under argon for 1 h, with individual Cu and Fe components labeled after similar processing.
Moreover, Fe/Cu bimetallic nanoparticles embedded in an ordered mesoporous carbon composite (Cu-Fe/MC) were synthesized using a “one-pot” block-copolymer self-assembly method, involving the preparation of Resol Resins from phenol and formaldehyde. The composite catalyst was prepared by dissolving Pluronic F127, adding iron and copper precursors, and mixing with Resol resin, followed by evaporation, polymerization, and pyrolysis at 800 °C under nitrogen [165]. Wu et al. [166] used biochar to produce Co-Fe with modified biochar (MB). Pristine biochar was treated with NaOH to produce modified biochar (MB), which was then used to prepare Fe/MB by soaking in a solution of FeSO4·7H2O and PEG-4000, followed by treatment with NaBH4 under nitrogen. The Fe/MB was further subjected to displacement plating with CoSO4·7H2O to synthesize Co-Fe/MB, then washed, dried at 60 °C for 24 h, and stored.
Wu and Feng [167] carried out the synthesis of Ag/Fe/modified biochar (MB) where 2.0 g of biochar was immersed in a solution containing ethanol, distilled water, PEG (4000), and FeSO4·7H2O, stirred at 1000 r/min, and ultrasonicated for 2 h. Nano-zero-valent iron (nZVI) was synthesized on biochar by the dropwise addition of 0.6 mol/L NaBH4 solution at 0 °C under intense stirring for 2 h, followed by thorough washing with distilled water and absolute ethanol to prepare nZVI/MB for further modification. Ag/Fe/MB was produced by adding AgNO3 solution to the nZVI/MB under intense stirring for 2 h, resulting in the reduction and deposition of Ag onto the nZVI surface, forming a thin discontinuous silver layer. In the work of Xing et al. [168], biochar (BC)@Fe/Ni composites were made under nitrogen protection in a 500 mL three-necked flask using ethanol and water as the reaction medium. FeCl2 solution was added to the flask with BC powder from heated ground straw (at 600 °C) and stirred under a nitrogen atmosphere, followed by the introduction of polyethylene glycol for enhanced dispersion, and the dropwise addition of NaBH4 solution using a peristaltic pump, with 15 min of stirring to complete the reduction reaction. Afterward, NiCl2 solution was added dropwise, triggering a displacement reaction on Fe0, leading to the deposition of Ni metal (Fe0 + Ni2+ → Fe2+ + Ni0). The reaction suspension was magnetically separated, washed with deionized water and ethanol, and vacuum-dried to yield BC@Fe/Ni composites. The preparation process is shown in Figure 13. Recently, biochar has been employed for Fe-bimetal synthesis, where a study developed magnetic Fe–Co-modified biochar (FMBC) containing iron (Fe) and cobalt (Co) bimetals after NaOH activation, using forestry waste cedar bark as the raw material through pyrolysis, to explore its capacity for the adsorption of ofloxacin (OFX) [169].
Figure 13. Schematic diagram of the BC@FeNi preparation process. Reprinted with permission from [168]. Copyright 2022, Elsevier.
Figure 13. Schematic diagram of the BC@FeNi preparation process. Reprinted with permission from [168]. Copyright 2022, Elsevier.
Metals 15 00603 g013
Ji et al. [170] prepared nZVIC (Fe-Cu)–municipal sludge-derived biochar (SBC). For nZVI-SBC, 4.2 g SBC was added to 250 mL of 0.15 M FeSO4∙7H2O solution and stirred rapidly at 25 °C for 10 min. Freshly prepared NaBH4 (100 mL, 0.79 M) stabilized with 0.5% NaOH was added, and the mixture was stirred for 30 min. The resulting carbon nanoparticles were vacuum-filtered, freeze-dried, and labeled as nZVI-SBC. For nZVIC-SBC, the FeSO4∙7H2O solution was replaced with a mixed solution of FeSO4∙7H2O (0.15 M) and CuSO4∙5H2O (0.00063 M), requiring additional NaBH4 (100 mL, 0.82 M) during synthesis. Also, Xia et al. [171] loaded Zn/Fe NPs onto a corncob biochar (CBC) surface (ZF@CBC). A mixture of ZnSO4∙7H2O (2.5 mmol) and FeCl3∙6H2O (5.0 mmol) in 300 mL ultrapure water was stirred for 15 min. NaOH (0.1 mol) was added to the Zn/Fe solution, and the mixture was oscillated at 45 °C for 30 min. In total, 5 g of CBC was added to the Zn/Fe mixture and oscillated at 45 °C for 24 h. The solid was rinsed with ultrapure water, oven-dried at 80 °C, and pyrolyzed at 450 °C for 2 h to produce ZF@CBC.
Activated carbon is often involved in the synthesis of carbon-based materials. In the study of Bose et al. [172], activated carbon (AC) was impregnated with iron using co-precipitation and reduction, forming Fe-AC by mixing AC with FeSO4·7H2O, adding NaOH and NaBH4, and reducing the ferrous iron, followed by washing and drying. Cobalt impregnation was carried out by mixing Fe-AC with CoCl2·6H2O, followed by the addition of NaBH4 and NaOH, heating to 150 °C for 3 h under reflux, resulting in the bimetallic Fe-Co/AC. The final Fe-Co/AC product was washed with double-distilled water to neutralize it, dried at 80 °C for 24 h, and stored for further use. Danmaliki and Saleh [173] used activated carbon (AC) derived from waste rubber tires which was dispersed in a mixture of deionized water, ethanol, and ethylene glycol, followed by the addition of cerium nitrate, refluxing at 90 °C, and drying at 110 °C overnight. The dried material was redispersed, treated with ferric nitrate, refluxed again, filtered, washed, dried, and calcined at 350 °C to localize the metals on the AC surface. Further, nZVI-Ni/AC particles were developed by Tian et al. [174]. Nanoscale zero-valent iron (nZVI) was synthesized using FeSO4·7H2O and PEG-4000 under an inert atmosphere, with activated carbon (AC) added to enhance dispersion and reactivity, followed by a reduction reaction with KBH4. After synthesizing nZVI, NiCl2·6H2O was added to the mixture, and the particles were stirred at room temperature before vacuum-drying for 48 h. The dried particles were heat-treated in a hydrogen atmosphere at 600 °C, resulting in Fe-Ni/AC activator particles.
In continuation, Kakavandi et al. [175] developed activated carbon (AC) modified by ZVI and silver (Fe-Ag) bimetallic and magnetic nanoparticles. Powder-activated carbon (PAC)-ZVI was synthesized via co-precipitation and reduction methods, involving the dissolution of FeSO4·7H2O in methanol–water, followed by the addition of PAC and pH adjustment to 7 with NaOH. NaBH4 was added to reduce ferrous iron to ZVI, which was coated onto the PAC, and the PAC-ZVI particles were separated, washed, and dried under nitrogen. PAC-ZVI was mixed with AgNO3 solution at 200 °C to form bimetallic nanoparticles, followed by NaBH4 reduction, magnetic separation, thorough washing, and drying under nitrogen. In another study by Fong et al. [176], granular activated carbon was sieved into a fine powder and used for the synthesis of Ag-Fe/CAC (CAC, commercial activated carbon) through metal ion impregnation, NaBH4 reduction, and a displacement reaction. Initially, 2.0 g of powdered activated carbon was added to 50 mL of a coating solution containing 0.05 g of polyethylene glycol (PEG-600), along with iron (II) sulfate heptahydrate (FeSO4·7H2O) in the range of 2.0 to 3.0 g. The nZVI/CAC was treated with NaBH4 solution, washed, and then combined with AgNO3 solution for silver deposition on nZVI surfaces, resulting in the final Ag-Fe/CAC product. Rahaman et al. [177] made Fe/Cu bimetallic nanoparticle-impregnated activated carbon derived from coconut husk (CAC). Iron solutions were prepared by dissolving FeSO4 and FeCl3, which were added to a CAC dispersion, followed by NaOH to reach a pH of 11, and the suspension was stirred and washed until neutral. The iron-impregnated CAC was dried, then treated with copper sulfate and NaBH4, resulting in the formation of bimetallic Fe-Cu/CAC nanoparticles. The bimetallic nanoparticles were washed and dried under vacuum.

4.7.2. Alumina (Al2O3) as Support

Some studies also involved alumina (Al2O3) in synthesizing bimetallic particles. Zhao et al. [178] used atomic layer deposition (ALD) to synthesize Fe-Ni/Al2O3. Initially, Ni/Al2O3 was prepared by dissolving nickel nitrate hexahydrate in deionized water, impregnating it onto alumina, followed by drying at 100 °C and calcination at 600 °C for 6 h. Iron was deposited onto Ni/Al2O3 using a fluidized atomic layer deposition (ALD) system, with nitrogen as a carrier gas and ozone for oxidizing the iron precursor, repeated for 3, 6, and 12 cycles to ensure uniform deposition. In another study, a Ni3Fe/Al2O3 catalyst was synthesized via a homogeneous deposition–precipitation method using nitrate precursors (Ni(NO3)2·6H2O and Fe(NO3)3·9H2O) in a 3:1 molar ratio, urea, and high-surface-area Al2O3 calcined at 600 °C. The precursor suspension was refluxed at 90 °C, processed through filtration and washing, dried at 110 °C, and finally calcined at 500 °C for 4 h under static air to complete the preparation [179]. In the study conducted by Sun et al. [180], Fe-Cu bimetallic catalysts were synthesized by mixing a surfactant solution of Pluronic P123 in ethanol with dissolved metal precursors (Fe(NO3)3∙9H2O, Cu(NO3)2∙3H2O, and Al(NO3)3∙9H2O) and citric acid monohydrate, followed by stirring, drying into a gel at 333 K for 3 days, and calcination at 1023 K for 4 h. The catalysts are named xFeyCu-Al2O3 to represent the mass fractions of Fe and Cu. Further, γ-Al2O3-supported Fe−Ru catalysts (with a Ru/Fe atomic ratio of 0.1) were synthesized by Liuzzi et al. [181] via a reduction–deposition method, ensuring Ru deposition on the outer layer of metal particles. The process involved reducing Fe(SO4) to metallic form using NaBH4 under a nitrogen atmosphere, followed by Ru(NO)(NO3)3 addition for Ru deposition on the reduced nanoparticles. The Fe−Ru nanoparticles were combined with γ-Al2O3 to achieve Ru and Fe loadings of 1 wt% and 5.5 wt%, respectively, with the final catalysts labeled as Ru-Fe/Al2O3.
Zhang et al. [182] prepared Pd-Fe bimetallic nanoparticles immobilized on an Al2O3/Polyvinylidene difluoride (PVDF) membrane. The modified Al2O3/PVDF membrane was coated with a solution of polyacrylic acid, ethylene glycol, and FeSO4·7H2O, thermally treated at 115 °C for 3 h to form a crosslinked composite, and then immersed in a 0.5 mol/L KBH4 solution for 15 min to form zero-valent iron nanoparticles on the membrane. The membrane with iron NPs was soaked in a palladium acetate solution, resulting in the deposition of palladium onto the iron nanoparticles, creating Pd/Fe NPs immobilized on the Al2O3/PVDF membrane. Pradhan et al. [183] carried out the synthesis of Co-Fe/Al2O3–MCM-41. In total, 0.5 mmol of Co(NO3)2·6H2O and 0.5 mmol of FeSO4·7H2O were mixed in ethanol and oleic acid. The total mixture was transferred to a stainless steel autoclave and heated for 20 h at 120 °C in a furnace. The resulting gel was washed with distilled water and ethanol, dried at 70 °C for 12 h, and calcined at 500 °C for 5 h in air. Additionally, there were also bimetallic catalysts, Pd-Fe/γ-Al2O3 and Rh-Fe/γ-Al2O3, that were made on γ-Al2O3 powder via incipient wetness impregnation using PdCl2, RhCl3, and Fe(NO3)3·9H2O as precursors, with palladium or rhodium and iron co-impregnated in a single step. The Pd and Rh loadings were set at 1% (w/w), and Fe loading was fixed at 4% (w/w). After impregnation, the materials were dried at 60 °C for 12 h, calcined at 200–400 °C for 4 h in air, and reduced under H2 flow for 2 h at 350 °C [184].

4.7.3. Silica as Support

Silica has also been used to support Fe/Cu and Fe/Al bimetals. Wang et al. [185] developed Fe/Cu–hollow mesoporous silica sphere (HMS) bimetallic composites, with 0.1 g of HMS added to 5 mL of aqueous solution containing 20 mg FeSO4 and 15 mg Cu(NO3)2, then stirred for 1 h under a nitrogen atmosphere. The concentration of FeSO4 and Cu(NO3)2 solutions were 4 g/L and 3 g/L, respectively, with a theoretical mass percentage of 3.7 wt% for both Fe and Cu on HMS. The suspension was dried under vacuum at 50 °C and then treated with 1 mL of NaBH4 aqueous solution for 3 h under nitrogen. The molar ratio of NaBH4 to the total metal (Fe + Cu was maintained at 6:1 to ensure sufficient NaBH4 for nanoparticle formation, after which the products were immersed, filtered, and washed with methanol three times. Another work by Wang et al. [186] produced 2Fe6Cu/HMS. In an Fe/Cu mass ratio of 2:6, 0.1 g of HMS was added to a 5 mL solution containing 0.01 g FeSO4·7H2O and 0.023 g Cu(NO3)2·3H2O, then stirred for 1 h under a nitrogen atmosphere. The suspension was dried under vacuum at 50 °C, followed by the addition of 1 mL of NaBH4 solution, and stirred for 3 h under nitrogen to form the nanoparticles. The molar ratio of NaBH4 to total metal was still at 6:1 for adequate reduction, and the products were immersed, filtered, and washed with methanol, yielding final catalysts denoted as 2Fe6Cu/HMS.
In Lin et al. [187], transition metal oxide catalysts with a total metal loading of 5 wt.% (Cu + Fe) were synthesized using an in situ auto-combustion method, involving the dissolution of metal nitrate precursors in water, the addition of glycine in a 1:1 molar ratio (glycine/NO3−), and mixing with SBA-15 support material, followed by aging and solvent evaporation at 100 °C. The dry powder underwent glycine combustion at 300 °C and calcination at 500 °C for 6 h, producing catalysts labeled as tM1-tM2/SBA-15, where “t” represents metal loading in wt.% and M1 and M2 are Fe and Cu, respectively. Additionally, Yan et al. [188] produced Fe-Al-SBA-15 catalysts with varying Fe and Al loadings using a microwave-assisted heating method, with tetramethyl orthosilicate (TMOS), iron nitrate, and aluminum isopropoxide as precursors, and P123 as the structure-directing agent. The synthesis involved preparing solution A (P123 in HCl) and solution B (precursors in water), followed by adding solution B dropwise to solution A to form an initial gel with a specific molar composition. The gel underwent microwave-assisted stirring at 313 K for 4 h, hydrothermal treatment in a Teflon autoclave at 373 K for 24 h, and subsequent calcination at 823 K to produce mesoporous Fe(x)-Al(y)-SBA-15-MW samples.

4.7.4. Minerals as Supports

Minerals have also been used as supports, as shown in Table 11. Bentonite has been widely used to remove hazardous substances from water and wastewater [189,190,191]. It has a high adsorption capacity for heavy metals and organic substances due to its high lattice charge and cation exchange capacity (CEC), which typically ranges from 40 to 130 meq/100 g [192]. In light of this, Sabouri et al. [193] used bentonite to immobilize Fe-Cu nanoparticles to degrade acidic dyes from aqueous media. Initially, nZVI was immobilized on bentonite (Be@Fe). Bentonite was mixed with FeCl3 solution, treated with NaBH4 under argon flow, stirred, and then separated, washed, and dried. For bimetallic Be@Fe-Cu nanoparticles, the Be@Fe nanoparticles were treated with CuSO4 solution under argon flow at 30 °C, followed by magnetic separation, washing with methanol, and drying. The entire process was carried out under argon to ensure oxidation protection and proper synthesis of Be@Fe and Be@Fe-Cu nanoparticles. In the study of Weng et al. [194], bentonite (B)-Fe/Ni particles were produced; the preparation involved mixing 1 g of bentonite with a solution containing ferric chloride and nickel sulfate in a water–ethanol mixture (1:4 ratio) under mechanical stirring. A 0.47 M NaBH4 solution was added dropwise to the mixture under nitrogen, fully reducing Fe3+ and Ni2+ to Fe0 and Ni0, respectively. The products were collected via vacuum filtration, rinsed thoroughly with water and ethanol to prevent oxidation, and dried under vacuum at 333 K for 12 h.
Kaolin is a low-cost and stable clay mineral [195] and could be a potential porous material for supporting nZVI in removing metal ions from contaminated water [196]. The work of Jin et al. [197] produced kaolinite (K)-Fe/Pd particles. K–Fe particles were first synthesized with a 1:1 iron/kaolin mass ratio by dissolving ferric chloride hexahydrate (4.84 g) in a 50 mL water–ethanol mixture, adding 1 g kaolinite, and reducing with dropwise addition of sodium borohydride solution (0.47 M, 100 mL) under nitrogen. The K–Fe particles were vacuum-filtered, rinsed with water and ethanol, and combined with palladium acetate (0.0106 g) dissolved in 30 mL ethanol under ultrasonic conditions for 5 min. The resulting K–Fe/Pd particles were vacuum-filtered, rinsed with ethanol, dried at 333 K under vacuum, and stored as a powder in a nitrogen atmosphere.
Zeolite, discovered in 1756 by F. Crondtedt [198], is a highly porous, cost-effective material with excellent cation-exchange capabilities, widely utilized as an adsorbent for wastewater treatment due to its natural abundance and industrial scalability [199]. In the work of Xu et al. [200], Cu/Fe@zeolite was synthesized using impregnation and reduction methods, with Cu/Fe@zeolite-1 prepared by stirring (60 min) zeolites (20–40 g/L) in a Cu2+ solution (27.31 g/L) under nitrogen, followed by static settlement and skimming. Cu/Fe@zeolite-2 was prepared similarly but involved adding Fe2+ to the Cu/zeolite suspension to reduce Cu2+ to Cu+ without precipitation separation. Both products were separated by centrifugation at 5000× g rpm for 15 min, rinsed with oxygen-free water, and freeze-dried. The freeze-drying process occurred at 223.2 K and 1 Pa for 24 h to obtain the final Cu/Fe@zeolite materials.
Diatomite, sepiolite, and palygorskite were also used as supports in Fe-Ni bimetals. For diatomite (Di)-Fe/Ni, a solution of 9.65 g FeCl3·6H2O and 0.90 g NiSO4·6H2O was prepared by dissolving them in 37.5 mL ethanol and 12.5 mL deionized water, then stirred for 20 min. A total of 2 g of purified diatomite (Di) was added to the bimetallic solution, which was then stirred for 2 h under a nitrogen atmosphere. A 1.1 M NaBH4 solution was prepared by dissolving 4.16 g of NaBH4 in 100 mL of deoxygenated deionized water and added dropwise to the mixture of Di and bimetallics. The Fe3+ and Ni2+ ions in the reaction mixture were reduced to Fe0 and Ni0 nanoparticles over 2 h, yielding a nanocomposite named “Di-Fe/Ni”, which was washed with ethanol (50 mL) for three times, dried at 60 °C, and stored in a vacuum desiccator [201].
For palygorskite (Pal)-Fe/Ni nanocomposite, a solution of 9.67 g FeCl3·6H2O and 0.92 g NiSO4·6H2O was prepared in a water–ethanol mixture and stirred for 20 min. Two grams of pretreated palygorskite was added to the Fe/Ni solution and stirred under a nitrogen atmosphere for 2 h. To ensure the saturation of palygorskite with Fe/Ni nanoparticles, excess bimetallic particles were removed by centrifugation at 3000× g rpm for 30 min. The final nanocomposite was created by adding a 1.1 M NaBH4 solution dropwise to the slurry, promoting the reduction of Fe3+ and Ni2+ to Fe0 and Ni0 nanoparticles. After stirring for 2 h, the product was filtered, washed with ethanol, and dried at 60 °C to obtain the Pal-Fe/Ni composite [202]. On the other hand, to develop sepiolite (Sep)-Fe/Ni, Fe/Ni bimetallic nanoparticles were prepared using FeCl3 (5.79 g) and NiSO4·6H2O (0.90 g) in a solution of distilled water and absolute alcohol (12.5 mL:37.5 mL) with a mixing ratio of Ni/Fe = 0.095:1 (wt/wt). Preliminary tests identified an optimized Ni/Fe ratio of 0.095:1, achieving 100% removal of 2,4-dichlorophenol while balancing efficiency, cost, and environmental safety, as excessive Fe increased pH and low Fe/Ni loading was ineffective. Sepiolite (Sep) was then suspended in a bimetal solution, stirred under a nitrogen atmosphere for 2 h, and combined with 1.1 M NaBH4 solution added dropwise to synthesize Sep-Fe/Ni [203].
Table 11. Synthesis of bimetals with minerals as supports.
Table 11. Synthesis of bimetals with minerals as supports.
Bimetal SystemExperimental Materials AdvantagesDisadvantagesReference
Be@Fe-Cu Metal precursors: FeCl3∙6H2O,
CuSO4∙5H2O
Others: Bentonite, NaBH4 (reducing agent)
Support: Bentonite (Be)
Mild synthesis conditions
Sequential metal impregnation for controlled deposition
Use of an inert argon atmosphere
Magnetic recoverability of the particles
Multi-step synthesis process
Long processing time
[193]
B-Fe/NiMetal precursors: FeCl3∙6H2O,
NiSO4∙6H2O
Others: Bentonite, NaBH4
Support: Bentonite (B)
Scalable and simple synthesis process
Controlled chemical reduction via NaBH4
Use of inert nitrogen atmosphere
Energy-intensive drying requirement
Long processing time
[194]
K-Fe/PdMetal precursors: FeCl3∙6H2O,
CuCl2∙2H2O
Others: Natural kaolinite, NaBH4
Support: Kaolinite (K)
Versatile synthesis process
Ultrasonic treatment improves Pd deposition
Use of inert nitrogen atmosphere
Energy-intensive drying requirement
Multiple ethanol washing steps
Long processing time
[197]
Cu/Fe@zeoliteMetal precursors: FeSO4·7H2O, CuCl2·2H2O
Other(s): Zeolite
Support: Zeolite
Multiple synthesis routes provide flexibility
Simplicity of the synthesis methods
Controlled stirring and centrifugation in the syntheses
Time-Consuming Processes
Energy-intensive drying requirement for Cu/Fe@zeolite-2 synthesis
[200]
Di-Fe/NiMetal precursors: FeCl3∙6H2O,
NiSO4∙6H2O
Others: Diatomite (Di), NaBH4
Support: Diatomite (Di)
Simplicity and versatility of the synthesis
Controlled addition of NaBH4 solution
Use of nitrogen atmosphere
Multi-step process
Multiple ethanol washing steps
Pre-processing of diatomite is required
Moderate temperature (60 °C) drying is required
Long processing time
[201]
Pal-Fe/NiMetal precursors: FeCl3∙6H2O,
NiSO4∙6H2O
Others: Palygorskite, NaBH4
Support: Palygorskite (Pal)
Versatility of the synthesis method
Controlled addition of NaBH4 solution
Controlled stirring and centrifugation in the process
Use of nitrogen atmosphere
Multi-step procedure
Pre-processing of palygorskite is required
Drying requirements of the overall process
Long processing time
[202]
Sep-Fe/NiMetal precursors:
FeCl3, NiSO4∙6H2O
Others: Sepiolite, NaBH4
Support: Sepiolite (Sep)
Versatility of the process
Controlled Ni/Fe composition
Controlled addition of NaBH4 solution
Use of nitrogen atmosphere
Drying requirements of the overall process
Pre-processing of sepiolite is required
Long processing time
[203]

4.7.5. Other Material Supports

In the study of Svarovskaya et al. [204], a novel hierarchical micro/nanostructured flower-like composite, AlOOH/AlFe, was synthesized using a simple one-pot method under mild conditions, utilizing water and Al/Fe(N2) nanopowder as the precursor. Bimetallic Al/Fe(N2) nanoparticles were synthesized via the electric explosion of aluminum and iron wires (50 wt% Fe and 50 wt% Al) in a nitrogen atmosphere at 3 × 10⁵ Pa, using a capacitor bank with 2.8 μF capacitance and 26 kV charging voltage. The synthesis involved reacting 4.00 g of Al/Fe nanopowder with 400 mL of water at 60 °C for 3 h with air bubbling, followed by centrifugation and drying at 100 °C for 4 h. Hina et al. [205] worked on Pd-Fe bimetallic catalysts supported in an AlF3 matrix. Aluminum ethoxide (Al(OEt)3) was synthesized by dissolving AlCl3 in ethanol, and a 0.5%Pd–0.5%Fe/AlF3 catalyst was prepared by adding PdCl2 and FeCl3, with ethylenediamine included as a pore-forming agent prior to HF addition. HF solution (40% v/v) was gradually introduced to the mixture, forming a viscous gel that was dried at 355 K, calcined in air at 473 K for 1 h, and subsequently reduced in hydrogen at 723 K for 4 h.
Xi et al. [206] synthesized Pd/Fe alloy nanoparticles on N-doped carbon layer functionalized on aluminum silicate fibers (ASF@NC)/PdFe. Aluminum silicate fibers (ASFs) were immersed in a Tris buffer solution with dopamine hydrochloride, K2PdCl4, and FeCl3, allowing dopamine to polymerize into polydopamine (PDA) while the metal salts reduced and anchored within the PDA layer, and the resulting ASF@PDA/PdFe composite was washed, collected, and freeze-dried. The ASF@PDA/PdFe composite was subjected to carbonization under an argon atmosphere at 900 °C for 2 h, producing the ASF@NC/PdFe catalytic fiber. The annealing process involved gradual heating at 10 °C/min, holding the temperature steady, and cooling naturally under argon protection to obtain the final product. On the other hand, Aftab et al. [207] worked on Fe-Co@polyacrylamide (PAM) hydrogel, and that the PAM hydrogel was initially synthesized via free-radical polymerization using MBA as a crosslinker and APS/TEMED as an initiator, with the mixture solidifying into a gel at 25 °C and stored for 24 h to ensure complete polymerization. The hydrogel was washed with deionized water to remove unreacted components and then immersed in a bimetallic solution of FeCl3·6H2O and Co(NO3)2·6H2O at a 1:1 weight ratio for 24 h. After immersion, the hydrogel was rinsed in deionized water for 12 h to remove unbound metal ions, producing the filtered bimetallic hydrogel (FeCo@PAM). The FeCo@PAM hydrogel was treated with a solution of ethylene glycol and 0.5 M hydrazine monohydrate at 60 °C for 6 h to achieve the in situ reduction of metal ions. The FeCo@PAM hydrogel, thoroughly washed with deionized water, freeze-dried for 24 h, and stored in a sealed container at ambient temperature, retained its catalytic activity.
Fe/Pd nanoparticle-assembled filter paper was also produced; the filter paper was first immersed in a polyethylenimine (PEI) solution (20 mg/mL) for 30 min, washed with water, and then treated with an FeCl3 solution (0.4 µM) under shaking for 3 h. The Fe3+ ions complexed onto the PEI-coated filter paper were reduced using an ice-cold NaBH4 solution (0.94 M) for 30 min to form Fe nanoparticles. The Fe nanoparticle-coated filter paper was subsequently reacted with a K2PdCl4 solution (0.04 µM) for 1 h to produce Fe/Pd nanoparticle-assembled filter paper [208] (see Figure 14), while in Ge et al. [209], fly ash (FA), an industrial waste from coal power plants [210,211], was used in the preparation of Cu/Fe-BM@FA (Cu/Fe bimetallic modified fly ash): CuCl2·2H2O and FeCl3 were dissolved in deionized water at concentrations of 222 mmol/L Cu2+ and 111 mmol/L Fe3+. The Cu2+ and Fe3+ solution was co-precipitated with 0.666 mol/L NaOH, maintaining pH 8 with adjustments using 1 mol/L HCl and 2 mol/L NaOH. A total of 0.4 g of FA was added to the Cu/Fe bimetallic compound suspension, forming Cu/Fe-BM@FA through gradual mixing. The mixture was filtered and dried at 80 °C for 10 h, resulting in Cu/Fe-BM@FA as the sorbent.

4.8. Influence of Characteristics of Synthesized Bimetals by Chemical Methods to Their Properties

Chemical methods have been widely utilized for synthesizing iron-based and aluminum-based bimetallic materials, as discussed in Section 4.1, Section 4.2, Section 4.3, Section 4.4, Section 4.5, Section 4.6, and Section 4.7. Table 12 summarizes the selected synthesized bimetallic products, highlighting their key characteristics and resulting properties.
Chemical reduction has been widely recognized as an effective synthesis method for producing bimetallic materials with strong adsorptive and catalytic properties. For instance, in the work by Ou et al. [80], Fe/Al bimetallic nanoparticles were prepared with sodium borohydride as the reducing agent. The synthesis yielded spherically shaped nanoparticles, which provided a high surface-to-volume ratio. This morphology increased surface reactivity and promoted enhanced interaction with contaminants. The uniform shape also enabled efficient electron transfer between Fe and Al, improving the material’s catalytic activity. Muradova et al. [85] synthesized Fe/Cu bimetallic nanoparticles via chemical reduction, forming a core–shell structure. The core–shell structure featured a discontinuous Cu shell on an nZVI core, which facilitated electron flow from Fe to Cu. This configuration facilitated the catalytic performance by increasing active sites for redox reactions, aiding nitrate removal. In another study, Torres-Blancas et al. [86] also developed Fe/Cu nanoparticles with spherical morphology. The spherical structure improved surface contact and electron transfer, resulting in a strong degradation ability for pollutant remediation.
A distinctive application of chemical reduction is the fabrication of bimetallic materials that exhibit both high adsorptive and catalytic performance, while also possessing magnetic recoverability for enhanced reusability. Sepúlveda et al. [82] synthesized Fe–Cu bimetallic nanoclusters via chemical reduction with NaBH4, adjusting the Fe:Cu mass ratios to 0.9:0.1, 0.75:0.25, and 0.5:0.5. These nanoclusters exhibited a high surface area and numerous active sites, which significantly enhanced their adsorptive removal capabilities. The presence of iron imparted magnetic properties, enabling convenient magnetic separation from solution. Varying the Fe/Cu ratios allowed for optimization between magnetic response and adsorption performance for targeted applications. In Koryam et al. [93], Fe–Co nanoparticles were produced through chemical reduction using NaOH and hydrazine hydrate, forming spherical particles. This spherical structure increased surface contact and, combined with the magnetic nature of Co and Fe, promoted both effective adsorption and magnetic recoverability. Naser and Shahwan [87] developed Fe/Ni bimetallic nanoparticles via chemical reduction, producing a unique chain-like structure. This morphology offered increased surface area and active sites, which improved the material’s adsorption efficiency. The combination of iron and nickel provided strong magnetic properties, allowing for straightforward magnetic separation post-use. The interconnected chain-like design also supported better dispersion and accessibility in solution, enhancing contaminant removal. In Zhou et al. [90], Fe/Ni nanoparticles were also synthesized using KBH4, resulting in nickel particles evenly distributed on iron surfaces. This arrangement boosted surface reactivity and electron transfer between metals, contributing to a high adsorptive, reductive, and magnetic performance.
Chemical dealloying is recognized for developing materials with nanoporous structures [95,98] and has been reported by some researchers for bimetallic synthesis. For instance, in the study by Han and Xu [101], an NP-Pd/Fe electrocatalyst was created using chemical dealloying of a Pd/Fe/Al alloy, producing an open nanosponge-like structure. This porous form offered a large surface area and interconnected channels, which improved access to active catalytic sites. The greater exposure of these sites enhanced electrocatalytic efficiency, especially for the oxygen reduction reaction. Furthermore, the open framework enabled effective electron and mass transport, boosting the catalyst’s activity and durability. On the other hand, Tian et al. [102] synthesized an NP-Pt/Fe alloy from Pt5Fe15Al80 alloy ribbons, resulting in interconnected strips with nanoporous features and pores measuring a few nanometers. This intricate structure provided a high surface area and improved access to active sites. The nanoscale porosity facilitated faster electron transfer and reactant diffusion, significantly enhancing electrocatalytic performance. As a result, the NP-Pt/Fe alloy showed electrocatalytic activity, making it suitable for electrochemical sensor applications.
Seed-mediated growth has also attracted attention for synthesizing bimetallic particles that exhibit both catalytic and degradation properties. In the study by Wang et al. [105], Ag/Fe bimetals were synthesized via seed-mediated growth, resulting in microparticles with diameters ranging from 2 to 10 μm. This relatively large particle size provided structural stability and facilitated ease of handling in catalytic applications. The distribution of Ag on Fe surfaces enhanced the availability of catalytic active sites, promoting efficient reaction processes. Consequently, the microparticle morphology contributed to the bimetal’s notable catalytic capability. Additionally, in another work by Huang et al. [106], a Cu/Al bimetal was fabricated forming a core–shell structure where Cu particles appeared as rod-like aggregations on aluminum surfaces. This configuration increased the active surface area and provided more reactive sites for interaction with pollutants. The close contact between the Cu shell and Al core facilitated efficient electron transfer, which is essential for breaking down contaminants. As a result, the structural features of the bimetal directly enhanced its degradation capability.
Electrochemical synthesis has been utilized to fabricate bimetallic nanowires with magnetic properties. Riva et al. [118] produced Fe/Rh bimetallic materials using electrochemical synthesis, forming dispersed polycrystalline nanowires approximately 18 nm in diameter and 1 μm in length. This elongated nanoscale structure provided a high aspect ratio, enhancing magnetic anisotropy. The fine dispersion and dimensional characteristics supported stable magnetic alignment, particularly under low-temperature conditions. As a result, the nanowire morphology contributed significantly to the material’s low temperature magnetic capability. Another work by Riva et al. [119] developed Fe/Rh bimetals, forming polycrystalline nanowire arrays with diameters of 20 nm and lengths ranging from 1 to 3 mm. The polycrystalline structure contributed to enhanced magnetic anisotropy due to grain boundary effects. The high aspect ratio of the nanowires supported strong directional magnetic alignment. As a result, the structural characteristics directly influenced and improved the magnetic properties of the material.
The galvanic replacement approach has been widely applied to synthesize common bimetallic materials like Fe/Cu and Fe/Al intended for catalytic and adsorptive processes. Additionally, one study reported the fabrication of a Mg/Fe bimetal, which was evaluated for both its degradation performance and electrochemical behavior. In the work by Yuan et al. [123], Fe–Cu bimetallic particles were prepared, leading to dispersed Cu particles on Fe surfaces. This dispersion increased the number of catalytic active sites. The close interface between Cu and Fe supported effective electron transfer, promoting catalytic efficiency. Therefore, the bimetal demonstrated favorable catalytic capability due to its surface structure. Mahmoud and Mahmoud [124] also developed Fe–Cu nanoparticles with sizes between 20 and 30 nm and irregular surface structures through galvanic replacement. The small particle size provided a high surface area, improving contact with contaminants. The irregular morphology also offered numerous active sites, enhancing adsorption. These features collectively improved the material’s adsorptive performance. In the study by Liu et al. [139], Fe/Cu particles were fabricated with microscale dimensions, with Cu particles deposited on the surface of Fe. This surface deposition of Cu enhanced the availability of reactive sites, facilitating effective interaction with contaminants. The close contact between Cu and Fe promoted efficient electron transfer, which is essential for redox-based degradation processes.
In the study by Xiang et al. [133], Fe/Al bimetallic materials were also produced via galvanic replacement, forming a core–shell configuration where Fe particles were layered onto Al surfaces. This design increased surface exposure and presented more active Fe sites for interacting with pollutants. The intimate contact between Fe and Al promoted the development of reactive interfaces, boosting the material’s adsorptive performance. Similarly, in the study by He et al. [137], the Fe/Al bimetallic particles synthesized displayed a similar core–shell structure with Fe coating Al surfaces. This structure allowed for greater access to reactive sites, improving contaminant binding efficiency. The synergistic interface between Fe and Al played a central role in strengthening the adsorptive capability of the bimetal. Another study by Fu et al. [131] developed Fe/Al bimetallic materials producing necklace-like or ball-like Fe particles (200–400 nm) distributed across larger Al particles (20–30 μm). This multiscale structure created a high surface area and multiple reactive zones, promoting both adsorption and catalysis. The finely dispersed Fe particles served as active sites for contaminant binding, while the strong Fe–Al interface supported electron transfer for catalytic activity. In the study by Yang et al. [134], Mg/Fe bimetallic particles formed through galvanic replacement were composed of numerous sheet-like crystal structures. This sheet morphology increased surface exposure, aiding in pollutant degradation. It also supported efficient electron flow, leading to superior electrochemical activity, as demonstrated by a higher reduction current density compared to pure Mg.
The thermogravimetric method has also been noted for effectively producing Ni/Fe bimetallic materials through gas–solid interaction processes. Meshkini Far et al. [152] synthesized Ni/Fe bimetallic catalysts using said method by reducing Fe2O3 and NiO powders in a hydrogen–helium atmosphere at 300 °C for four hours. The resulting material possessed a nanoporous structure, which significantly increased its surface area and exposed active catalytic sites. This porous design facilitated efficient reactant diffusion and enhanced interaction between the catalyst and reactants. The combination of high surface area and synergistic effects between Ni and Fe contributed to the material’s strong catalytic performance. In Liu et al. [157], Ni/Fe bimetallic catalysts were prepared using the same method, starting from Ni–Fe layered double hydroxides. The synthesized material consisted of spherically shaped nanoparticles, yielding a uniform morphology with a high surface-to-volume ratio. This spherical structure improved the accessibility of active sites and supported effective catalytic interactions. De Masi et al. [158] also used the thermogravimetric method to produce Fe/Ni nanoparticles averaging 18.6 ± 2.4 nm, with nickel distributed on their surface. The nanoscale size and surface-exposed Ni enhanced both the catalytic activity and magnetic response, enabling selective CO2-to-methane conversion under low magnetic fields.
Several Fe-based bimetallic materials have also been reported to be synthesized using supports like carbon-based materials, alumina-based materials, silica-based materials, minerals, and various other substrates. These materials are primarily developed for catalytic and adsorptive applications. The following discussion highlights selected studies focusing on their structural characteristics and properties.
In the study by Wang et al. [161], Fe–Ce bimetallic particles were supported on nitrogen-doped carbon nanotubes (NCNTs), forming hollow CNTs that encapsulate nanocrystals. This structure exhibited an ordered distribution of carbon layers with CNT diameters between 100 and 200 nm. The hollow tubular architecture enhanced electron transport and increased accessibility to active sites. As a result, the features significantly facilitated the material’s electrocatalytic performance. In Tian et al. [162], Fe–Al bimetallic particles were supported on multi-walled carbon nanotubes (MWCNTs), producing long, straight nanotubes with a graphite interlayer spacing of 0.34 nm. This well-organized structure improved electronic conductivity by promoting fast electron transfer along the carbon layers. The combination of Fe–Al with MWCNTs also supplied numerous catalytic sites, enhancing catalytic activity. Together, these structural qualities contributed to both favorable catalytic capability and electronic conductivity, making the material ideal for high-rate rechargeable Li-ion battery use. In the work by Wu and Feng [167], Ag–Fe bimetallic nanoparticles averaging 51 nm in diameter were uniformly dispersed on modified biochar (MB), creating small globular formations. This even distribution increased the surface area and exposed more active sites for contaminant interaction. The synergy between Ag and Fe on the biochar enhanced both adsorption and reduction processes, improving removal efficiency. Xing et al. [168], Fe/Ni bimetallic nanoparticles formed chain-like structures distributed throughout the pores and surfaces of biochar (BC). This distinctive morphology expanded the surface area and improved access to active adsorption sites. Additionally, the chain structures imparted magnetic properties, allowing for easy recovery of the material.
In the study by Mutz et al. [179], Ni3Fe nanoparticles were uniformly dispersed on an Al2O3 support, creating a well-distributed bimetallic catalyst structure. This even dispersion maximized the exposure of active sites, enhancing the interaction between reactants and the catalyst surface. The strong interaction between Ni and Fe within the nanoparticles contributed to synergistic effects that improved catalytic performance. The Ni3Fe/Al2O3 nanocatalyst demonstrated notable catalytic capability due to its structure and support integration. In Zhang et al. [182], Pd–Fe nanoparticles were supported on an Al2O3/PVDF membrane, displaying a smooth, spherical morphology with particle sizes between 50 and 100 nm. This uniform and nanoscale structure increased the surface area and enhanced the availability of reactive sites. The strong interaction between Pd and Fe facilitated effective electron transfer, promoting redox reactions essential for contaminant degradation. In the work by Wang et al. [185], Fe and Cu nanoparticles were uniformly dispersed within the hollow mesoporous silica sphere (HMS) support, with an average particle size of around 18 nm. This even distribution enhanced the availability of active metal sites and ensured consistent contact with reactants. The small nanoparticle size contributed to a high surface area, promoting efficient catalytic interactions and electron transfer. The structural features of the FeCu/HMS material directly supported its strong catalytic and degradation capabilities. Additionally, Yan et al. [188] made Fe–Al bimetallic particles supported on SBA-15, forming a structure with well-ordered hexagonal mesopores and one-dimensional channels. This porous architecture enhanced the diffusion of reactants and provided a stable framework for catalytic activity. The observed agglomeration of Fe, indicating the presence of FexOy clusters within the channels, introduced abundant active sites for catalytic reactions.
In the work by Weng et al. [194], Fe/Ni spherical particles ranging from 30 to 60 nm were uniformly dispersed on bentonite. This nanoscale dispersion enhanced the overall surface area and increased the number of accessible active sites. The integration of Fe/Ni with bentonite facilitated both adsorption of contaminants and catalytic interactions due to improved particle contact and reactivity. In Jin et al. [197], Fe/Pd bimetallic particles with diameters ranging from 20 to 70 nm were supported on kaolinite, forming short chain-like spherical structures. This morphology increased the surface area and ensured a uniform distribution of active sites. The nanoscale size and chain-like arrangement promoted effective electron transfer and enhanced interaction with reactants. The structural features of K–Fe/Pd contributed significantly to its catalytic capability. Ezzatahmadi et al. [201] made Fe/Ni spherical nanoparticles (50–80 nm) well dispersed within the pores and on the surface of diatomite, forming a porous Di–Fe/Ni composite. This porous structure increased the accessible surface area and improved the distribution of active catalytic sites. The nanoscale particles enhanced electron transfer and contact with contaminants, facilitating degradation reactions. Similarly, Ezzatahmadi et al. [202] also produced Fe/Ni spherical nanoparticles ranging from 20 to 60 nm in diameter that were well dispersed and stabilized on the surface of palygorskite. This uniform dispersion maximized the exposure of active sites, enhancing catalytic efficiency. The stabilization on the palygorskite support prevented nanoparticle agglomeration, maintaining high reactivity for degradation processes.
Finally, Shi et al. [208] fabricated Fe/Pd bimetallic nanoparticles with a quasi-spherical shape and an average diameter of 10.1 ± 1.7 nm that were uniformly distributed on the surface of filter paper. This homogeneous dispersion ensured the maximum exposure of active sites, promoting efficient interaction with target contaminant. The nanoscale size and even coverage enhanced electron transfer and surface reactivity, which are critical for catalytic functions. As a result, the structural characteristics of the Fe/Pd-assembled filter paper supported its catalytic and reduction capabilities. In the study by Ge et al. [209], Cu and Fe were successfully detected on spherical fly ash (FA) microparticles, indicating the effective loading of the Cu/Fe bimetal. This supported configuration enhanced the distribution of active sites on the FA surface, increasing interaction points with pollutants. The spherical morphology of the FA particles contributed to a high surface area, promoting both adsorption and catalytic reactions.

5. Biological Methods for Synthesizing Iron-Based and Aluminum-Based Bimetals

The biological method for synthesizing metallic nanoparticles offers an eco-friendly, green alternative to traditional physical and chemical methods, providing nontoxic, energy-efficient, and low-cost procedures [39]. Physical methods are energy-intensive and expensive [40,41], while chemical methods often involve harmful solvents and toxic by-products [39]. Biological synthesis can occur through bio-reduction, where metal ions are reduced to stable forms using nontoxic reductants, or by biosorption, where metal ions bind to organisms and are converted into stable nanoparticles [212,213,214]. Table 13 summarizes the reducing agents and stabilizers for the green synthesis of bimetallic particles.
Recently, Fe/Ni, Fe/Pd, and Fe/Cu are amongst the bimetallic particles synthesized biologically. For instance, in the work of Alruqi et al. [215], Fe/Ni core–shell bimetallic nanoparticles were synthesized using a seed-growth co-reduction method with Fe(NO3)3 and Ni(NO3)2, utilizing Pithecellobium dulce legume mesocarp extract as a reducing agent. The extract, prepared by heating powdered mesocarp in deionized water, facilitated the reduction of Fe3+ and Ni2+ ions, confirmed by UV–visible spectroscopy. The nanoparticles were aged, filtered, washed with deionized water and acetone, vacuum-dried, and stored for characterization and batch adsorption experiments. In the study of Lin et al. [216], eucalyptus leaf extracts were prepared by heating crushed leaves in deionized water at 80 °C for 1 h. Fe-Pd nanoparticles (NPs) were synthesized by mixing metal salt solutions with the extract under nitrogen, followed by vacuum filtration, washing, and freeze-drying. Calcined Fe/Pd NPs were prepared by calcination at 500 °C for 4 h in a nitrogen atmosphere. Additionally, Zhang et al. [217] prepared a mixed Fe and Cu solution by combining CuSO4 and FeSO4·7H2O in deionized water, followed by pH adjustment using NaOH and HCl. Green tea extract is added as a reducing agent, resulting in a black solid nanoparticle suspension that is filtered, dried, and used for nitrate removal in wastewater.
Zhu et al. [218] prepared an Fe(II)-Cu(II) solution by dissolving 1.39 g FeSO4·7H2O and 0.18 g CuSO4·5H2O in 100 mL of deionized water, to which green tea extract (50 mL) was added under nitrogen and stirred for 20 min, resulting in a black suspension. The GT-nZVI/Cu suspension was filtered, washed with water and alcohol, and dried overnight in a vacuum oven. In the work of Lin et al. [219], Fe/Ni nanoparticles (NPs) were synthesized using a solution of 0.1 M FeCl3·6H2O and 0.01 M NiCl2·6H2O mixed with eucalyptus leaf extract, prepared by heating dried leaves in deionized water at 80 °C, followed by cooling and filtration. The metal salt solution was added to the extract under nitrogen, stirred, filtered, washed with deionized water and ethanol, and freeze-dried for 48 h to yield Fe/Ni NPs. C–Fe/Ni NPs were obtained by calcining Fe/Ni NPs at 500 °C under a nitrogen atmosphere for 4 h. Additionally, Gopal et al. [220] synthesized Fe/Pd bimetallic nanoparticles by reducing 0.1 M iron (III) chloride with pomegranate peel extract, followed by capping the reduced Fe nanoparticles with 0.1 w/v% carboxymethyl cellulose (CMC) for stabilization. Palladium was introduced by adding various concentrations (5, 10, 15, and 20 mM) of Pd precursor solution to the stabilized Fe nanoparticles, forming Fe/Pd bimetallic nanoparticles. Figure 15 shows the core–shell structure formation of the Fe/Pd bimetal.
Chitosan, recognized for its environmental biodegradability, has gained significant attention in green synthesis processes [221] due to its amine and hydroxyl functional groups, which enable strong binding with Fe0 and effective stabilization of nanoparticles [222]. Chitosan (CS)-stabilized Fe/Cu nanoparticles were synthesized by Jiang et al. [223] for hexavalent chromium remediation by dissolving FeSO4·7H2O in an oxygen-free solution with chitosan (1–4 wt%) under nitrogen, followed by stirring. A NaBH4 solution was added dropwise under mechanical agitation to form zero-valent iron nanoparticles (CS-nZVI), and CuSO4·5H2O was subsequently introduced under vacuum for copper incorporation. The CS-Fe-Cu nanoparticles were washed with ethanol, vacuum-dried at 60 °C for 12 h, and collected for further applications. Cheng et al. [224], on the other hand, used chitosan as a crosslinking agent to load Fe–Al bimetal particles onto bentonite, forming the Fe/Al bimetal chitosan bentonite (Fe–Al bimetal @ bent) complex designed for efficient nitrate removal from wastewater at low temperatures. The Fe/Al bimetal was synthesized by treating aluminum powder with concentrated hydrochloric acid and Fe2SO4 solution, followed by vacuum filtration, while chitosan solution was prepared by dissolving chitosan in 1% nitric acid. The composite was formed by mixing the Fe/Al bimetal with the chitosan solution, adding bentonite, adjusting the pH to 5.6, stirring at 45 °C for 1 h, and then centrifuging and drying the sample overnight in a vacuum oven. Chitosan (CS)-Fe/Ni by Anju et al. [225] was also synthesized when FeCl3·6H2O (0.973 g) and NiSO4·6H2O (0.089 g) solutions were prepared in a 10 mL of deionized water and combined with the chitosan solution under a nitrogen atmosphere. Sodium borohydride (NaBH4, 0.544 g) in an ethanol–water system was added dropwise to the mixture and stirred vigorously for 30 min, forming black CS-Fe precipitates. NiSO4·6H2O solution was subsequently added to CS-Fe and agitated for another 30 min, yielding CS-Fe/Ni nanoparticles (see Figure 16). Additionally, a chitosan–Cu–Fe bimetal complex was also developed by dissolving chitosan in an acidic aqueous solution, followed by the addition of FeCl3·6H2O and CuSO4·5H2O and stirring for 4 h. Ethanol was added to precipitate the complex, which was then washed with ethanol to remove excess reagents and dried at 80 °C. Chitosan hydrogel swelling depends on polymer composition and water interactions, so low-swelling bimetal complexes were preferred for column research and catalytic studies [226].
Alginate, a polysaccharide derived from brown algae, is valued for its functionalized backbone, biodegradability, renewability, and nontoxicity, making it a versatile material for water treatment [227]. It can be combined with materials like chitosan, hydroxyapatite, or activated carbon to create materials for applications in medicine, pharmacy, and environmental protection [228]. Limestone, an abundant and cost-effective adsorbent, effectively removes contaminants like heavy metals, dyes, and pharmaceuticals owing to its heterogeneous surface, buffering capacity, and secondary binding sites [229]. In the work of Ahmed et al. [230], Fe–Cu/Alg–LS nanocomposites were prepared with a mixture of 2% (w/v) sodium alginate and 7 g of limestone in 100 mL distilled water, stirred, and heated to 80 °C for homogenization. A solution of ZVI/Cu (0.5 Fe–0.5 Cu in 100 mL) was added to the mixture, followed by adding 0.3 M calcium chloride via syringe injection to form beads. The beads were hardened by submerging them in calcium chloride for 12 h.
Cellulose is a highly renewable biopolymer found in plant cell walls, some algae, and bacteria [231] and has been widely studied for its applications in various fields [232]. Cellulose-based materials are ideal for immobilizing metal nanoparticles and have been explored for their use in organic transformations and catalytic applications [233,234,235]. In the study of Karami et al. [236], bimetallic Fe/Cu nanoparticles were successfully immobilized on microcrystalline cellulose (MCC) to create an efficient, magnetically recoverable nanocatalyst for the NaBH4 reduction of nitroarenes to arylamines in water. Cellulose was sonicated in deionized water, then mixed with an FeCl2·4H2O solution and sonicated further, followed by stirring under nitrogen to eliminate dissolved oxygen. NaBH4 was added to the mixture under inert conditions, immobilizing iron nanoparticles (Fe NPs) on microcrystalline cellulose (Fe@MCC), which was separated using a magnet, washed, and dried. The Fe@MCC nanocomposite was then treated with a CuCl2·2H2O solution, and metallic copper nanoparticles (Cu NPs) were formed by reducing Cu2+ ions with NaBH4, resulting in the Fe–Cu@MCC nanocomposite, which was washed, dried, and collected.
Table 13. Biological methods for bimetal synthesis.
Table 13. Biological methods for bimetal synthesis.
Bimetal SystemExperimental Materials AdvantagesDisadvantagesReference
Fe-NiMetal precursors: Fe(NO3)3, Ni(NO3)2
Reducing agent: Pithecellobium dulce legume mesocarp extract
Eco-friendly synthesis
Simple and cost-effective
Efficient metal ion reduction
Surfactant-assisted stability
Controlled nanoparticle formation
Process variability
Longer preparation time (multi-step process)
[215]
Fe-Pd Metal precursors: FeCl3·6H2O, PdCl2
Reducing agent: Eucalyptus leaf extract
Green synthesis approach
Simple process
Enhanced stability
Reduced contamination
Variability in leaf extract composition
Time-intensive preparation
Energy consumption
Limited reduction efficiency
[216]
Fe-Cu Metal precursors: FeSO4·7H2O, CuSO4
Reducing agent: Green tea extract
Eco-friendly synthesis
Controlled copper loading
Improved stability
Simple and scalable process
Time-consuming process
Batch-to-batch variability
Energy consumption
Limited control over particle size
[217]
Fe-Cu Metal precursors: FeSO4·7H2O, CuSO4·5H2O
Reducing agent: Green tea extract
Green synthesis approach
Simple and efficient process
Potential batch variability
Energy consumption
Limited control over particle size
[218]
C-Fe-Ni Metal precursors: FeCl3·6H2O, NiCl2·6H2O
Reducing agent: Eucalyptus leaf extract
Eco-friendly synthesis
Cost-effective materials
Simple preparation method
Improved stability via calcination
Time-intensive process
Batch-to-batch variability
Energy-intensive steps
Potential agglomeration
[219]
Fe-PdMetal precursors:
Fe (III) chloride, Potassium hexachloropalladate (IV)
Reducing agent: Pomegranate peel extract
Eco-friendly synthesis
Cost-effective and sustainable process
Enhanced stability
Controlled bimetallic composition
Time-intensive synthesis
Batch-to-batch variability
Energy consumption
[220]
Chitosan(CS)-stabilized Fe-CuMetal precursors: FeSO4·7H2O, CuSO4·5H2O
Reducing agent: NaBH4
Stabilizing agent: Chitosan
Enhanced stability
Controlled cu loading
Efficient reduction process
Oxygen-free synthesis
Energy-intensive process
Complex synthesis procedure
[223]
Fe–Al bimetal chitosan bentonite (Fe–Al bimetal@bent) complexMetal precursors: Fe2SO4, Al powder
Stabilizing agents: Chitosan, Na-bentonite
Simple synthesis process
Effective ph control
Improved structural stability
Vacuum-drying enhancing purity
Use of concentrated acid
Energy and time-intensive drying process
Centrifugation step complexity
[224]
Chitosan (CS)-Fe-NiMetal precursors: FeCl3·6H2O, NiSO4·6H2O
Reducing agent: NaBH4
Stabilizing agent: Chitosan
Green synthesis approach
Controlled reduction process
Effective metal loading
Lyophilization for long-term stability
Intricate synthesis method
Labor-intensive process
[225]
Chitosan–Cu–Fe bimetal complexMetal precursors: FeCl3·6H2O, NiSO4·6H2O
Reducing agent: NaBH4
Stabilizing agent: Chitosan
Simple and efficient synthesis
Good metal loading control
Improved swelling properties
Chitosan involving acid dissolution
Limited structural control
[226]
ZVFe–Cu/Alg–LSMetal precursors: FeSO4·7H2O, CuSO4·5H2O
Stabilizing agents: Sodium alginate, limestone
Environmentally friendly synthesis
Improved stability
Controlled release of zero-valent iron (ZVI) and copper
Multi-step process
Requires multiple washing steps
[230]
Fe–Cu@MCCMetal precursors: FeCl2·4H2O, CuCl2·2H2O
Reducing agent: NaBH4
Stabilizing agent: Microcrystalline cellulose (MCC)
Magnetic recoverability of the catalysts
Support material (MCC) improves stability
MCC as biodegradable and nontoxic support
Complex multi-step process [236]

Influence of Characteristics of Synthesized Bimetals by Biological Methods to Their Properties

Over the past decade, a variety of bio-reducing agents and bio-based stabilizers have been utilized in the synthesis of mostly Fe-based bimetallic materials and composites. These materials were predominantly nanoscale in size and were primarily applied in adsorption and catalysis-related processes for environmental remediation purposes. Table 14 presents the various characteristics and properties of bimetals using some bio-reducing agents and stabilizers.
Green tea extracts have been primarily utilized in the synthesis of Fe/Cu bimetals. In the study by Zhang et al. [217], Fe-Cu bimetallic materials were synthesized using green tea extract as a reducing agent, producing spherically shaped nanoparticles. This spherical morphology provided a high surface area, which enhanced contact with nitrate pollutants in wastewater. The increased surface interaction contributed to a strong adsorptive capacity, enabling effective pollutant removal. Additionally, the uniform shape promoted efficient electron transfer between Fe and Cu, thereby improving the reduction capability essential for nitrate remediation. Similarly, Zhu et al. [218] employed green tea extract to synthesize Fe/Cu nanoparticles with spherical morphology, which provided a high surface area and uniform dispersion in solution. These features significantly enhanced the material’s adsorptive performance, making it well suited for environmental remediation purposes.
Eucalyptus leaf extracts have also been utilized for bimetallic synthesis. In the study by Lin et al. [216], Fe/Pd bimetallic nanoparticles were synthesized resulting in spherically shaped nanoparticles. This morphology provided a high surface area, facilitating effective contact with target molecules. The uniform spherical structure enhanced the material’s catalytic capability, promoting efficient reaction kinetics. Additionally, the increased surface area and surface activity contributed to a strong adsorptive capacity, making the material suitable for environmental and catalytic applications. Also, in Lin et al. [219], calcined (C) Fe/Ni bimetallic materials were also developed resulting in polydisperse regular spherical nanoparticles. This uniform yet varied morphology provided a high surface area and multiple active sites, enhancing the material’s interaction with reactants. The calcination process further improved crystallinity and stability, which are essential for catalytic performance.
Additional bio-based reducing agents used include extracts from Pithecellobium dulce legume mesocarp and pomegranate peel. In the study by Alruqi et al. [215], Fe/Ni bimetallic materials were synthesized employing Pithecellobium dulce legume mesocarp extract as a natural reducing agent. The resulting materials exhibited a core–shell structure and nanospheres with some irregularly shaped nanoparticles, which provided a high surface-to-volume ratio beneficial for adsorption processes. This morphology enhanced the material’s ability to interact with and capture contaminants, contributing to its notable adsorptive capacity. The combination of nanoscale features and bio-assisted synthesis also promoted eco-friendliness and functional efficiency in environmental applications. Furthermore, in Gopal et al. [220], Fe/Pd bimetallic materials were also used through pomegranate peel extract, resulting in spherically shaped nanoparticles with a core–shell structure. The spherical morphology provided a large surface area, promoting efficient catalytic interactions. The core–shell architecture enhanced electron transfer between the Fe and Pd components, which contributed to the material’s catalytic performance. Together, these structural features enabled the Fe-Pd nanoparticles to exhibit strong catalytic capability, suitable for environmental applications.
For bio-based stabilizers, chitosan was used primarily in bimetallic synthesis. For instance, Jiang et al. [223] developed Fe-Cu bimetallic nanoparticles stabilized with chitosan, forming spherically shaped nanoparticles. This spherical morphology provided a high surface area, which enhanced the nanoparticles’ interaction with hexavalent chromium contaminants. The uniform shape and stabilization by chitosan prevented particle agglomeration, promoting consistent dispersion in aqueous environments. As a result, the material demonstrated a strong catalytic capability, making it effective for the remediation of hexavalent chromium. Another study by Anju et al. [225] also prepared CS-Fe-Ni bimetallic materials using chitosan as a stabilizer, forming nanoparticles with an Fe core and a chitosan shell. This core–shell structure provided structural integrity and enhanced stability, preventing particle aggregation. The nanoscale size increased the surface area available for reactions, improving interaction with target molecules. As a result, the material exhibited a strong catalytic capability, making it effective for applications that require efficient and sustained catalytic activity.
Moreover, Cheng et al. [224] prepared Fe–Al bimetal@bent complexes using chitosan as a crosslinking agent to load Fe–Al bimetallic particles onto bentonite. The structure featured abundant Fe–Al bimetals encapsulated within a Cs-bentonite matrix, enhancing particle stability and dispersion. This configuration provided a high surface area and accessible active sites, promoting efficient interaction with nitrate ions. As a result, the material demonstrated a strong adsorptive capability, making it suitable for effective nitrate removal from wastewater, even at low temperatures. In the study by Rashid et al. [226], a chitosan–Cu–Fe bimetal complex was also synthesized, resulting in an irregular and relatively nonporous structure. The incorporation of Cu and Fe within the chitosan matrix facilitated the formation of a stable bimetallic system. Despite its nonporous nature, the irregular morphology allowed sufficient exposure of the active metal sites. This structural configuration contributed to the material’s catalytic capability, enabling it to participate effectively in catalytic processes.
Other bio-stabilizers were used as well such as alginate, limestone, and cellulose. In the study by Ahmed et al. [230], a ZVFe–Cu/Alg–LS bimetallic nanocomposite was synthesized using alginate and limestone (Alg–LS) as stabilizers, resulting in a multilayer structure with a rough surface. This multilayered architecture provided an extensive surface area and numerous active sites, which enhanced the material’s ability to interact with contaminants. The rough surface further promoted efficient adsorption by increasing surface roughness and contact points. Consequently, the nanocomposite demonstrated both strong adsorptive and catalytic capabilities, making it effective for environmental remediation applications. Further, in Karami et al. [236], Fe–Cu@MCC nanocomposite particles were created by immobilizing Fe–Cu nanoparticles onto microcrystalline cellulose (MCC), resulting in particle sizes ranging from 27 to 35 nm. The nanoscale size provided a high surface area, enhancing the exposure of active sites for catalytic reactions. The integration with MCC also contributed to structural stability and facilitated dispersion, supporting consistent catalytic performance. Additionally, the presence of iron conferred the material with a magnetic property, enabling easy magnetic recovery from solution after use.

6. Summary and Future Directions

This work provides a detailed overview of methods for synthesizing Fe-based and Al-based bimetals. Following PRISMA guidelines, 122 studies from 2014 to 2023 were systematically and bibliometrically reviewed and summarized. The synthesis of Fe-Al bimetals involved physical, chemical, and biological methods.
The research on the synthesis of iron-based and aluminum-based bimetals initially declined but saw a sharp increase in 2018, followed by a stable trend in recent years, with the last five years accounting for 50% of total studies. China led in research publications (62.3%), followed by Russia (5.7%), Australia (4.9%), and India (4.9%), while Saudi Arabia ranked highest in the number of citations per document (95.0). RSC Advances was the most active journal, with Applied Catalysis B: Environmental having the highest number of citations per document (203.0), and all listed journals ranked Q1 in impact. Tongji University (China) was the leading institution, with Chinese institutions dominating, while A. Sharipova (Israel/Russia) ranked as the most active author. Chemical synthesis methods dominated, particularly supported particles (41 studies), galvanic replacement (23 studies), and chemical reduction (15 studies), while biological and physical methods were gaining traction. The most commonly synthesized bimetals were Fe/Cu (20 studies), Fe/Al (16 studies), and Fe/Ni (16 studies), with Fe/Ag and other noble metal-based bimetals appearing in fewer studies.
Physical methods include mechanical alloying, the electrical explosion of metal wires, radiolysis, sonochemical methods, and MF-LAL methods. Mechanical alloying remains an effective method for producing bimetallic composite powders and alloys using high-energy milling to achieve fine distribution and nanometer-scale refinement, with applications in magnetic and biomedical fields. Optimizing process parameters (e.g., milling time, speed) may reduce energy consumption without compromising quality. The electrical explosion of dissimilar metal wires is an effective recent synthesis method for producing bimetallic nanoparticles with tunable compositions and properties. Future research may consider optimizing electrical explosion parameters, exploring a wider range of metal combinations, and conducting performance evaluations to better correlate synthesis conditions with the structure and functional properties of bimetallic nanoparticles. Femtosecond laser irradiation enables the synthesis of Fe-Pt bimetallic nanoparticles from aqueous metal salt solutions without chemical reducing agents, resulting in particles with enhanced surface reactivity. The addition of polyvinylpyrrolidone (PVP) further improves dispersibility and size control, emphasizing the impact of synthesis conditions on the nanoparticles’ characteristics and properties. Future undertakings may focus on optimizing radiolysis synthesis parameters, exploring alternative dispersing agents, and evaluating the performance of Fe/Pt bimetallic nanoparticles to better understand the relationships between synthesis methods, material characteristics, and their practical applications. The sonochemical method has been effectively used to produce Fe-based bimetallic materials with favorable magnetic properties and to fabricate porous bimetallic structures for energy and catalytic applications. Future studies may consider optimizing sonochemical synthesis parameters, such as, but not limited to, ultrasonic power, precursor concentration, pH, and reaction time, to achieve more uniform particle size, morphology, and compositional control in bimetallic nanoparticles. The MF-LAL method produced Fe-based bimetallic nanochains (e.g., Fe/Pt, Fe/Co, and Fe/Ni) with one-dimensional morphology and strong ferromagnetic properties—characterized by high saturation magnetization, low coercivity, and low remanence—making them ideal for advanced magnetic applications. Future research may also consider refining MF-LAL synthesis parameters including, but not limited to, laser pulse energy, irradiation duration, and magnetic field strength, while exploring various metal pairings and solvent systems to achieve more precise control over the morphology, magnetic alignment, and structural stability of bimetallic nanochains.
In comparison, the chemical techniques are chemical reduction, dealloying, seed-mediated growth, electrochemical deposition, galvanic replacement, the thermogravimetric method, and the supported particle method. Chemical reduction is a widely recognized synthesis method for producing bimetallic materials that combine adsorptive and catalytic properties with magnetic recoverability for improved reusability. Future studies may be conducted to investigate the optimization of chemical reduction parameters including, but not limited to, the types of metal salts, concentrations of reducing agents, and composition of the reaction media, to potentially reduce synthesis time while enhancing control over the morphology, composition, and structural arrangement of bimetallic particles. Chemical dealloying has been employed to synthesize nanoporous bimetallic materials with electrocatalytic capabilities. Future research may consider optimizing dealloying parameters such as, but not limited to alloy composition, etchant concentration, and treatment duration, to achieve better control over pore size distribution and structural uniformity in bimetallic materials. Investigation of alternative precursor alloys and the use of more environmentally friendly dealloying agents could support the development of more sustainable synthesis methods while maintaining desirable functional properties. Seed-mediated growth is also a method for synthesizing bimetallic particles with catalytic and degradation capabilities. Future research efforts may benefit from optimizing synthesis parameters including, but not limited to, metal precursor concentrations, reaction conditions, and deposition methods—to enhance control over the particle size, morphology, and overall structure of bimetallic materials. Electrodeposition techniques allow control over the morphology and composition of magnetic Fe/Rh nanowires and Fe/Zn dendritic structures. Future studies may consider refining key parameters—such as, but not limited to, voltage, frequency, and electrolyte formulation—to further enhance the precision in shaping the structural, compositional, and dimensional features of these bimetallic materials.
Galvanic replacement is an established method for synthesizing Fe/Cu, Fe/Al, and Mg/Fe bimetallic materials, enabling controlled deposition of secondary metals onto primary metal surfaces. Future investigations may focus on optimizing synthesis conditions such as, but not limited to, metal precursor ratios, pH levels, reaction temperatures, and agitation rates—to improve the uniformity and distribution of the secondary metal. Additionally, utilizing aluminum scraps and alloys as a raw material for Fe/Al bimetal synthesis presents a cost-efficient and sustainable approach. The thermogravimetric method is a reliable technique for synthesizing Ni/Fe bimetallic materials via gas–solid interactions. Upcoming research may refine synthesis strategies including, but not limited to, co-precipitation, thermal decomposition, and gas-phase reduction, to achieve better control over the particle size, elemental composition, and structural morphology of the resulting bimetallic materials. Various Fe-based bimetallic materials have been synthesized using a variety of support materials, including carbon-based, alumina-based, and silica-based substrates, minerals, etc. These supported bimetallic systems are mainly designed for use in catalytic and adsorption-related applications. Future studies may refine synthesis methods for supported bimetallic materials by exploring advanced techniques to achieve uniform particle dispersion and prevent agglomeration. Optimizing the process flow may also reduce both synthesis time and operational costs, further enhancing the practicality of these methods.
Biological synthesis methods are green approaches that have shown promise in fabricating bimetallic nanoparticles and nanocomposites using plant extracts, biopolymers, and natural supports. These methods offer environmentally friendly alternatives that yield bimetallic materials with desirable structures, enhancing properties such as stability, reactivity, and adsorption capacity. The resulting materials demonstrate potential for various applications, including wastewater treatment, catalysis, and environmental remediation, owing to their tunable morphology and surface chemistry. Future work could focus on optimizing green synthesis methods by exploring a wider range of bio-reducing agents, bio-based stabilizers, and process parameters to achieve better control over particle size, morphology, and uniformity, thereby enhancing functional properties and reducing the time-consuming preparation processes.

Author Contributions

Conceptualization, J.K.B.B., C.B.T. and V.J.T.R.; methodology, J.K.B.B., C.B.T., T.P., J.B.Z., T.A., I.P., W.M., M.I., R.D.A., S.J., K.H. and V.J.T.R.; formal analysis, J.K.B.B., C.B.T., T.P., J.B.Z., T.A., I.P., W.M., M.I., R.D.A., A.H.O., A.B.B., M.A.B.P., S.J., K.H. and V.J.T.R.; resources, T.P., T.A., I.P., W.M., M.I., R.D.A., A.H.O., A.B.B., M.A.B.P., S.J. and K.H.; data curation, J.K.B.B., C.B.T. and T.P.; writing—original draft preparation, J.K.B.B.; writing—review and editing, J.K.B.B., C.B.T., T.P., J.B.Z., T.A., I.P., W.M., M.I., R.D.A., A.H.O., A.B.B., M.A.B.P., S.J., K.H. and V.J.T.R.; visualization, J.K.B.B.; supervision, C.B.T., R.D.A. and V.J.T.R.; project administration, C.B.T., J.B.Z. and V.J.T.R.; funding acquisition, I.P., A.H.O., A.B.B., M.A.B.P. and V.J.T.R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors wish to thank the Department of Science and Technology-Philippine Council for Industry, Energy, and Emerging Technology Research and Development (DOST-PCIEERD) for funding HyFIBE Program Project 1 (PCIEERD Project No. 9546).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors also thank the DOST-Engineering Research and Development for Technology (DOST-ERDT) program for providing Jeffrey Ken B. Balangao’s scholarship.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ZVMsZero-valent metals
ZVIZero-valent iron
HOCsHalogenated organic compounds
TCETrichloroethylene
PCPPentachlorophenol
ZVAl Zero-valent aluminum
AMDAcid mine drainage
MF-LALMagnetic field-assisted laser ablation in liquid
PVPPolyvinylpyrrolidone
TCDTrisodium citrate dehydrate
PSTTPotassium sodium tartrate tetrahydrate
EDTADisodium ethylenediaminetetraacetate dehydrate
EnEthylenediamine
TEATriethanolamine
AASAtomic absorption spectroscopy
ORROxygen reduction reaction
LDHLayered double hydroxide
CNTsCarbon nanotubes
MWCNTsMulti-walled carbon nanotubes
CNFCarbon nanofibers
PANPolyacrylonitrile
DMFDimethylformamide
MCMesoporous carbon
MBModified biochar
BCBiochar
FMBCFe–Co-modified biochar
nZVIC-SBCFe-Cu-municipal sludge-derived biochar nanoparticles
ZF@CBCZn/Fe nanoparticles on corncob biochar (CBC)
ACActivated carbon
PACPowder-activated carbon
CACCommercial activated carbon
ALDAtomic layer deposition
PVDFPolyvinylidene difluoride
HMSHollow mesoporous silica sphere
TMOSTetramethyl orthosilicate
CECCation exchange capacity
Be@Fe-CuFe-Cu on bentonite
B-Fe/NiFe/Ni on bentonite
K-Fe/PdK-Fe/Pd on kaolinite
Cu/Fe@zeoliteCu/Fe on zeolite
Di-Fe/NiFe/Ni on diatomite
Pal-Fe/NiFe/Ni on palygorskite
Sep-Fe/NiFe/Ni on sepiolite
ASF@NCN-doped carbon layer functionalized on aluminum silicate fibers
PDAPolydopamine
PAMPolyacrylamide
PEIPolyethylenimine
Cu/Fe-BM@FACu/Fe bimetallic modified fly ash
GT-nZVI/CuGreen tea extract-based nZVI/Cu particles
C–Fe/Ni NPsCalcined Fe/Ni nanoparticles
CMC Carboxymethyl cellulose
CSChitosan
CS-Fe-CuChitosan (CS)-stabilized Fe/Cu
Fe–Al bimetal @ bentFe–Al bimetal chitosan bentonite complex
Alg–LSAlginate–limestone
Fe–Cu@MCCFe-Cu immobilized on microcrystalline cellulose

References

  1. Yan, Z.; Ouyang, J.; Wu, B.; Liu, C.; Wang, H.; Wang, A.; Li, Z.A.B. Nonmetallic modified zero-valent iron for remediating halogenated organic compounds and heavy metals: A comprehensive review. Environ. Sci. Ecotechnol. 2024, 21, 100417. [Google Scholar] [CrossRef] [PubMed]
  2. Silwamba, M.; Ito, M.; Tabelin, C.B.; Park, I.; Jeon, S.; Takada, M.; Kubo, Y.; Hokari, N.; Tsunekawa, M.; Hiroyoshi, N. Simultaneous extraction and recovery of lead using citrate and micro-scale zero-valent iron for decontamination of polluted shooting range soils. Environ. Adv. 2021, 5, 100115. [Google Scholar] [CrossRef]
  3. Silwamba, M.; Ito, M.; Hiroyoshi, N.; Tabelin, C.B.; Fukushima, T.; Park, I.; Jeon, S.; Igarashi, T.; Sato, T.; Nyambe, I.; et al. Detoxification of lead-bearing zinc plant leach residues from Kabwe, Zambia by coupled extraction-cementation method. J. Environ. Chem. Eng. 2020, 8, 104197. [Google Scholar] [CrossRef]
  4. Silwamba, M.; Ito, M.; Hiroyoshi, N.; Tabelin, C.B.; Hashizume, R.; Fukushima, T.; Park, I.; Jeon, S.; Igarashi, T.; Sato, T.; et al. Alkaline leaching and concurrent cementation of dissolved Pb and Zn from zinc plant leach residues. Minerals 2022, 12, 393. [Google Scholar] [CrossRef]
  5. Seng, S.; Tabelin, C.B.; Kojima, M.; Hiroyoshi, N.; Ito, M. Galvanic microencapsulation (GME) using zero-valent aluminum and zero-valent iron to suppress pyrite oxidation. Mater. Trans. 2019, 60, 277–286. [Google Scholar] [CrossRef]
  6. Seng, S.; Tabelin, C.B.; Makino, Y.; Chea, M.; Phengsaart, T.; Park, I.; Hiroyoshi, N.; Ito, M. Improvement of flotation and suppression of pyrite oxidation using phosphate-enhanced galvanic microencapsulation (GME) in a ball mill with steel ball media. Miner. Eng. 2019, 143, 105931. [Google Scholar] [CrossRef]
  7. Choi, S.; Jeon, S.; Park, I.; Tabelin, C.B.; Ito, M.; Hiroyoshi, N. Enhanced cementation of Cd2+, Co2+, Ni2+, and Zn2+ on Al from sulfate solutions by activated carbon addition. Hydrometallurgy 2021, 201, 105580. [Google Scholar] [CrossRef]
  8. Jeon, S.; Bright, S.; Park, I.; Tabelin, C.B.; Ito, M.; Hiroyoshi, N. A simple and efficient recovery technique for gold ions from ammonium thiosulfate medium by galvanic interactions of zero-valent aluminum and activated carbon: A parametric and mechanistic study of cementation. Hydrometallurgy 2022, 208, 105815. [Google Scholar] [CrossRef]
  9. Chiu, P.C. Applications of zero-valent iron (ZVI) and nanoscale ZVI to municipal and decentralized drinking water systems—A review. Nov. Solut. Water Pollut. 2013, 237–249. [Google Scholar] [CrossRef]
  10. Fu, F.; 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]
  11. Mueller, N.C.; Nowack, B. Nanoparticles for remediation: Solving big problems with little particles. Elements 2010, 6, 395–400. [Google Scholar] [CrossRef]
  12. Zhao, X.; Liu, W.; Cai, Z.; Han, B.; Qian, T.; Zhao, D. An overview of preparation and applications of stabilized zero-valent iron nanoparticles for soil and groundwater remediation. Water Res. 2016, 100, 245–266. [Google Scholar] [CrossRef] [PubMed]
  13. Jeon, S.; Ito, M.; Tabelin, C.B.; Pongsumrankul, R.; Tanaka, S.; Kitajima, N.; Saito, A.; Park, I.; Hiroyoshi, N. A physical separation scheme to improve ammonium thiosulfate leaching of gold by separation of base metals in crushed mobile phones. Miner. Eng. 2019, 138, 168–177. [Google Scholar] [CrossRef]
  14. Jeon, S.; Tabelin, C.B.; Park, I.; Nagata, Y.; Ito, M.; Hiroyoshi, N. Ammonium thiosulfate extraction of gold from printed circuit boards (PCBs) of end-of-life mobile phones and its recovery from pregnant leach solution by cementation. Hydrometallurgy 2020, 191, 105214. [Google Scholar] [CrossRef]
  15. Chen, S.Y.; Chen, W.H.; Shih, C.J. Heavy metal removal from wastewater using zero-valent iron nanoparticles. Water Sci. Technol. 2008, 58, 1947–1954. [Google Scholar] [CrossRef]
  16. Wang, C.B.; Zhang, W.X. Synthesizing nanoscale iron particles for rapid and complete dechlorination of TCE and PCBs. Environ. Sci. Technol. 1997, 31, 2154–2156. [Google Scholar] [CrossRef]
  17. Cheng, S.F.; Wu, S.C. The enhancement methods for the degradation of TCE by zero-valent metals. Chemosphere 2000, 41, 1263–1270. [Google Scholar] [CrossRef]
  18. Sun, Y.; Li, J.; Huang, T.; Guan, X. The influences of iron characteristics, operating conditions and solution chemistry on contaminants removal by zero-valent iron: A review. Water Res. 2016, 100, 277–295. [Google Scholar] [CrossRef]
  19. Zaleska-Medynska, A.; Marchelek, M.; Diak, M.; Grabowska, E. Noble metal-based bimetallic nanoparticles: The effect of the structure on the optical, catalytic and photocatalytic properties. Adv. Colloid Interface Sci. 2016, 229, 80–107. [Google Scholar] [CrossRef]
  20. Rodriguez, J.A.; Goodman, D.W. The nature of the metal-metal bond in bimetallic surfaces. Science 1992, 257, 897–903. [Google Scholar] [CrossRef]
  21. Kitchin, J.R.; Nørskov, J.K.; Barteau, M.A.; Chen, J.G. Modification of the surface electronic and chemical properties of Pt(111) by subsurface 3d transition metals. J. Chem. Phys. 2004, 120, 10240–10246. [Google Scholar] [CrossRef] [PubMed]
  22. Guan, X.; Sun, Y.; Qin, H.; Li, J.; Lo, I.M.; He, D.; Dong, H. The limitations of applying zero-valent iron technology in contaminants sequestration and the corresponding countermeasures: The development in zero-valent iron technology in the last two decades (1994–2014). Water Res. 2015, 75, 224–248. [Google Scholar] [CrossRef]
  23. Huo, X.; Zhou, P.; Liu, Y.; Cheng, F.; Liu, Y.; Cheng, X.; Zhang, Y.; Wang, Q. Removal of contaminants by activating peroxymonosulfate (PMS) using zero valent iron (ZVI)-based bimetallic particles (ZVI/Cu, ZVI/Co, ZVI/Ni, and ZVI/Ag). RSC Adv. 2020, 10, 28232–28242. [Google Scholar] [CrossRef]
  24. Nidheesh, P.V.; Khatri, J.; Singh, T.A.; Gandhimathi, R.; Ramesh, S.T. Review of zero-valent aluminium based water and wastewater treatment methods. Chemosphere 2018, 200, 621–631. [Google Scholar] [CrossRef]
  25. Fu, F.; Cheng, Z.; Lu, J. Synthesis and use of bimetals and bimetal oxides in contaminants removal from water: A review. RSC Adv. 2015, 5, 85395–85409. [Google Scholar] [CrossRef]
  26. Sankar, M.; Dimitratos, N.; Miedziak, P.J.; Wells, P.P.; Kiely, C.J.; Hutchings, G.J. Designing bimetallic catalysts for a green and sustainable future. Chem. Soc. Rev. 2012, 41, 8099–8139. [Google Scholar] [CrossRef]
  27. Yu, W.; Porosoff, M.D.; Chen, J.G. Review of Pt-based bimetallic catalysis: From model surfaces to supported catalysts. Chem. Rev. 2012, 112, 5780–5817. [Google Scholar] [CrossRef]
  28. Bokare, A.D.; Choi, W. Zero-valent aluminum for oxidative degradation of aqueous organic pollutants. Environ. Sci. Technol. 2009, 43, 7130–7135. [Google Scholar] [CrossRef]
  29. Kim, J.; Choi, H.; Kim, D.; Park, J.Y. Operando surface studies on metal-oxide interfaces of bimetal and mixed catalysts. ACS Catal. 2021, 11, 8645–8677. [Google Scholar] [CrossRef]
  30. Scaria, J.; Nidheesh, P.V.; Kumar, M.S. Synthesis and applications of various bimetallic nanomaterials in water and wastewater treatment. J. Environ. Manag. 2020, 259, 110011. [Google Scholar] [CrossRef]
  31. Quiton, K.G.N.; Lu, M.C.; Huang, Y.H. Synthesis and catalytic utilization of bimetallic systems for wastewater remediation: A review. Chemosphere 2021, 262, 128371. [Google Scholar] [CrossRef]
  32. Liu, W.J.; Qian, T.T.; Jiang, H. Bimetallic Fe nanoparticles: Recent advances in synthesis and application in catalytic elimination of environmental pollutants. Chem. Eng. J. 2014, 236, 448–463. [Google Scholar] [CrossRef]
  33. Aryee, A.A.; Liu, Y.; Han, R.; Qu, L. Bimetallic adsorbents for wastewater treatment: A review. Environ. Chem. Lett. 2023, 21, 1811–1835. [Google Scholar] [CrossRef]
  34. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G.; The PRISMA Group. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. Int. J. Surg. 2010, 8, 336–341. [Google Scholar] [CrossRef]
  35. Andrews, R. The place of systematic reviews in education research. Br. J. Educ. Stud. 2005, 53, 399–416. [Google Scholar] [CrossRef]
  36. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. bmj 2021, 372, n71. [Google Scholar] [CrossRef]
  37. James, S.L.; Adams, C.J.; Bolm, C.; Braga, D.; Collier, P.; Friščić, T.; Grepioni, F.; Harris, K.D.M.; Hyett, G.; Jones, W.; et al. Mechanochemistry: Opportunities for new and cleaner synthesis. Chem. Soc. Rev. 2012, 41, 413–447. [Google Scholar] [CrossRef]
  38. Imawaka, K.; Sugita, M.; Takewaki, T.; Tanaka, S. Mechanochemical synthesis of bimetallic CoZn-ZIFs with sodalite structure. Polyhedron 2019, 158, 290–295. [Google Scholar] [CrossRef]
  39. Thakkar, K.N.; Mhatre, S.S.; Parikh, R.Y. Biological synthesis of metallic nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 257–262. [Google Scholar] [CrossRef]
  40. Chen, J.C.; Lin, Z.H.; Ma, X.X. Evidence of the production of silver nanoparticles via pretreatment of Phoma sp. 3.2883 with silver nitrate. Lett. Appl. Microbiol. 2003, 37, 105–108. [Google Scholar] [CrossRef]
  41. Li, Y.; Duan, X.; Qian, Y.; Yang, L.; Liao, H. Nanocrystalline silver particles: Synthesis, agglomeration, and sputtering induced by electron beam. J. Colloid Interface Sci. 1999, 209, 347–349. [Google Scholar] [CrossRef] [PubMed]
  42. Benjamin, J.S. Dispersion strengthened superalloys by mechanical alloying. Metall. Trans. 1970, 1, 2943–2951. [Google Scholar] [CrossRef]
  43. Lyu, H.; Gao, B.; He, F.; Ding, C.; Tang, J.; Crittenden, J.C. Ball-milled carbon nanomaterials for energy and environmental applications. ACS Sustain. Chem. Eng. 2017, 5, 9568–9585. [Google Scholar] [CrossRef]
  44. Wang, P.; Hu, J.; Liu, T.; Han, G.; Ma, W.M.; Li, J. New insights into ball-milled zero-valent iron composites for pollution remediation: An overview. J. Clean. Prod. 2023, 385, 135513. [Google Scholar] [CrossRef]
  45. Ai, D.; Wei, T.; Meng, Y.; Chen, X.; Wang, B. Ball milling sulfur-doped nano zero-valent iron@ biochar composite for the efficient removal of phosphorus from water: Performance and mechanisms. Bioresour. Technol. 2022, 357, 127316. [Google Scholar] [CrossRef]
  46. Gao, J.; Wang, W.; Rondinone, A.J.; He, F.; Liang, L. Degradation of trichloroethene with a novel ball milled Fe–C nanocomposite. J. Hazard. Mater. 2015, 300, 443–450. [Google Scholar] [CrossRef]
  47. Gu, Y.; Wang, B.; He, F.; Bradley, M.J.; Tratnyek, P.G. Mechanochemically sulfidated microscale zero valent iron: Pathways, kinetics, mechanism, and efficiency of trichloroethylene dechlorination. Environ. Sci. Technol. 2017, 51, 12653–12662. [Google Scholar] [CrossRef]
  48. Friščić, T.; Mottillo, C.; Titi, H.M. Mechanochemistry for synthesis. Angew. Chem. 2020, 132, 1030–1041. [Google Scholar] [CrossRef]
  49. Hawili, A.A.; Ghommem, M.; Alami, A.H.; Alasad, S.; Egilmez, M.; Zaid, W.A. Utilizing aluminum sheets with FeCu deposits as cheap water cleaning electrodes. Appl. Surf. Sci. Adv. 2022, 7, 100193. [Google Scholar] [CrossRef]
  50. Vishlaghi, M.B.; Ataie, A. Investigation on solid solubility and physical properties of Cu–Fe/CNT nano-composite prepared via mechanical alloying route. Powder Technol. 2014, 268, 102–109. [Google Scholar] [CrossRef]
  51. Schoenitz, M.; Dreizin, E.L. Structure and properties of Al–Mg mechanical alloys. J. Mater. Res. 2003, 18, 1827–1836. [Google Scholar] [CrossRef]
  52. Sharipova, A.; Swain, S.K.; Gotman, I.; Starosvetsky, D.; Psakhie, S.G.; Unger, R.; Gutmanas, E.Y. Mechanical, degradation and drug-release behavior of nano-grained Fe-Ag composites for biomedical applications. J. Mech. Behav. Biomed. Mater. 2018, 86, 240–249. [Google Scholar] [CrossRef]
  53. Sharipova, A.; Gotman, I.; Psakhie, S.G.; Gutmanas, E.Y. Biodegradable nanocomposite Fe–Ag load-bearing scaffolds for bone healing. J. Mech. Behav. Biomed. Mater. 2019, 98, 246–254. [Google Scholar] [CrossRef]
  54. Sharipova, A.F.; Psakhie, S.G.; Gotman, I.; Gutmanas, E.Y. Smart nanocomposites based on Fe–Ag and Fe–Cu nanopowders for biodegradable high-strength implants with slow drug release. Phys. Mesomech. 2020, 23, 128–134. [Google Scholar] [CrossRef]
  55. Sharipova, A.F.; Psakhye, S.G.; Gotman, I.; Lerner, M.I.; Lozhkomoev, A.S.; Gutmanas, E.Y. Cold Sintering of Fe–Ag and Fe–Cu Nanocomposites by Consolidation in the High-Pressure Gradient. Russ. J. Non-Ferr. Met. 2019, 60, 162–168. [Google Scholar] [CrossRef]
  56. Romanova, V.M.; Ivanenkov, G.V.; Mingaleev, A.R.; Ter-Oganesyan, A.E.; Shelkovenko, T.A.; Pikuz, S.A. Electric explosion of fine wires: Three groups of materials. Plasma Phys. Rep. 2015, 41, 617–636. [Google Scholar] [CrossRef]
  57. Sahai, A.; Goswami, N.; Kaushik, S.D.; Tripathi, S. Cu/Cu2O/CuO nanoparticles: Novel synthesis by exploding wire technique and extensive characterization. Appl. Surf. Sci. 2016, 390, 974–983. [Google Scholar] [CrossRef]
  58. Park, G.H.; Lee, G.Y.; Kim, H.A.; Lee, A.Y.; Oh, H.R.; Kim, S.Y.; Kim, D.H.; Lee, M.H. Evaluation of alloying effect on the formation of Ni-Fe nanosized powders by pulsed wire discharge. Mater. Sci. Eng. B 2016, 212, 24–29. [Google Scholar] [CrossRef]
  59. Bland, S.N.; Krasik, Y.E.; Yanuka, D.; Gardner, R.; MacDonald, J.; Virozub, A.; Efimov, S.; Gleizer, S.; Chaturvedi, N. Generation of highly symmetric, cylindrically convergent shockwaves in water. Phys. Plasmas 2017, 24, 082702. [Google Scholar] [CrossRef]
  60. Shelkovenko, T.A.; Pikuz, S.A.; Hammer, D.A. A review of projection radiography of plasma and biological objects in X-pinch radiation. Plasma Phys. Rep. 2016, 42, 226–268. [Google Scholar] [CrossRef]
  61. Tokoi, Y.; Orikawa, T.; Suzuki, T.; Nakayama, T.; Suematsu, H.; Niihara, K. Phase Control of Ti–Fe Nanoparticles Prepared by Pulsed Wire Discharge. Jpn. J. Appl. Phys. 2011, 50, 01BJ06. [Google Scholar] [CrossRef]
  62. Ishihara, S.; Koishi, T.; Orikawa, T.; Suematsu, H.; Nakayama, T.; Suzuki, T.; Niihara, K. Synthesis of intermetallic NiAl compound nanoparticles by pulsed wire discharge of twisted Ni and Al wires. Intermetallics 2012, 23, 134–142. [Google Scholar] [CrossRef]
  63. Pervikov, A.; Glazkova, E.; Lerner, M. Energy characteristics of the electrical explosion of two intertwined wires made of dissimilar metals. Phys. Plasmas 2018, 25, 070701. [Google Scholar] [CrossRef]
  64. Lerner, M.I.; Psakhie, S.G.; Lozhkomoev, A.S.; Sharipova, A.F.; Pervikov, A.V.; Gotman, I.; Gutmanas, E.Y. Fe–Cu nanocomposites by high pressure consolidation of powders prepared by electric explosion of wires. Adv. Eng. Mater. 2018, 20, 1701024. [Google Scholar] [CrossRef]
  65. Chau, J.L.H.; Chen, C.Y.; Yang, C.C. Facile synthesis of bimetallic nanoparticles by femtosecond laser irradiation method. Arab. J. Chem. 2017, 10, S1395–S1401. [Google Scholar] [CrossRef]
  66. Suslick, K.S.; Fang, M.; Hyeon, T. Sonochemical synthesis of iron colloids. J. Am. Chem. Soc. 1996, 118, 11960–11961. [Google Scholar] [CrossRef]
  67. Gogate, P.R.; Sutkar, V.S.; Pandit, A.B. Sonochemical reactors: Important design and scale up considerations with a special emphasis on heterogeneous systems. Chem. Eng. J. 2011, 166, 1066–1082. [Google Scholar] [CrossRef]
  68. Shirsath, S.R.; Pinjari, D.V.; Gogate, P.R.; Sonawane, S.H.; Pandit, D.A. Ultrasound assisted synthesis of doped TiO2 nano-particles: Characterization and comparison of effectiveness for photocatalytic oxidation of dyestuff effluent. Ultrason. Sonochem. 2013, 20, 277–286. [Google Scholar] [CrossRef]
  69. Zhao, D.; Zheng, Y.; Li, M.; Baig, S.A.; Wu, D.; Xu, X. Catalytic dechlorination of 2, 4-dichlorophenol by Ni/Fe nanoparticles prepared in the presence of ultrasonic irradiation. Ultrason. Sonochem. 2014, 21, 1714–1721. [Google Scholar] [CrossRef]
  70. Yu, Y.; Huang, Z.; Deng, D.; Ju, Y.; Ren, L.; Xiang, M.; Li, L.; Li, H. Synthesis of millimeter-scale sponge Fe/Cu bimetallic particles removing TBBPA and insights of degradation mechanism. Chem. Eng. J. 2017, 325, 279–288. [Google Scholar] [CrossRef]
  71. Tabrizian, P.; Ma, W.; Bakr, A.; Rahaman, M.S. pH-sensitive and magnetically separable Fe/Cu bimetallic nanoparticles supported by graphene oxide (GO) for high-efficiency removal of tetracyclines. J. Colloid Interface Sci. 2019, 534, 549–562. [Google Scholar] [CrossRef] [PubMed]
  72. Li, F.; Shi, P.; Wu, J.; Qi, X.; Liu, Y.; Li, G. Trace Bimetallic Iron/Manganese Co-Doped N-Ketjenblack Carbon Electrocatalyst for Robust Oxygen Reduction Reaction. J. Electrochem. Soc. 2021, 168, 060502. [Google Scholar] [CrossRef]
  73. Geng, C.; Lin, R.; Yang, P.; Liu, P.; Guo, L.; Fang, Y.; Cui, B. Mild routine to prepare Fe-Mn bimetallic nano-cluster (Fe-Mn NCs) and its magnetic starch-based composite adsorbent (Fe-Mn@ SCAs) for wide pH range adsorption for Hg (II) sewage. J. Taiwan Inst. Chem. Eng. 2023, 144, 104768. [Google Scholar] [CrossRef]
  74. Liang, Y.; Liu, P.; Xiao, J.; Li, H.B.; Wang, C.X.; Yang, G.W. A general strategy for one-step fabrication of one-dimensional magnetic nanoparticle chains based on laser ablation in liquid. Laser Phys. Lett. 2014, 11, 056001. [Google Scholar] [CrossRef]
  75. Liang, Y.; Liu, P.; Xiao, J.; Li, H.; Wang, C.; Yang, G. A microfibre assembly of an iron-carbon composite with giant magnetisation. Sci. Rep. 2013, 3, 3051. [Google Scholar] [CrossRef]
  76. Liang, Y.; Liu, P.; Yang, G. Fabrication of one-dimensional chain of iron-based bimetallic alloying nanoparticles with unique magnetizations. Cryst. Growth Des. 2014, 14, 5847–5855. [Google Scholar] [CrossRef]
  77. Wang, D.; Li, Y. Bimetallic nanocrystals: Liquid-phase synthesis and catalytic applications. Adv. Mater. 2011, 23, 1044–1060. [Google Scholar] [CrossRef]
  78. Gilroy, K.D.; Ruditskiy, A.; Peng, H.C.; Qin, D.; Xia, Y. Bimetallic nanocrystals: Syntheses, properties, and applications. Chem. Rev. 2016, 116, 10414–10472. [Google Scholar] [CrossRef]
  79. Raut, S.S.; Shetty, R.; Raju, N.M.; Kamble, S.P.; Kulkarni, P.S. Screening of zero valent mono/bimetallic catalysts and recommendation of Raney Ni (without reducing agent) for dechlorination of 4-chlorophenol. Chemosphere 2020, 250, 126298. [Google Scholar] [CrossRef]
  80. Ou, J.H.; Sheu, Y.T.; Tsang, D.C.; Sun, Y.J.; Kao, C.M. Application of iron/aluminum bimetallic nanoparticle system for chromium-contaminated groundwater remediation. Chemosphere 2020, 256, 127158. [Google Scholar] [CrossRef]
  81. Wang, Q.; Tian, C.; Shi, B.; Wang, D.; Feng, C. Efficiency and mechanism of micro-and nano-plastic removal with polymeric Al-Fe bimetallic coagulants: Role of Fe addition. J. Hazard. Mater. 2023, 448, 130978. [Google Scholar] [CrossRef] [PubMed]
  82. Sepúlveda, P.; Rubio, M.A.; Baltazar, S.E.; Rojas-Nunez, J.; Llamazares, J.S.; Garcia, A.G.; Arancibia-Miranda, N. As (V) removal capacity of FeCu bimetallic nanoparticles in aqueous solutions: The influence of Cu content and morphologic changes in bimetallic nanoparticles. J. Colloid Interface Sci. 2018, 524, 177–187. [Google Scholar] [CrossRef] [PubMed]
  83. Ulucan-Altuntas, K.; Kuzu, S.L. Modelling and optimization of dye removal by Fe/Cu bimetallic nanoparticles coated with different Cu ratios. Mater. Res. Express 2019, 6, 1150a4. [Google Scholar] [CrossRef]
  84. Abdel-Aziz, A.B.; Mohamed, N.; El-taweel, R.M.; Husien, S.; Fahim, I.S.; Said, L.A.; Radwan, A.G. Crystal violet removal using bimetallic Fe0–Cu and its composites with fava bean activated carbon. Results Eng. 2023, 20, 101420. [Google Scholar] [CrossRef]
  85. Muradova, G.G.; Gadjieva, S.R.; Di Palma, L.; Vilardi, G. Nitrates removal by bimetallic nanoparticles in water. Chem. Eng. Trans. 2016, 47, 205–210. [Google Scholar]
  86. Torres-Blancas, T.; Roa-Morales, G.; Ureña-Núñez, F.; Barrera-Díaz, C.; Dorazco-González, A.; Natividad, R. Ozonation enhancement by Fe–Cu biometallic particles. J. Taiwan Inst. Chem. Eng. 2017, 74, 225–232. [Google Scholar] [CrossRef]
  87. Naser, R.; Shahwan, T. Comparative assessment of the decolorization of aqueous bromophenol blue using Fe nanoparticles and Fe-Ni bimetallic nanoparticles. Desalination Water Treat. 2019, 159, 64416. [Google Scholar] [CrossRef]
  88. Weng, X.; Sun, Q.; Lin, S.; Chen, Z.; Megharaj, M.; Naidu, R. Enhancement of catalytic degradation of amoxicillin in aqueous solution using clay supported bimetallic Fe/Ni nanoparticles. Chemosphere 2014, 103, 80–85. [Google Scholar] [CrossRef]
  89. Zhou, X.; Jing, G.; Lv, B.; Zhou, Z.; Zhu, R. Highly efficient removal of chromium (VI) by Fe/Ni bimetallic nanoparticles in an ultrasound-assisted system. Chemosphere 2016, 160, 332–341. [Google Scholar] [CrossRef]
  90. Zhou, S.; Li, Y.; Chen, J.; Liu, Z.; Wang, Z.; Na, P. Enhanced Cr (VI) removal from aqueous solutions using Ni/Fe bimetallic nanoparticles: Characterization, kinetics and mechanism. RSC Adv. 2014, 4, 50699–50707. [Google Scholar] [CrossRef]
  91. Mansouriieh, N.; Sohrabi, M.R.; Khosravi, M. Adsorption kinetics and thermodynamics of organophosphorus profenofos pesticide onto Fe/Ni bimetallic nanoparticles. Int. J. Environ. Sci. Technol. 2016, 13, 1393–1404. [Google Scholar] [CrossRef]
  92. Liao, H.; Zhu, W.; Duan, T.; Zhang, Y.; He, G.; Wei, Y.; Zhou, J. Beaded segments like bi-metallic nano-zero-valent iron-titanium for the fast and efficient adsorption and reduction of U (VI) in aqueous solutions. Colloids Surf. A Physicochem. Eng. Asp. 2021, 613, 126080. [Google Scholar] [CrossRef]
  93. Koryam, A.A.; El-Wakeel, S.T.; Radwan, E.K.; Darwish, E.S.; Abdel Fattah, A.M. One-step room-temperature synthesis of bimetallic nanoscale zero-valent FeCo by hydrazine reduction: Effect of metal salts and application in contaminated water treatment. ACS Omega 2022, 7, 34810–34823. [Google Scholar] [CrossRef]
  94. Qureshi, S.S.; Memon, S.A.; Ram, N.; Saeed, S.; Mubarak, N.M.; Karri, R.R. Rapid adsorption of selenium removal using iron manganese-based micro adsorbent. Sci. Rep. 2022, 12, 17207. [Google Scholar] [CrossRef]
  95. Guo, D.J.; Ding, Y. Porous nanostructured metals for electrocatalysis. Electroanalysis 2012, 24, 2035–2043. [Google Scholar] [CrossRef]
  96. Shen, W.N.; Dunn, B.; Moore, C.D.; Goorsky, M.S.; Radetic, T.; Gronsky, R. Synthesis of nanoporous bismuth films by liquid-phase deposition. J. Mater. Chem. 2000, 10, 657–662. [Google Scholar] [CrossRef]
  97. Luo, H.; Sun, L.; Lu, Y.; Yan. Electrodeposition of mesoporous semimetal and magnetic metal films from lyotropic liquid crystalline phases. Langmuir 2004, 20, 10218–10222. [Google Scholar] [CrossRef]
  98. Wang, X.; Qi, Z.; Zhao, C.; Wang, W.; Zhang, Z. Influence of alloy composition and dealloying solution on the formation and microstructure of monolithic nanoporous silver through chemical dealloying of Al−Ag alloys. J. Phys. Chem. C 2009, 113, 13139–13150. [Google Scholar] [CrossRef]
  99. Chen, L.; Guo, H.; Fujita, T.; Hirata, A.; Zhang, W.; Inoue, A.; Chen, M. Nanoporous PdNi bimetallic catalyst with enhanced electrocatalytic performances for electro-oxidation and oxygen reduction reactions. Adv. Funct. Mater. 2011, 21, 4364–4370. [Google Scholar] [CrossRef]
  100. Ding, Y.; Chen, M. Nanoporous metals for catalytic and optical applications. MRS Bull. 2009, 34, 569–576. [Google Scholar] [CrossRef]
  101. Han, B.; Xu, C. Nanoporous PdFe alloy as highly active and durable electrocatalyst for oxygen reduction reaction. Int. J. Hydrogen Energy 2014, 39, 18247–18255. [Google Scholar] [CrossRef]
  102. Tian, C.; Chen, D.; Lu, N.; Li, Y.; Cui, R.; Han, Z.; Zhang, G. Electrochemical bisphenol A sensor based on nanoporous PtFe alloy and graphene modified glassy carbon electrode. J. Electroanal. Chem. 2018, 830, 27–33. [Google Scholar] [CrossRef]
  103. Gu, J.; Zhang, Y.W.; Tao, F.F. Shape control of bimetallic nanocatalysts through well-designed colloidal chemistry approaches. Chem. Soc. Rev. 2012, 41, 8050–8065. [Google Scholar] [CrossRef]
  104. Chen, S.; Jenkins, S.V.; Tao, J.; Zhu, Y.; Chen, J. Anisotropic seeded growth of Cu–M (M= Au, Pt, or Pd) bimetallic nanorods with tunable optical and catalytic properties. J. Phys. Chem. C 2013, 117, 8924–8932. [Google Scholar] [CrossRef]
  105. Wang, Z.; Huang, W.; Fennell, D.E. Rapid dechlorination of 1, 2, 3, 4-TCDD by Ag/Fe bimetallic particles. Chem. Eng. J. 2015, 273, 465–471. [Google Scholar] [CrossRef]
  106. Huang, C.C.; Lo, S.L.; Lien, H.L. Vitamin B12-mediated hydrodechlorination of dichloromethane by bimetallic Cu/Al particles. Chem. Eng. J. 2015, 273, 413–420. [Google Scholar] [CrossRef]
  107. Zhou, X.J.; Harmer, A.J.; Heinig, N.F.; Leung, K.T. Parametric study on electrochemical deposition of copper nanoparticles on an ultrathin polypyrrole film deposited on a gold film electrode. Langmuir 2004, 20, 5109–5113. [Google Scholar] [CrossRef]
  108. Yancey, D.F.; Carino, E.V.; Crooks, R.M. Electrochemical synthesis and electrocatalytic properties of Au@ Pt dendrimer-encapsulated nanoparticles. J. Am. Chem. Soc. 2010, 132, 10988–10989. [Google Scholar] [CrossRef]
  109. Tremont, R.J.; Cruz, G.; Cabrera, C.R. Pt electrodeposition on a copper surface modified with 3-mercaptopropyltrimethoxysilane and 1-propanethiol. J. Electroanal. Chem. 2003, 558, 65–74. [Google Scholar] [CrossRef]
  110. Stiger, R.M.; Gorer, S.; Craft, B.; Penner, R.M. Investigations of electrochemical silver nanocrystal growth on hydrogen-terminated silicon (100). Langmuir 1999, 15, 790–798. [Google Scholar] [CrossRef]
  111. Ortega, J.M. Electrodeposition of copper on poly (o-aminophenol) modified platinum electrode. Thin Solid Film. 2000, 360, 159–165. [Google Scholar] [CrossRef]
  112. Claussen, J.C.; Franklin, A.D.; ul Haque, A.; Porterfield, D.M.; Fisher, T.S. Electrochemical biosensor of nanocube-augmented carbon nanotube networks. ACS Nano 2009, 3, 37–44. [Google Scholar] [CrossRef] [PubMed]
  113. Yang, J.; Zhang, W.D.; Gunasekaran, S. An amperometric non-enzymatic glucose sensor by electrodepositing copper nanocubes onto vertically well-aligned multi-walled carbon nanotube arrays. Biosens. Bioelectron. 2010, 26, 279–284. [Google Scholar] [CrossRef] [PubMed]
  114. Xu, Y.; Chen, L.; Wang, X.; Yao, W.; Zhang, Q. Recent advances in noble metal based composite nanocatalysts: Colloidal synthesis, properties, and catalytic applications. Nanoscale 2015, 7, 10559–10583. [Google Scholar] [CrossRef]
  115. Hyeon, T. Chemical synthesis of magnetic nanoparticles. Chem. Commun. 2003, 9, 927–934. [Google Scholar] [CrossRef]
  116. Kang, K.M.; Kim, H.W.; Shim, I.W.; Kwak, H.Y. Catalytic test of supported Ni catalysts with core/shell structure for dry reforming of methane. Fuel Process. Technol. 2011, 92, 1236–1243. [Google Scholar] [CrossRef]
  117. Duan, S.; Wang, R. Bimetallic nanostructures with magnetic and noble metals and their physicochemical applications. Prog. Nat. Sci. Mater. Int. 2013, 23, 113–126. [Google Scholar] [CrossRef]
  118. Riva, J.S.; Pozo-López, G.; Condo, A.M.; Fabietti, L.M.; Urreta, S.E. Low temperature ferromagnetism in Rh-rich Fe-Rh granular nanowires. J. Alloys Compd. 2018, 747, 1008–1017. [Google Scholar] [CrossRef]
  119. Riva, J.S.; Pozo-López, G.; Condo, A.M.; Viqueira, M.S.; Urreta, S.E.; Cornejo, D.R.; Fabietti, L.M. Biphasic FeRh nanowires synthesized by AC electrodeposition. J. Alloys Compd. 2016, 688, 804–813. [Google Scholar] [CrossRef]
  120. Riva, J.S.; Pozo-López, G.; Condó, A.M.; Levingston, J.M.; Fabietti, L.M.; Urreta, S.E. Magnetic viscosity in iron-rhodium nanowires. J. Alloys Compd. 2017, 709, 531–534. [Google Scholar] [CrossRef]
  121. Hao, Y.; Wang, J.; Xia, Q.; Zhang, X.; Song, Y.; Yao, Z. Accelerating Fe (iii)/Fe (ii) redox cycling by Zn 0 in micro-nano dendritic Fe–Zn alloy for enhanced Fenton-like degradation of phenol. New J. Chem. 2023, 47, 17508–17516. [Google Scholar] [CrossRef]
  122. Zhang, H.; Jin, M.; Wang, J.; Li, W.; Camargo, P.H.; Kim, M.J.; Yang, D.; Xie, Z.; Xia, Y. Synthesis of Pd−Pt bimetallic nanocrystals with a concave structure through a bromide-induced galvanic replacement reaction. J. Am. Chem. Soc. 2011, 33, 6078–6089. [Google Scholar] [CrossRef] [PubMed]
  123. Yuan, Y.; Li, H.; Lai, B.; Yang, P.; Gou, M.; Zhou, Y.; Sun, G. Removal of high-concentration CI acid orange 7 from aqueous solution by zerovalent iron/copper (Fe/Cu) bimetallic particles. Ind. Eng. Chem. Res. 2014, 53, 2605–2613. [Google Scholar] [CrossRef]
  124. Mahmoud, M.S.; Mahmoud, A.S. Wastewater treatment using nano bimetallic iron/copper, adsorption isotherm, kinetic studies, and artificial intelligence neural networks. Emergent Mater. 2021, 4, 1455–1463. [Google Scholar] [CrossRef]
  125. Mahmoud, A.S.; Mohamed, N.Y.; Mostafa, M.K.; Mahmoud, M.S. Effective chromium adsorption from aqueous solutions and tannery wastewater using bimetallic Fe/Cu nanoparticles: Response surface methodology and artificial neural network. Air Soil Water Res. 2021, 14, 11786221211028162. [Google Scholar] [CrossRef]
  126. Lai, B.; Zhang, Y.; Chen, Z.; Yang, P.; Zhou, Y.; Wang, J. Removal of p-nitrophenol (PNP) in aqueous solution by the micron-scale iron–copper (Fe/Cu) bimetallic particles. Appl. Catal. B Environ. 2014, 144, 816–830. [Google Scholar] [CrossRef]
  127. Lai, B.; Zhang, Y.H.; Yuan, Y.; Chen, Z.Y.; Yang, P. Influence of preparation conditions on characteristics, reactivity, and operational life of microsized Fe/Cu bimetallic particles. Ind. Eng. Chem. Res. 2014, 53, 12295–12304. [Google Scholar] [CrossRef]
  128. Xiong, Z.; Lai, B.; Yang, P.; Zhou, Y.; Wang, J.; Fang, S. Comparative study on the reactivity of Fe/Cu bimetallic particles and zero valent iron (ZVI) under different conditions of N2, air or without aeration. J. Hazard. Mater. 2015, 297, 261–268. [Google Scholar] [CrossRef]
  129. Ren, Y.; Lai, B. Comparative study on the characteristics, operational life and reactivity of Fe/Cu bimetallic particles prepared by electroless and displacement plating process. RSC Adv. 2016, 6, 58302–58314. [Google Scholar] [CrossRef]
  130. Liu, X.; Fan, J.H.; Ma, L.M. Simultaneously degradation of 2, 4-Dichlorophenol and EDTA in aqueous solution by the bimetallic Cu–Fe/O2 system. Environ. Sci. Pollut. Res. 2015, 22, 1186–1198. [Google Scholar] [CrossRef]
  131. Fu, F.; Cheng, Z.; Dionysiou, D.D.; Tang, B. Fe/Al bimetallic particles for the fast and highly efficient removal of Cr (VI) over a wide pH range: Performance and mechanism. J. Hazard. Mater. 2015, 298, 261–269. [Google Scholar] [CrossRef] [PubMed]
  132. Cheng, Z.; Fu, F.; Dionysiou, D.D.; Tang, B. Adsorption, oxidation, and reduction behavior of arsenic in the removal of aqueous As (III) by mesoporous Fe/Al bimetallic particles. Water Res. 2016, 96, 22–31. [Google Scholar] [CrossRef]
  133. Xiang, S.; Cheng, W.; Nie, X.; Ding, C.; Yi, F.; Asiri, A.M.; Marwani, H.M. Zero-valent iron-aluminum for the fast and effective U (VI) removal. J. Taiwan Inst. Chem. Eng. 2018, 85, 186–192. [Google Scholar] [CrossRef]
  134. Yang, Z.; Zhang, X.; Pu, S.; Ni, R.; Lin, Y.; Liu, Y. Novel Fenton-like system (Mg/Fe-O2) for degradation of 4-chlorophenol. Environ. Pollut. 2019, 250, 906–913. [Google Scholar] [CrossRef]
  135. Aghaei, E.; Wang, Z.; Tadesse, B.; Tabelin, C.B.; Quadir, Z.; Alorro, R.D. Performance evaluation of Fe-Al bimetallic particles for the removal of potentially toxic elements from combined acid mine drainage-effluents from refractory gold ore processing. Minerals 2021, 11, 590. [Google Scholar] [CrossRef]
  136. Aghaei, E.; Tadesse, B.; Tabelin, C.B.; Alorro, R.D. Mercury sequestration from synthetic and real gold processing wastewaters using Fe–Al bimetallic particles. J. Clean. Prod. 2022, 372, 133482. [Google Scholar] [CrossRef]
  137. He, Y.; Sun, H.; Liu, W.; Yang, W.; Lin, A. Study on removal effect of Cr (VI) and surface reaction mechanisms by bimetallic system in aqueous solution. Environ. Technol. 2020, 41, 1867–1876. [Google Scholar] [CrossRef]
  138. Yeh, L.; Yen, C.H.; Kao, Y.L.; Lien, H.L.; Chang, S.M. Inactivation of Escherichia coli by dual-functional zerovalent Fe/Al composites in water. Chemosphere 2022, 299, 134371. [Google Scholar] [CrossRef]
  139. Liu, X.; Fan, J.H.; Hao, Y.; Ma, L.M. The degradation of EDTA by the bimetallic Fe–Cu/O2 system. Chem. Eng. J. 2014, 250, 354–365. [Google Scholar] [CrossRef]
  140. Park, I.; Ito, M.; Jeon, S.; Tabelin, C.B.; Phengsaart, T.; Silwamba, M.; Hiroyoshi, N. A novel recycling route for aluminum alloys: Synthesis of Fe/Al bimetallic materials and magnetic separation. Miner. Eng. 2023, 201, 108202. [Google Scholar] [CrossRef]
  141. Lien, H.L.; Yu, C.H.; Kamali, S.; Sahu, R.S. Bimetallic Fe/Al system: An all-in-one solid-phase Fenton reagent for generation of hydroxyl radicals under oxic conditions. Sci. Total Environ. 2019, 673, 480–488. [Google Scholar] [CrossRef]
  142. Tabelin, C.B.; Resabal, V.J.T.; Park, I.; Villanueva, M.G.B.; Choi, S.; Ebio, R.; Cabural, P.J.; Villacorte-Tabelin, M.; Orbecido, A.; Alorro, R.D.; et al. Repurposing of aluminum scrap into magnetic Al0/ZVI bimetallic materials: Two-stage mechanical-chemical synthesis and characterization of products. J. Clean. Prod. 2021, 317, 128285. [Google Scholar] [CrossRef]
  143. Sun, Y.; Xia, Y. Mechanistic study on the replacement reaction between silver nanostructures and chloroauric acid in aqueous medium. J. Am. Chem. Soc. 2004, 126, 3892–3901. [Google Scholar] [CrossRef]
  144. Liu, X.; Fan, J.H.; Ma, L.M. Elimination of 4-chlorophenol in aqueous solution by the bimetallic Al–Fe/O2 at normal temperature and pressure. Chem. Eng. J. 2014, 236, 274–284. [Google Scholar] [CrossRef]
  145. Fan, J.H.; Liu, X.; Ma, L.M. EDTA enhanced degradation of 4-bromophenol by Al0–Fe0–O2 system. Chem. Eng. J. 2015, 263, 71–82. [Google Scholar] [CrossRef]
  146. Fan, J.; Wang, H.; Ma, L. Oxalate-assisted oxidative degradation of 4-chlorophenol in a bimetallic, zero-valent iron–aluminum/air/water system. Environ. Sci. Pollut. Res. 2016, 23, 16686–16698. [Google Scholar] [CrossRef]
  147. Khoshandam, B.; Kumar, R.V.; Jamshidi, E. Simulation of non-catalytic gas–solid reactions: Application of grain model for the reduction of cobalt oxide with methane. Miner. Process. Extr. Metall. 2005, 114, 10–22. [Google Scholar] [CrossRef]
  148. Khoshandam, B.; Kumar, R.V.; Jamshidi, E. Kinetics of reduction of manganese oxide by methane. Can. Metall. Q. 2007, 46, 365–371. [Google Scholar] [CrossRef]
  149. Yape, E.O.; Anacleto, N.M. Direct smelting of chromite and laterite ores with carbon under argon atmosphere. Adv. Mater. Res. 2014, 849, 170–176. [Google Scholar] [CrossRef]
  150. Balangao, J.K.B.; Podiotan, F.J.C.; Ambolode, A.E.C.; Anacleto, N.M.; Namoco, C.S., Jr. Production of Iron-Chromium-Nickel Metal Alloys via Reduction of Mixed Chromite Ore from Zambales and Laterite Ore from Taganito, Surigao del Norte under Argon Atmosphere. Indian J. Sci. Technol. 2018, 11, 1–11. [Google Scholar]
  151. Balangao, J.K.B.; Podiotan, F.J.C.; Ambolode, A.E.C.; Anacleto, N.M. Isothermal Reduction Smelting of Mixed Chromite-Laterite Samples with Coconut Charcoal as Reductant under Argon Atmosphere in a Vertical Electric Arc Furnace. Int. J. Mech. Eng. Technol. 2022, 13, 46–53. [Google Scholar]
  152. Meshkini Far, R.; Ischenko, O.V.; Dyachenko, A.G.; Bieda, O.; Gaidai, S.V.; Lisnyak, V.V. CO2 hydrogenation into CH4 over Ni–Fe catalysts. Funct. Mater. Lett. 2018, 11, 1850057. [Google Scholar] [CrossRef]
  153. Tang, H.; Li, S.; Gong, D.; Guan, Y.; Liu, Y. Bimetallic Ni-Fe catalysts derived from layered double hydroxides for CO methanation from syngas. Front. Chem. Sci. Eng. 2017, 11, 613–623. [Google Scholar] [CrossRef]
  154. Shao, Y.; Wang, J.; Sun, K.; Gao, G.; Li, C.; Zhang, L.; Zhang, S.; Xu, L.; Hu, G.; Hu, X. Selective hydrogenation of furfural and its derivative over bimetallic NiFe-based catalysts: Understanding the synergy between Ni sites and Ni–Fe alloy. Renew. Energy 2021, 170, 1114–1128. [Google Scholar] [CrossRef]
  155. Li, D.; Koike, M.; Wang, L.; Nakagawa, Y.; Xu, Y.; Tomishige, K. Regenerability of hydrotalcite-derived nickel–iron alloy nanoparticles for syngas production from biomass tar. ChemSusChem 2014, 7, 510–522. [Google Scholar] [CrossRef]
  156. Zhao, J.; Shi, R.; Waterhouse, G.I.; Zhang, T. Selective photothermal CO2 reduction to CO, CH4, alkanes, alkenes over bimetallic alloy catalysts derived from layered double hydroxide nanosheets. Nano Energy 2022, 102, 107650. [Google Scholar] [CrossRef]
  157. Liu, Q.; Wang, J.; An, K.; Zhang, S.; Liu, G.; Liu, Y. Highly dispersed Ni–Fe alloy catalysts on MgAl2O4 derived from hydrotalcite for direct ethanol synthesis from syngas. Energy Technol. 2020, 8, 2000205. [Google Scholar] [CrossRef]
  158. De Masi, D.; Asensio, J.M.; Fazzini, P.F.; Lacroix, L.M.; Chaudret, B. Engineering iron–nickel nanoparticles for magnetically induced CO2 methanation in continuous flow. Angew. Chem. Int. Ed. 2020, 59, 6187–6191. [Google Scholar] [CrossRef]
  159. Ferrando, R.; Jellinek, J.; Johnston, R.L. Nanoalloys: From theory to applications of alloy clusters and nanoparticles. Chem. Rev. 2008, 108, 845–910. [Google Scholar] [CrossRef]
  160. Zhang, L.; Lepper, M.; Stark, M.; Menzel, T.; Lungerich, D.; Jux, N.; Hieringer, W.; Steinrück, H.P.; Marbach, H. On the critical role of the substrate: The adsorption behaviour of tetrabenzoporphyrins on different metal surfaces. Phys. Chem. Chem. Phys. 2017, 19, 20281–20289. [Google Scholar] [CrossRef]
  161. Wang, J.; Wu, A.; Qiu, Z.; Li, A.; Qin, W.; Huang, H. N-CNT supported Fe/Ce bimetallic catalyst for Al-air aqueous batteries. Appl. Surf. Sci. 2023, 608, 155185. [Google Scholar] [CrossRef]
  162. Tian, F.; Zhong, S.; Nie, W.; Zeng, M.; Chen, B.; Liu, X. Multi-walled carbon nanotubes prepared with low-cost Fe-Al bimetallic catalysts for high-rate rechargeable Li-ion batteries. J. Solid State Electrochem. 2020, 24, 667–674. [Google Scholar] [CrossRef]
  163. Wang, J.; Liu, C.; Li, J.; Luo, R.; Hu, X.; Sun, X.; Shen, J.; Han, W.; Wang, L. In-situ incorporation of iron-copper bimetallic particles in electrospun carbon nanofibers as an efficient Fenton catalyst. Appl. Catal. B Environ. 2017, 207, 316–325. [Google Scholar] [CrossRef]
  164. Nam, G.; Park, J.; Choi, M.; Oh, P.; Park, S.; Kim, M.G.; Park, N.; Cho, J.; Lee, J.S. Carbon-coated core–shell Fe–Cu nanoparticles as highly active and durable electrocatalysts for a Zn–air battery. ACS Nano 2015, 9, 6493–6501. [Google Scholar] [CrossRef]
  165. Wang, Y.; Zhao, H.; Zhao, G. Iron-copper bimetallic nanoparticles embedded within ordered mesoporous carbon as effective and stable heterogeneous Fenton catalyst for the degradation of organic contaminants. Appl. Catal. B Environ. 2015, 164, 396–406. [Google Scholar] [CrossRef]
  166. Wu, H.; Feng, Q.; Lu, P.; Chen, M.; Yang, H. Degradation mechanisms of cefotaxime using biochar supported Co/Fe bimetallic nanoparticles. Environ. Sci. Water Res. Technol. 2018, 4, 964–975. [Google Scholar] [CrossRef]
  167. Wu, H.; Feng, Q. Fabrication of bimetallic Ag/Fe immobilized on modified biochar for removal of carbon tetrachloride. J. Environ. Sci. 2017, 54, 346–357. [Google Scholar] [CrossRef]
  168. Xing, X.; Ren, X.; Alharbi, N.S.; Chen, C. Biochar-supported Fe/Ni bimetallic nanoparticles for the efficient removal of Cr (VI) from aqueous solution. J. Mol. Liq. 2022, 359, 119257. [Google Scholar] [CrossRef]
  169. Hao, J.; Wu, L.; Lu, X.; Zeng, Y.; Jia, B.; Luo, T.; He, S.; Liang, L. A stable Fe/Co bimetallic modified biochar for ofloxacin removal from water: Adsorption behavior and mechanisms. RSC Adv. 2022, 12, 31650–31662. [Google Scholar] [CrossRef]
  170. Ji, J.; Xu, S.; Ma, Z.; Mou, Y. Trivalent antimony removal using carbonaceous nanomaterial loaded with zero-valent bimetal (iron/copper) and their effect on seed growth. Chemosphere 2022, 296, 134047. [Google Scholar] [CrossRef]
  171. Xia, H.; Zhang, Y.; Chen, Q.; Liu, R.; Wang, H. Unraveling adsorption characteristics and removal mechanism of novel Zn/Fe-bimetal-loaded and starch-coated corn cobs biochar for Pb (II) and Cd (II) in wastewater. J. Mol. Liq. 2023, 391, 123375. [Google Scholar] [CrossRef]
  172. Bose, S.; Mukherjee, T.; Rahaman, M. Simultaneous adsorption of manganese and fluoride from aqueous solution via bimetal impregnated activated carbon derived from waste tire: Response surface method modeling approach. Environ. Prog. Sustain. Energy 2021, 40, e13600. [Google Scholar] [CrossRef]
  173. Danmaliki, G.I.; Saleh, T.A. Effects of bimetallic Ce/Fe nanoparticles on the desulfurization of thiophenes using activated carbon. Chem. Eng. J. 2017, 307, 914–927. [Google Scholar] [CrossRef]
  174. Tian, H.; Chen, C.; Zhu, T.; Zhu, B.; Sun, Y. Characterization and degradation mechanism of bimetallic iron-based/AC activated persulfate for PAHs-contaminated soil remediation. Chemosphere 2021, 267, 128875. [Google Scholar] [CrossRef]
  175. Kakavandi, B.; Kalantary, R.R.; Farzadkia, M.; Mahvi, A.H.; Esrafili, A.; Azari, A.; Yari, A.R.; Javid, A.B. Enhanced chromium (VI) removal using activated carbon modified by zero valent iron and silver bimetallic nanoparticles. J. Environ. Health Sci. Eng. 2014, 12, 115. [Google Scholar] [CrossRef]
  176. Fong, W.M.; Affam, A.C.; Chung, W.C. Synthesis of Ag/Fe/CAC for colour and COD removal from methylene blue dye wastewater. Int. J. Environ. Sci. Technol. 2020, 17, 3485–3494. [Google Scholar] [CrossRef]
  177. Rahaman, M.; Das, A.; Bose, S. Development of copper—Iron bimetallic nanoparticle impregnated activated carbon derived from coconut husk and its efficacy as a novel adsorbent toward the removal of chromium (VI) from aqueous solution. Water Environ. Res. 2021, 93, 1417–1427. [Google Scholar] [CrossRef]
  178. Zhao, R.; Du, X.; Cao, K.; Gong, M.; Li, Y.; Ai, J.; Ye, R.; Chen, R.; Shan, B. Highly dispersed Fe-decorated Ni nanoparticles prepared by atomic layer deposition for dry reforming of methane. Int. J. Hydrogen Energy 2023, 48, 28780–28791. [Google Scholar] [CrossRef]
  179. Mutz, B.; Belimov, M.; Wang, W.; Sprenger, P.; Serrer, M.A.; Wang, D.; Pfeifer, P.; Kleist, W.; Grunwaldt, J.D. Potential of an alumina-supported Ni3Fe catalyst in the methanation of CO2: Impact of alloy formation on activity and stability. ACS Catal. 2017, 7, 6802–6814. [Google Scholar] [CrossRef]
  180. Sun, Y.; Yang, Z.; Tian, P.; Sheng, Y.; Xu, J.; Han, Y.F. Oxidative degradation of nitrobenzene by a Fenton-like reaction with Fe-Cu bimetallic catalysts. Appl. Catal. B Environ. 2019, 244, 1–10. [Google Scholar] [CrossRef]
  181. Liuzzi, D.; Pérez-Alonso, F.J.; Rojas, S. Ru-M (M= Fe or Co) Catalysts with high Ru surface concentration for Fischer-Tropsch synthesis. Fuel 2021, 293, 120435. [Google Scholar] [CrossRef]
  182. Zhang, L.; Meng, Z.; Zang, S. Preparation and characterization of Pd/Fe bimetallic nanoparticles immobilized on Al2O3/PVDF membrane: Parameter optimization and dechlorination of dichloroacetic acid. J. Environ. Sci. 2015, 31, 194–202. [Google Scholar] [CrossRef] [PubMed]
  183. Pradhan, A.C.; Sahoo, M.K.; Bellamkonda, S.; Parida, K.M.; Rao, G.R. Enhanced photodegradation of dyes and mixed dyes by heterogeneous mesoporous Co–Fe/Al2O3–MCM-41 nanocomposites: Nanoparticles formation, semiconductor behavior and mesoporosity. RSC Adv. 2016, 6, 94263–94277. [Google Scholar] [CrossRef]
  184. Munoz, M.; de Pedro, Z.M.; Casas, J.A.; Rodriguez, J.J. Improved γ-alumina-supported Pd and Rh catalysts for hydrodechlorination of chlorophenols. Appl. Catal. A Gen. 2014, 488, 78–85. [Google Scholar] [CrossRef]
  185. Wang, J.; Liu, C.; Tong, L.; Li, J.; Luo, R.; Qi, J.; Li, Y.; Wang, L. Iron–copper bimetallic nanoparticles supported on hollow mesoporous silica spheres: An effective heterogeneous Fenton catalyst for orange II degradation. RSC Adv. 2015, 5, 69593–69605. [Google Scholar] [CrossRef]
  186. Wang, J.; Liu, C.; Hussain, I.; Li, C.; Li, J.; Sun, X.; Shen, J.; Han, W.; Wang, L. Iron–copper bimetallic nanoparticles supported on hollow mesoporous silica spheres: The effect of Fe/Cu ratio on heterogeneous Fenton degradation of a dye. RSC Adv. 2016, 6, 54623–54635. [Google Scholar] [CrossRef]
  187. Lin, H.; Zhong, X.; Ciotonea, C.; Fan, X.; Mao, X.; Li, Y.; Deng, B.; Zhang, H.; Royer, S. Efficient degradation of clofibric acid by electro-enhanced peroxydisulfate activation with Fe-Cu/SBA-15 catalyst. Appl. Catal. B Environ. 2018, 230, 1–10. [Google Scholar] [CrossRef]
  188. Yan, H.; Qu, H.; Bai, H.; Zhong, Q. Property, active species and reaction mechanism of NO and NH3 over mesoporous Fe-Al-SBA-15 via microwave assisted synthesis for NH3-SCR. J. Mol. Catal. A Chem. 2015, 403, 1–9. [Google Scholar] [CrossRef]
  189. Kurniawan, A.; Sutiono, H.; Ju, Y.H.; Soetaredjo, F.E.; Ayucitra, A.; Yudha, A.; Ismadji, S. Utilization of rarasaponin natural surfactant for organo-bentonite preparation: Application for methylene blue removal from aqueous effluent. Microporous Mesoporous Mater. 2011, 142, 184–193. [Google Scholar] [CrossRef]
  190. Campos, B.; Aguilar-Carrillo, J.; Algarra, M.; Gonçalves, M.A.; Rodríguez-Castellón, E.; da Silva, J.C.E.; Bobos, I. Adsorption of uranyl ions on kaolinite, montmorillonite, humic acid and composite clay material. Appl. Clay Sci. 2013, 85, 53–63. [Google Scholar] [CrossRef]
  191. Zhou, C.H.; Zhang, D.; Tong, D.S.; Wu, L.M.; Yu, W.H.; Ismadji, S. Paper-like composites of cellulose acetate–organo-montmorillonite for removal of hazardous anionic dye in water. Chem. Eng. J. 2012, 209, 223–234. [Google Scholar] [CrossRef]
  192. Murray, H.H. Applied Clay Mineralogy: Occurrences, Processing and Applications of Kaolins, Bentonites, Palygorskite–Sepiolite, and Common Clays; Elsevier: Amsterdam, The Netherlands, 2006. [Google Scholar]
  193. Sabouri, M.R.; Sohrabi, M.R.; Zeraatkar Moghaddam, A. Response surface method Optimization of the Dyes Degradation using Zero-Valent Iron based Bimetallic Nanoparticle on the Bentonite Clay Surface. Pollution 2020, 6, 581–595. [Google Scholar]
  194. Weng, X.; Cai, W.; Lan, R.; Sun, Q.; Chen, Z. Simultaneous removal of amoxicillin, ampicillin and penicillin by clay supported Fe/Ni bimetallic nanoparticles. Environ. Pollut. 2018, 236, 562–569. [Google Scholar] [CrossRef] [PubMed]
  195. Jiang, M.Q.; Jin, X.Y.; Lu, X.Q.; Chen, Z.L. Adsorption of Pb (II), Cd (II), Ni (II) and Cu (II) onto natural kaolinite clay. Desalination 2010, 252, 33–39. [Google Scholar] [CrossRef]
  196. Zhang, X.; Lin, S.; Chen, Z.; Megharaj, M.; Naidu, R. Kaolinite-supported nanoscale zero-valent iron for removal of Pb2+ from aqueous solution: Reactivity, characterization and mechanism. Water Res. 2011, 45, 3481–3488. [Google Scholar] [CrossRef]
  197. Jin, X.; Chen, Z.; Wang, T.; Chen, Z.; Megharaj, M.; Naidu, R. Simultaneous removal of co-contaminants: Acid brilliant violet and Cu 2+ by functional bimetallic Fe/Pd nanoparticles. J. Nanopart. Res. 2014, 16, 2657. [Google Scholar] [CrossRef]
  198. Masters, A.F.; Maschmeyer, T. Zeolites—From curiosity to cornerstone. Microporous Mesoporous Mater. 2011, 142, 423–438. [Google Scholar] [CrossRef]
  199. Rahman, R.O.A.; El-Kamash, A.M.; Hung, Y.T. Applications of nano-zeolite in wastewater treatment: An overview. Water 2022, 14, 137. [Google Scholar] [CrossRef]
  200. Xu, L.; Zhao, L.; Mao, Y.; Zhou, Z.; Wu, D. Enhancing the degradation of bisphenol A by dioxygen activation using bimetallic Cu/Fe@ zeolite: Critical role of Cu (I) and superoxide radical. Sep. Purif. Technol. 2020, 253, 117550. [Google Scholar] [CrossRef]
  201. Ezzatahmadi, N.; Bao, T.; Liu, H.; Millar, G.J.; Ayoko, G.A.; Zhu, J.; Zhu, R.; Liang, X.; He, H.; Xi, Y. Catalytic degradation of Orange II in aqueous solution using diatomite-supported bimetallic Fe/Ni nanoparticles. RSC Adv. 2018, 8, 7687–7696. [Google Scholar] [CrossRef]
  202. Ezzatahmadi, N.; Millar, G.J.; Ayoko, G.A.; Zhu, J.; Zhu, R.; Liang, X.; He, H.; Xi, Y. Degradation of 2,4-dichlorophenol using palygorskite-supported bimetallic Fe/Ni nanocomposite as a heterogeneous catalyst. Appl. Clay Sci. 2019, 168, 276–286. [Google Scholar] [CrossRef]
  203. Ezzatahmadi, N.; Marshall, D.L.; Hou, K.; Ayoko, G.A.; Millar, G.J.; Xi, Y. Simultaneous adsorption and degradation of 2, 4-dichlorophenol on sepiolite-supported bimetallic Fe/Ni nanoparticles. J. Environ. Chem. Eng. 2019, 7, 102955. [Google Scholar] [CrossRef]
  204. Svarovskaya, N.; Bakina, O.; Glazkova, E.; Rodkevich, N.; Lerner, M.; Vornakova, E.; Chzhou, V.; Naumova, L. Synthesis of novel hierarchical micro/nanostructures AlOOH/AlFe and their application for As (V) removal. Environ. Sci. Pollut. Res. 2022, 29, 1246–1258. [Google Scholar] [CrossRef]
  205. Hina, R.; Arafa, I.; Al-Khateeb, F. Gas phase hydrodechlorination of CCl4 over Pd–Cu and Pd–Fe bimetallic catalysts supported on an AlF3 matrix. Prog. React. Kinet. Mech. 2016, 41, 29–38. [Google Scholar] [CrossRef]
  206. Xi, J.; Wang, Q.; Duan, X.; Zhang, N.; Yu, J.; Sun, H.; Wang, S. Continuous flow reduction of organic dyes over Pd-Fe alloy based fibrous catalyst in a fixed-bed system. Chem. Eng. Sci. 2021, 231, 116303. [Google Scholar] [CrossRef]
  207. Aftab, T.B.; Hussain, A.; Li, D. Application of a novel bimetallic hydrogel based on iron and cobalt for the synergistic catalytic degradation of Congo Red dye. J. Chin. Chem. Soc. 2019, 66, 919–927. [Google Scholar] [CrossRef]
  208. Shi, D.; Ouyang, Z.; Zhao, Y.; Xiong, J.; Shi, X. Catalytic reduction of hexavalent chromium using iron/palladium bimetallic nanoparticle-assembled filter paper. Nanomaterials 2019, 9, 1183. [Google Scholar] [CrossRef]
  209. Ge, W.; Li, S.; Jiang, M.; He, G.; Zhang, W. Cu/Fe bimetallic modified fly ash: An economical adsorbent for enhanced phosphorus removal from aqueous solutions. Water Air Soil Pollut. 2022, 233, 182. [Google Scholar] [CrossRef]
  210. Balangao, J.K.B.; Ramos, M.S.K. Production of Autoclaved Aerated Concretes with Fly Ash from a Coal Power Plant in Misamis Oriental, Philippines. Indian J. Sci. Technol. 2019, 12, 1–16. [Google Scholar] [CrossRef]
  211. Balangao, J.K.B.; Caingles, V.K.S.; Baguhin, I.A. Morphological and environmental characterization of lime sludge/fly ash stabilized sub-base materials. Sci. Int. 2023, 35, 283–290. [Google Scholar]
  212. Deplanche, K.; Caldelari, I.; Mikheenko, I.P.; Sargent, F.; Macaskie, L.E. Involvement of hydrogenases in the formation of highly catalytic Pd (0) nanoparticles by bioreduction of Pd (II) using Escherichia coli mutant strains. Microbiology 2010, 156, 2630–2640. [Google Scholar] [CrossRef] [PubMed]
  213. Pantidos, N.; Horsfall, L.E. Biological synthesis of metallic nanoparticles by bacteria, fungi and plants. J. Nanomed. Nanotechnol. 2014, 5, 1000233. [Google Scholar] [CrossRef]
  214. Yong, P.; Rowson, N.A.; Farr, J.P.G.; Harris, I.R.; Macaskie, L.E. Bioaccumulation of palladium by Desulfovibrio desulfuricans. J. Chem. Technol. Biotechnol. 2002, 77, 593–601. [Google Scholar] [CrossRef]
  215. Alruqi, S.S.; Al-Thabaiti, S.A.; Khan, Z. Iron-nickel bimetallic nanoparticles: Surfactant assisted synthesis and their catalytic activities. J. Mol. Liq. 2019, 282, 448–455. [Google Scholar] [CrossRef]
  216. Lin, Y.; Jin, X.; Khan, N.I.; Owens, G.; Chen, Z. Efficient removal of As (III) by calcined green synthesized bimetallic Fe/Pd nanoparticles based on adsorption and oxidation. J. Clean. Prod. 2021, 286, 124987. [Google Scholar] [CrossRef]
  217. Zhang, J.; Niu, Y.; Zhou, Y.; Ju, S.; Gu, Y. Green preparation of nano-zero-valent iron-copper bimetals for nitrate removal: Characterization, reduction reaction pathway, and mechanisms. Adv. Powder Technol. 2022, 33, 103807. [Google Scholar] [CrossRef]
  218. Zhu, F.; Ma, S.; Liu, T.; Deng, X. Green synthesis of nano zero-valent iron/Cu by green tea to remove hexavalent chromium from groundwater. J. Clean. Prod. 2018, 174, 184–190. [Google Scholar] [CrossRef]
  219. Lin, Y.; Jin, X.; Khan, N.I.; Owens, G.; Chen, Z. Bimetallic Fe/Ni nanoparticles derived from green synthesis for the removal of arsenic (V) in mine wastewater. J. Environ. Manag. 2022, 301, 113838. [Google Scholar] [CrossRef]
  220. Gopal, G.; Ravikumar, K.V.G.; Salma, M.; Chandrasekaran, N.; Mukherjee, A. Green synthesized Fe/Pd and in-situ Bentonite-Fe/Pd composite for efficient tetracycline removal. J. Environ. Chem. Eng. 2020, 8, 104126. [Google Scholar] [CrossRef]
  221. Kazemi, M.; Jahanshahi, M.; Peyravi, M. Hexavalent chromium removal by multilayer membrane assisted by photocatalytic couple nanoparticle from both permeate and retentate. J. Hazard. Mater. 2018, 344, 12–22. [Google Scholar] [CrossRef]
  222. Liu, T.; Yang, X.; Wang, Z.L.; Yan, X. Enhanced chitosan beads-supported Fe0-nanoparticles for removal of heavy metals from electroplating wastewater in permeable reactive barriers. Water Res. 2013, 47, 6691–6700. [Google Scholar] [CrossRef] [PubMed]
  223. Jiang, D.; Huang, D.; Lai, C.; Xu, P.; Zeng, G.; Wan, J.; Tang, L.; Dong, H.; Huang, B.; Hu, T. Difunctional chitosan-stabilized Fe/Cu bimetallic nanoparticles for removal of hexavalent chromium wastewater. Sci. Total Environ. 2018, 644, 1181–1189. [Google Scholar] [CrossRef] [PubMed]
  224. Cheng, H.; Zhu, Q.; Wang, A.; Weng, M.; Xing, Z. Composite of chitosan and bentonite cladding Fe–Al bimetal: Effective removal of nitrate and by-products from wastewater. Environ. Res. 2020, 184, 109336. [Google Scholar] [CrossRef]
  225. Anju, R.P.P.; Sunil, J.T.; Dinoop, L. Chitosan stabilized Fe/Ni bimetallic nanoparticles for the removal of cationic and anionic triphenylmethane dyes from water. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100295. [Google Scholar]
  226. Rashid, S.; Shen, C.; Chen, X.; Li, S.; Chen, Y.; Wen, Y.; Liu, J. Enhanced catalytic ability of chitosan–Cu–Fe bimetal complex for the removal of dyes in aqueous solution. RSC Adv. 2015, 5, 90731–90741. [Google Scholar] [CrossRef]
  227. Silva, E.C.; Soares, V.R.; Fajardo, A.R. Removal of pharmaceuticals from aqueous medium by alginate/polypyrrole/ZnFe2O4 beads via magnetic field enhanced adsorption. Chemosphere 2023, 316, 137734. [Google Scholar] [CrossRef]
  228. Wasilewska, M.; Deryło-Marczewska, A. Adsorption of non-steroidal anti-inflammatory drugs on alginate-carbon composites—Equilibrium and kinetics. Materials 2022, 15, 6049. [Google Scholar] [CrossRef]
  229. Ragab, A.H.; Hussein, H.S.; Ahmed, I.A.; Abualnaja, K.M.; AlMasoud, N. An efficient strategy for enhancing the adsorption of antibiotics and drugs from aqueous solutions using an effective limestone-activated carbon–alginate nanocomposite. Molecules 2021, 26, 5180. [Google Scholar] [CrossRef]
  230. Ahmed, I.A.; Hussein, H.S.; ALOthman, Z.A.; ALanazi, A.G.; Alsaiari, N.S.; Khalid, A. Green synthesis of Fe–Cu bimetallic supported on alginate-limestone nanocomposite for the removal of drugs from contaminated water. Polymers 2023, 15, 1221. [Google Scholar] [CrossRef]
  231. Wu, X.; Lu, C.; Zhang, W.; Yuan, G.; Xiong, R.; Zhang, X. A novel reagentless approach for synthesizing cellulose nanocrystal-supported palladium nanoparticles with enhanced catalytic performance. J. Mater. Chem. A 2013, 1, 8645–8652. [Google Scholar] [CrossRef]
  232. Klemm, D.; Heublein, B.; Fink, H.P.; Bohn, A. Cellulose: Fascinating biopolymer and sustainable raw material. Angew. Chem. Int. Ed. 2005, 44, 3358–3393. [Google Scholar] [CrossRef] [PubMed]
  233. Baran, N.Y.; Baran, T.; Menteş, A. Fabrication and application of cellulose Schiff base supported Pd (II) catalyst for fast and simple synthesis of biaryls via Suzuki coupling reaction. Appl. Catal. A Gen. 2017, 531, 36–44. [Google Scholar] [CrossRef]
  234. Mandal, B.H.; Rahman, M.L.; Yusoff, M.M.; Chong, K.F.; Sarkar, S.M. Bio-waste corn-cob cellulose supported poly (hydroxamic acid) copper complex for Huisgen reaction: Waste to wealth approach. Carbohydr. Polym. 2017, 156, 175–181. [Google Scholar] [CrossRef]
  235. Bhardwaj, M.; Sahi, S.; Mahajan, H.; Paul, S.; Clark, J.H. Novel heterogeneous catalyst systems based on Pd (0) nanoparticles onto amine functionalized silica-cellulose substrates [Pd (0)-EDA/SCs]: Synthesis, characterization and catalytic activity toward C–C and C–S coupling reactions in water under limiting basic conditions. J. Mol. Catal. A Chem. 2015, 408, 48–59. [Google Scholar]
  236. Karami, S.; Zeynizadeh, B.; Shokri, Z. Cellulose supported bimetallic Fe–Cu nanoparticles: A magnetically recoverable nanocatalyst for quick reduction of nitroarenes to amines in water. Cellulose 2018, 25, 3295–3305. [Google Scholar] [CrossRef]
Figure 1. A schematic diagram of the selection criteria and methodology to identify related research for the systematic review.
Figure 1. A schematic diagram of the selection criteria and methodology to identify related research for the systematic review.
Metals 15 00603 g001
Figure 2. Synthesis methods for Fe-based and Al-based bimetals.
Figure 2. Synthesis methods for Fe-based and Al-based bimetals.
Metals 15 00603 g002
Figure 3. Number of publications on Fe-based and Al-based bimetals annually from 2014 to 2023.
Figure 3. Number of publications on Fe-based and Al-based bimetals annually from 2014 to 2023.
Metals 15 00603 g003
Figure 4. Most popularly utilized synthesis methods for Fe-based and Al-based bimetals.
Figure 4. Most popularly utilized synthesis methods for Fe-based and Al-based bimetals.
Metals 15 00603 g004
Figure 5. Most synthesized bimetals in the last 10 years.
Figure 5. Most synthesized bimetals in the last 10 years.
Metals 15 00603 g005
Figure 14. Schematic diagram of the formation of bimetallic iron/palladium nanoparticle (Fe/Pd NP)-assembled filter paper for transforming hexavalent chromium (Cr(VI)) to chromium irons in trivalent state (Cr(III)). Adapted from [208].
Figure 14. Schematic diagram of the formation of bimetallic iron/palladium nanoparticle (Fe/Pd NP)-assembled filter paper for transforming hexavalent chromium (Cr(VI)) to chromium irons in trivalent state (Cr(III)). Adapted from [208].
Metals 15 00603 g014
Figure 15. A schematic representation of Fe/Pd core–shell formation using pomegranate peel extract as a reducing agent. Reprinted with permission from [220]. Copyright 2020, Elsevier.
Figure 15. A schematic representation of Fe/Pd core–shell formation using pomegranate peel extract as a reducing agent. Reprinted with permission from [220]. Copyright 2020, Elsevier.
Metals 15 00603 g015
Figure 16. A schematic representation of the synthesis of CS-Fe/Ni nanoparticles. Reprinted with permission from [225]. Copyright 2020, Elsevier.
Figure 16. A schematic representation of the synthesis of CS-Fe/Ni nanoparticles. Reprinted with permission from [225]. Copyright 2020, Elsevier.
Metals 15 00603 g016
Table 1. Active countries publishing research on the synthesis of Fe-based and Al-based bimetals (2014–2023).
Table 1. Active countries publishing research on the synthesis of Fe-based and Al-based bimetals (2014–2023).
CountryFrequency% (n = 122)C/D
China7662.345.2
Russia75.716.0
Australia64.923.8
India64.916.7
Taiwan54.127.2
Iran54.164.0
Egypt43.314.0
Israel43.312.8
Argentina32.44.3
Saudi Arabia32.495.0
Note: “C/D” means the number of citations per document.
Table 2. Active journals in publishing synthesis of iron-based and aluminum-based bimetals (2014–2023).
Table 2. Active journals in publishing synthesis of iron-based and aluminum-based bimetals (2014–2023).
JournalFrequency% (n = 122)C/DCountryJournal Rank
RSC Advances86.648.0UKQ1
Chemical Engineering Journal75.770.4The NetherlandsQ1
Chemosphere75.741.6UKQ1
Applied Catalysis B: Environmental54.1203.0The NetherlandsQ1
Journal of Cleaner Production43.359.0UKQ1
Environmental Science and Pollution Research32.410.3Germany Q1
Journal of the Taiwan Institute of Chemical Engineers32.417.3TaiwanQ1
Journal of Hazardous Materials32.480.7The NetherlandsQ1
Journal of Alloys and Compounds32.414.3The Netherlands Q1
Journal of Molecular Liquids32.48.3The Netherlands Q1
Note: “C/D” means the number of citations per document.
Table 3. Active institutions in publishing synthesis of Fe-based and Al-based bimetals (2014–2023).
Table 3. Active institutions in publishing synthesis of Fe-based and Al-based bimetals (2014–2023).
InstitutionFrequency% (n = 122)Country
Tongji University75.7China
Institute of Strength Physics and Materials Science 64.9Russia
Fujian Normal University54.1China
Sichuan University54.1China
Nanjing University of Science and Technology43.3China
Technion–Israel Institute of Technology43.3Israel
Tianjin University 32.4China
China University of Mining and Technology32.4China
Chinese Academy of Sciences32.4China
Queensland University of Technology 32.4Australia
Consejo Nacional de Investigaciones Científicas y Tecnicas (CONICET) 32.4 Argentina
Table 4. Most active authors in publishing synthesis of iron-based and aluminum-based bimetals (2014–2023).
Table 4. Most active authors in publishing synthesis of iron-based and aluminum-based bimetals (2014–2023).
AuthorFrequency% (n = 122)Country
A. Sharipova43.3Israel/Russia
Xin Liu32.4China
Jing Wang32.4China
J.S. Riva32.4Argentina
Naeim Ezzatahmadi 32.4Australia
Xiulan Weng21.6China
Hongwei Wu21.6China
Bo Lai21.6China
Yuanqiong Lin21.6China
Jin-Hong Fan21.6China
Elham Aghaei21.6Australia
Table 7. Characteristics and properties of synthesized bimetals by physical methods.
Table 7. Characteristics and properties of synthesized bimetals by physical methods.
Bimetal SystemMethodCharacteristics of Synthesized Bimetals PropertiesReference
Fe-Cu/aluminum collarMechanical AlloyingPresence of Fe and Cu seen deposited on aluminum substrate; coating of up to 500 nm layer thicknessMagnetic [49]
Cu-Fe/CNTMechanical AlloyingCNT additions to Cu-Fe at 2, 5, and 10 (wt.%); Cu-Fe shown as agglomerate microparticles, larger at 2 and 5 wt.% CNT, decreased in size at 10 wt.% CNTMagnetic[50]
Fe/Ag nanocomposite Mechanical Alloying Nanocomposite structure, Fe-5Ag, Fe-10Ag (vol.%) Optimal combined strength and ductility after annealing at 550 °C, biomedical suitability [52]
Fe/Ag nanocomposite Mechanical Alloying 70% and 75% Macroporous, Fe-5Ag and Fe-10Ag (vol%) High compressive strength, high bending strength, high ductility, biodegradability [53]
Fe/Ag, Fe/Cu nanocomposites Mechanical Alloying Nanocomposite structure, Fe–10% Ag, Fe–20% Ag, and Fe–25% Cu (vol%) High strength and ductility, biodegradability[54]
Fe/Ag, Fe/Cu nanocomposites Mechanical Alloying Nanocomposite structure, Fe–10% Ag, Fe–20% Ag, and Fe–25% Cu (vol%)Densities close to theoretical values, High plasticity, bending strength[55]
Fe/CuElectrical Explosion of Metal Wires Nanostructured, near-fully dense, 72 Fe–28 Cu, 47 Fe–53 Cu, and 28 Fe–72 Cu (wt%)High yield strength (Fe-rich, 72 Fe–28 Cu): 700 MPa; high bending strength (Fe-rich, 72 Fe–28 Cu): 920 MPa; greater ductility (Cu-rich, 28 Fe–72 Cu); lower electrical resistivity (Cu-rich, 28 Fe–72 Cu)[64]
Fe/PtRadiolysisNanometer-sized fine particlesEco-friendly, improved dispersibility with PVP addition, controllability of mean particle size with PVP addition[65]
Ni/FeSonochemicalSpherically shaped nanoparticlesAdsorptive and catalytic capability, enhanced dispersity, reduced agglomeration[69]
s-Fe/Cu SonochemicalIrregularly-shaped microparticles Magnetic [70]
Fe/Cu-GOSonochemicalFe/Cu nanoparticles integrated into a GO matrix forming Fe/Cu-GO nanocompositesMagnetically recoverable, good dispersion[71]
Fe-Mn/KB SonochemicalUniform nanoparticles with the mean diameters of ∼41 nm, with distinct mesoporesSuperior oxygen reduction reaction (ORR) activity, performance comparable to commercial Pt/C electrocatalyst, stable 1.50 V voltage platform, durability over 20 h in Al–air battery tests[72]
Fe-Mn@SCAsSonochemicalFe-Mn nanoclusters with porous carbon structureAdsorptive capability, high surface area and porosity[73]
Fe/Pt, Fe/Co, and Fe/NiMF-LALNano-sized, one-dimensional (1D) chainsFerromagnetism, high saturation magnetization, low coercivity, low remanent magnetization[76]
Table 8. Bimetal synthesis by chemical reduction.
Table 8. Bimetal synthesis by chemical reduction.
Bimetal SystemExperimental MaterialsAdvantagesDisadvantagesReferences
Fe/AlMetal precursors: FeSO4·7H2O, aluminum chloride
Reducing agent: concentrated HCl
Simple and cost-effective synthesis
Efficient bimetallic composition
Controlled synthesis conditions
Use of concentrated hydrochloric acid (HCl)
Energy-intensive drying process
[79]
Fe/AlMetal precursors: ferric chloride, aluminum chloride
Reducing agent: NaBH4
Controlled sequential reduction
Versatile metal coating configurations
Improved stability of bimetallic nanoparticles
Washing and lyophilization improve purity
Complex and lengthy synthesis process
High consumption of reducing agent
[80]
Fe/Al Metal precursors: FeCl3·6H2O, AlCl3·6H2O
Reducing agent: NaOH solution
Controlled coagulation process
Adjustable Al/Fe ratio
Stable storage conditions
Effective coagulant formation
Slow preparation
process
Storage requirements
[81]
Fe/CuMetal precursors: FeCl3·6H2O and CuCl2·2H2O
Reducing agent: NaBH4
Tunable Fe/Cu ratios
Efficient reduction method
Ultracentrifugation for separation
Lyophilization for stability
Complex processing steps
High energy consumption
[82]
Fe/CuMetal precursors: FeSO4·7H2O, CuSO4
Reducing agent: NaBH4
Controlled Fe/Cu ratios
Modified borohydride method
Dropwise addition of NaBH4
Addition of ethanol in the modified borohydride method
Complex multi-step process
Additional cost for ethanol usage
[83]
Fe/CuMetal precursors: FeCl3·6H2O, CuCl2·2H2O
Reducing agent: NaBH4
Simple and cost-effective
Homogeneity of the mixture
Efficient reduction process
Mild reaction conditions
Energy-intensive drying process[84]
Fe/CuMetal precursors: CuCl2, colloidal solution of nZVI particles
Reducing agent: NaBH4
Simple synthesis at room temperature
Efficient copper deposition
Customizable Cu content
Rapid reaction time
Discontinuous Cu shell[85]
Fe/CuMetal precursors: iron (II) sulfate, copper sulfate
Reducing agent: NaBH4
Simple process
Controlled pH adjustment
Rapid reduction process
Effective removal of impurities
Reaction under nitrogen atmosphere
Excess reducing agent usage[86]
Fe/NiMetal precursors: FeSO4·7H2O, NiCl2·6H2O
Reducing agent: NaBH4
Simple and efficient process
Rapid reaction time
Controlled addition of NaBH4
Energy-intensive drying process
Multiple ethanol washing of nanoparticles
[87]
Fe/NiMetal precursors: FeCl3·6H2O, NiSO4·6H2O
Reducing agent: NaBH4
Efficient reduction process
Prevention of oxidation by use of N2 atmosphere
improved purity of nanoparticles
Energy-intensive drying process
Multiple ethanol washing steps of nanoparticles
[88]
Fe/NiMetal precursors: FeCl3·6H2O, NiSO4·6H2O
Reducing agent: KBH4
Controlled reduction process
Prevention of oxidation by use of N2 atmosphere
Efficient mixing and homogeneity
Controlled particle recovery
Improved purity of nanoparticles
Energy-intensive centrifugation step[89]
Fe/NiMetal precursors: FeSO4·7H2O, NiSO4·7H2O
Reducing agent: KBH4
Controlled reduction process
Efficient washing and purification of nanoparticles
Stable storage conditions of nanoparticles
Magnetic recoverability of nanoparticles
Energy-intensive drying process
Multiple washing steps for nanoparticles with ethanol and distilled water
Longer synthesis time
[90]
Fe/NiMetal precursors: FeCl3·6H2O, Ni(NO3)2·6H2O
Reducing agent: NaBH4
Simple and scalable process
Room-temperature processing
Mechanical stirring for homogeneity
Controlled NaBH4 addition
Efficient washing and purification
Use of N2 atmosphere
Multi-step process [91]
Fe/TiMetal precursors: FeSO4·7H2O, Ti(SO4)2
Reducing agent: NaBH4
Efficient reduction process
Short reaction time
Centrifugal selection of Fe-Ti nanoparticles
Use of vacuum-drying to preserve nanoparticles
Use of N2 atmosphere in the process
Long vacuum-drying time [92]
Fe/CoMetal precursors: Fe(NO3)3·9H2O, FeCl3·6H2O, Co(NO3)2·6H2O, CoCl2·6H2O, CoSO4·7H2O
Reducing agents: NaOH and hydrazine hydrate (N2H4·7H2O)
Simple and efficient synthesis
Controlled Fe/Co ratio
Magnetic recoverability of nanoparticles
Low-temperature drying
Multiple ethanol washing steps [93]
Fe/MnMetal precursors: iron (II) chloride hexahydrate, manganese chloride
Reducing agent: sodium tetraborate
Simple and cost-effective method
Controlled bimetallic composition
Controlled addition of sodium tetraborate solution
Moderate drying temperature
Long drying time [94]
Table 10. Carbon-based materials as supports for synthesis.
Table 10. Carbon-based materials as supports for synthesis.
Bimetal SystemExperimental MaterialsAdvantagesDisadvantagesReference
Fe-Ce/NCNTMetal precursors: FeCl3·6H2O, Ce(NO3)3·9H2O
Other(s): melamine
Support: nitrogen-doped carbon nanotubes (NCNT)
Controlled bimetallic composition
High-temperature pyrolysis in the process
Ultrasonication with HCl
Protective argon atmosphere
Time-consuming process
High energy consumption
Use of strong acid (HCl) for purification
Equipment-dependent process
[161]
Fe-Al/MWCNTsMetal precursors: Fe(NO3)3·9H2O, Al(NO3)3·9H2O
Other(s): citric acid
Support: multi-walled carbon nanotubes (MWCNTs)
Simple catalyst preparation
Controlled bimetallic composition
Involvement of calcination
Use of N2 atmosphere
High energy consumption [162]
Fe-Cu/CNFMetal precursors: iron (III) acetylacetonate (Fe(acac)3), copper (II) acetate monohydrate (Cu(ac)2·H2O)
Others: polyacrylonitrile (PAN), dimethylformamide (DMF)
Support: carbon nanofibers (CNFs)
Controlled morphology and composition
Strong metal-support interaction
Good thermal stability
Enhanced mechanical properties
Complex synthesis process
High energy consumption
[163]
Fe-Cu/Graphitic carbon Metal precursors: Fe(II) acetylacetonate, chlorophyllin (Cu precursor)
Support: graphitic carbon
Simple and scalable synthesis
Use of natural precursor (chlorophyllin)
Controlled heating profile
Use of high-temperature treatment
Use of argon (Ar) atmosphere
High energy consumption[164]
Cu-Fe/MC Metal precursors: Fe(NO3)3·9H2O,
Cu(NO3)2·3H2O
Others: Pluronic F127, phenol and formalin solution (carbon precursors)
Support: mesoporous carbon (MC)
Controlled metal loading
Improved thermal stability
Porous carbon structure
High energy consumption
Long processing time
[165]
Co-Fe/MBMetal precursors: FeSO4·7H2O, CoSO4·7H2O
Others: pristine sawdust biochar, PEG-4000 (dispersant), NaBH4 (reducing agent)
Support: modified biochar (MB)
Enhanced stability of Co/Fe nanoparticles
Controlled reduction process
Use of biochar as support
Enhanced metal–support interaction
Multiple ethanol washing steps
Long processing time
Energy-intensive drying
[166]
Ag-Fe/MB Metal precursors: FeSO4·7H2O, AgNO3
Others: original biochar (OB), NaBH4 (reducing agent), PEG-4000
Support: modified biochar (MB)
Versatile synthesis approach
Controlled stirring of the mixture
Controlled addition of NaBH4 solution
Use of biochar as support
Complex multi-step process
High energy consumption
Long processing time
[167]
BC@Fe/NiMetal precursors: FeCl2, NiCl2
Others: ground straw (biochar source), NaBH4 (reducing agent), polyethylene glycol (PEG) (dispersant)
Support: biochar (BC)
Biochar as a support material
Controlled metal deposition
Controlled biochar loading
Magnetic recoverability of the particles
Multi-step synthesis procedure
High energy consumption
[168]
Fe–Co-modified biochar (FMBC)Metal precursors: Fe(NO3)3 9H2O,
Co(NO3)2·6H2O
Other(s): cedar bark (biochar source)
Support: modified biochar (MBC)
Forestry waste (cedar bark) as a precursor for biochar
Biochar as a support material
High temperature pyrolysis
Magnetic recoverability of the adsorbents
High energy consumption [169]
nZVIC (Fe-Cu)-municipal sludge-derived biochar (SBC)Metal precursors: FeSO4·7H2O,
CuSO4·5H2O
Others: modified sewage sludge (biochar source), NaBH4
Support: municipal sludge-derived biochar (SBC)
Low-temperature synthesis
Biochar as a support material
Controlled reduction conditions
Municipal sludge utilization
Use of excess NaBH4 solution[170]
Zn-Fe/CBCMetal precursors: FeCl3·6H2O,
ZnSO4·7H2O
Other(s): corncob (biochar source)
Support: corncob biochar (CBC)
Use of corncob as biochar source
Biochar as a support material
Moderate-temperature pyrolysis (450 °C)
Centrifugation and washing ensuring selection of Zn-Fe/CBC
Long processing time
Energy-intensive drying for biochar preparation
Multiple washing steps
Energy-intensive drying synthesis requirement
Pyrolysis energy demand
[171]
Fe-Co/ACMetal precursors: FeSO4·7H2O,
CoCl2·6H2O
Other(s): scrap tires (activated carbon source), NaBH4
Support: activated carbon (AC)
Use of scrap tires as activated carbon source
Versatile synthesis method
Controlled reduction with NaBH4
Uses water as the primary solvent in the process
High temperature requirements during synthesis
Long processing time
[172]
Fe-Ce/AC Metal precursors: Fe (NO3)3·9H2O,
Ce (NO3)3·6H2O
Other(s): waste rubber tires (WRT) (activated carbon source)
Support: activated carbon (AC)
Use of waste rubber tires (wrt) as activated carbon source
ensured
metal–carbon interaction
Controlled precipitation conditions
Moderate calcination temperature (350 °C)
Energy-intensive process
Long processing time
[173]
nZVI-Ni/ACMetal precursors: FeSO4·7H2O, NiCl2·6H2O
Others: KBH4 (reducing agent), PEG-4000
Support: activated carbon (AC)
Controlled reduction process
Improved dispersion with peg-4000
Use of inert gas environment in the synthesis
Multi-step process
High energy requirements
Long processing time
[174]
PAC-Fe/AgMetal precursors: FeSO4·7H2O,
AgNO3
Others: powder activated carbon, NaBH4
Support: powder activated carbon (PAC)
Versatile synthesis method
Controlled reduction with NaBH4
Improved Ag adhesion onto ZVI/PAC
Magnetic recoverability of the bimetallic nanoparticles
Muti-step process
High energy requirements
Use of excess NaBH4 solution
[175]
Ag-Fe/CACMetal precursors: FeSO4·7H2O,
AgNO3
Others: commercial activated carbon (CAC), NaBH4 (reducing agent),
polyethylene glycol 600 (PEG-600)
Support: commercial activated carbon (CAC)
Simple versatile synthesis procedure
Mild synthesis conditions
Involves uncontrolled stirring [176]
Fe-Cu/CACMetal precursors: FeSO4, FeCl3, CuSO4
Others: coconut husk (activated carbon source), NaBH4 (reducing agent)
Support: coconut husk-derived activated carbon (CAC)
Use of coconut husk as activated carbon source
Versatile synthesis procedure
(Sequential metal impregnation)
Muti-step synthesis procedure
Process complexity
High energy requirements
Multiple chemical consumption
[177]
Table 12. Characteristics and properties of synthesized bimetals by chemical methods.
Table 12. Characteristics and properties of synthesized bimetals by chemical methods.
Bimetal SystemMethod Characteristics of Synthesized BimetalsPropertiesReference
Fe-AlChemical reductionSpherically shaped nanoparticles Adsorptive and catalytic capability[80]
Fe-CuChemical reductionNanoclusters, Fe/Cu mass ratios (0.9:0.1, 0.75:0.25, and 0.5:0.5)Magnetic recoverability, adsorptive removal capacity [82]
Fe-CuChemical reduction Core–shell structure, discontinuous Cu shell on an nZVI coreCatalytic capability [85]
Fe-CuChemical reduction Spherically shaped nanoparticlesDegradation capability [86]
Fe-NiChemical reduction Nanoparticles with chain-like structureMagnetic recoverability, adsorption capability [87]
Fe-NiChemical reduction Ni particles dispersed on Fe nanoparticle surface Magnetic recoverability, adsorptive and reductive capability [90]
Fe-Co Chemical reduction Spherically shaped nanoparticles Magnetic recoverability, adsorptive capability [93]
NP-Pd/FeChemical dealloying Open nanosponge structureElectrocatalytic capability [101]
NP-Pt/Fe alloyChemical dealloying Interconnected strips with nanoporous morphology with pore size of
about several nanometer
Electrocatalytic capability [102]
Ag/FeSeed-mediated growth Microparticles with diameters of 2–10 μmCatalytic capability [105]
Cu/Al Seed-mediated growth Core–shell structure, Cu particles deposited as rod-like aggregations on the aluminum surfacesDegradation capability [106]
Fe/RhElectrochemical synthesis Dispersed nanowires about 18 nm in diameter and 1 mm longLow temperature magnetic capability [118]
Fe/RhElectrochemical synthesis polycrystalline nanowire arrays, 20 nm in diameter, and about 1–3 mm in lengthMagnetic properties [119]
Fe-CuGalvanic replacementDispersed Cu particles on Fe surface Catalytic capability[123]
Fe-CuGalvanic replacementNanoparticles had an irregular surface structure with particle sizes ranging from 20 to 30 nmAdsorptive capability[124]
Fe-AlGalvanic replacementFine Fe particles (about 200–400 nm) on the surface of Al were necklace-like or ball-like, the size of the bimetal was about 20–30 μmAdsorptive and catalytic capability[131]
Fe-AlGalvanic replacementCore–shell structure, Fe particles deposited on Al surface Adsorptive capability [133]
Mg-FeGalvanic replacementMg/Fe particles were indicated by many sheet crystal particlesDegradation capability, electrochemical property [134]
Fe-Al Galvanic replacementCore–shell structure, Fe particles deposited on Al surface Adsorptive capability [137]
Fe-Cu Galvanic replacementMicro-scale particles, Cu particles deposited on Fe surfaceDegradation capability [139]
Ni/FeThermogravimetric method Nanoporous structureCatalytic capability [152]
Ni/FeThermogravimetric method Spherically shaped nanoparticlesCatalytic capability [157]
Fe/Ni NPsThermogravimetric method Nanoparticles have an average size of 18.6 ± 2.4 nm, with Ni observed on their surfaceCatalytic capability, magnetic property [158]
Fe-Ce/NCNTSupported particles Hollow CNTs encapsulated nanocrystals structure was evident, along with the ordered carbon layer distribution; and CNTs diameter distribution was between 100 and 200 nmElectrocatalytic capability [161]
Fe-Al/MWCNTsSupported particlesStraight and long carbon nanotubes (MWNTs) appear on the Fe-Al catalyst, MWNTs have a graphite interlayer spacing of 0.34 nmCatalytic capability, electronic conductivity [162]
Ag-Fe/MB Supported particlesDispersion of Ag/Fe NPs (estimated diameter equal to 51 nm) on biochar surface, as well as the formation of small globular structuresAdsorptive and reductive capability [167]
BC@Fe/NiSupported particlesFe/Ni NPs existing in chain forms and distributed in some pores or other places of BCMagnetic recoverability, adsorptive capability [168]
Ni3Fe/Al2O3Supported particlesNi-Fe nanoparticles are well dispersed on the Al2O3 support Catalytic capability [179]
Pd-Fe/Al2O3/PVDFSupported particlesPd/Fe NPs seen on the surface of the Al2O3/PVDF membrane, exhibiting a smooth, spherical morphology with particle sizes ranging from approximately 50 to 100 nmDegradation capability [182]
FeCu/HMSSupported particlesFe and Cu evenly dispersed throughout the silica matrix, metal nanoparticles in the Fe-Cu/HMS having an average size of approximately 18 nmCatalytic and degradation capability [185]
Fe-Al-SBA-15Supported particlesWell-ordered hexagonal arrays of mesopores with one-dimensional channels, also agglomeration of Fe seen clearly, indicating considerable FexOy clusters inside the channelCatalytic capability [188]
B-Fe/NiSupported particlesFe/Ni spherical particles ranging in size from 30 to 60 nm well dispersed on bentoniteAdsorptive and catalytic capability [194]
K-Fe/PdSupported particlesFe/Pd particles with a diameter of about 20–70 nm, and consisting of short chain-like spherical particles seen on kaoliniteCatalytic capability[197]
Di-Fe/NiSupported particlesporous structure of Di-Fe/Ni, dispersion of Fe/Ni spherical nanoparticles (in the range of 50–80 nm) into the pores and on the surface of diatomiteCatalytic and degradation capability[201]
Pal-Fe/Ni Supported particlesFe/Ni spherical nanoparticles, with a diameter range of 20–60 nm, well dispersed and stabilized onto the palygorskite surfaceCatalytic and degradation capability[202]
Fe/Pd-assembled filter paperSupported particlesFe/Pd white, quasi-spherical NPs (mean diameter of 10.1 ± 1.7 nm) distributed homogeneously onto the filter paper surfaceCatalytic and reduction capability[208]
Cu/Fe-BM@FASupported particlesFe and Cu detected on FA spherical microparticles Adsorptive and catalytic capability[209]
Table 14. Characteristics and properties of synthesized bimetals by biological methods.
Table 14. Characteristics and properties of synthesized bimetals by biological methods.
Bimetal SystemMethod Characteristics of Synthesized BimetalsPropertiesReference
Fe-NiBiologicalCore–shell structure, Nanosphere with some irregularly-shaped nanoparticlesAdsorptive capacity[215]
Fe-PdBiologicalSpherically shaped nanoparticlesCatalytic capability, adsorptive capacity[216]
Fe-CuBiologicalSpherically shaped nanoparticlesAdsorptive capacity, reduction capability[217]
Fe-CuBiologicalSpherically shaped nanoparticlesAdsorptive capacity[218]
C-Fe-NiBiologicalPolydisperse regular spherical nanoparticlesCatalytic capability[219]
Fe-PdBiologicalcore–shell structure Spherically shaped nanoparticlesCatalytic capability[220]
Chitosan (CS)-stabilized Fe-CuBiologicalSpherically shaped nanoparticlesCatalytic capability[223]
Fe–Al bimetal chitosan bentonite (Fe–Al bimetal@bent) complexBiologicalPlenty of bimetal surrounded by chitosan (Cs)–bentoniteAdsorptive capability[224]
Chitosan (CS)-Fe-NiBiologicalNanoparticle, Fe as core, chitosan as shellCatalytic capability[225]
Chitosan–Cu–Fe bimetal complexBiologicalIrregular and relatively nonporous complex with presence of Fu and CuCatalytic capability[231]
ZVFe–Cu/Alg–LSBiologicalNanocomposite exhibiting multilayer structure and rough surfaceAdsorptive and catalytic capability[230]
Fe–Cu@MCCBiologicalNanocomposite particles within range of 27–35 nmCatalytic capability, magnetic recoverability[236]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Balangao, J.K.B.; Tabelin, C.B.; Phengsaart, T.; Zoleta, J.B.; Arima, T.; Park, I.; Mufalo, W.; Ito, M.; Alorro, R.D.; Orbecido, A.H.; et al. Synthesis of Iron-Based and Aluminum-Based Bimetals: A Systematic Review. Metals 2025, 15, 603. https://doi.org/10.3390/met15060603

AMA Style

Balangao JKB, Tabelin CB, Phengsaart T, Zoleta JB, Arima T, Park I, Mufalo W, Ito M, Alorro RD, Orbecido AH, et al. Synthesis of Iron-Based and Aluminum-Based Bimetals: A Systematic Review. Metals. 2025; 15(6):603. https://doi.org/10.3390/met15060603

Chicago/Turabian Style

Balangao, Jeffrey Ken B., Carlito Baltazar Tabelin, Theerayut Phengsaart, Joshua B. Zoleta, Takahiko Arima, Ilhwan Park, Walubita Mufalo, Mayumi Ito, Richard D. Alorro, Aileen H. Orbecido, and et al. 2025. "Synthesis of Iron-Based and Aluminum-Based Bimetals: A Systematic Review" Metals 15, no. 6: 603. https://doi.org/10.3390/met15060603

APA Style

Balangao, J. K. B., Tabelin, C. B., Phengsaart, T., Zoleta, J. B., Arima, T., Park, I., Mufalo, W., Ito, M., Alorro, R. D., Orbecido, A. H., Beltran, A. B., Promentilla, M. A. B., Jeon, S., Haga, K., & Resabal, V. J. T. (2025). Synthesis of Iron-Based and Aluminum-Based Bimetals: A Systematic Review. Metals, 15(6), 603. https://doi.org/10.3390/met15060603

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