Next Article in Journal
Study on the Distribution Characteristics and Influencing Factors of Traditional Villages in the Yunnan, Guangxi, and Guizhou Rocky Desertification Area
Next Article in Special Issue
Study on Co-Pyrolysis of Coal and Biomass and Process Simulation Optimization
Previous Article in Journal
The Impact of Land Use/Cover Change on Ecological Environment Quality and Its Spatial Spillover Effect under the Coupling Effect of Urban Expansion and Open-Pit Mining Activities
Previous Article in Special Issue
Research on Magnetic Rollers for Recovering Non-Ferrous Metals from End-of-Life Vehicles Employing Machine Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sourcing, Refining and Recycling of Rare-Earth Magnets

1
John Ormerod Consulting Limited Liability Company, Loudon, TN 37774, USA
2
Critical Materials Innovation Hub, Ames National Laboratory, Division of Critical Materials, US DOE, Ames, IA 50011, USA
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(20), 14901; https://doi.org/10.3390/su152014901
Submission received: 26 July 2023 / Revised: 27 September 2023 / Accepted: 28 September 2023 / Published: 16 October 2023
(This article belongs to the Special Issue Waste Recycling and Circular Economy: From Trash to Treasure)

Abstract

:
Permanent magnets today are used in a wide range of transportation, industrial, residential/commercial, consumer electronics, defense, domestic, data storage, wind energy, and medical markets and applications. There are five classes of commercial permanent magnet materials; however, magnets based on Nd-Fe-B account for over 60% of the global magnet production by value. They typically contain around 31 wt.% of rare earth elements (REEs), principally, Nd and Pr, plus Dy for higher-temperature performance. Nd-Fe-B magnets are forecast to grow throughout this decade, largely driven by the growth in electric vehicles of all types. However, several studies forecast a shortfall of the primary REEs from mined resources. In this paper, the sourcing, processing, and recycling of REEs are discussed. Additionally, presented are the advantages and disadvantages of the major recovery and recycling technologies for REEs.

1. Rare-Earth Elements (REE): Sources, Classification, Concentration, and Refinement

Rare-earth elements (REEs) are a group of 15 elements in the periodic table, ranging from lanthanum to lutetium, also known as lanthanides, plus scandium and yttrium. Depending on their atomic numbers, REEs can be separated into light rare-earth elements (LREEs) and heavy rare-earth elements (HREEs). LREEs include lanthanum, cerium, praseodymium, neodymium, samarium, and europium, and HREEs are defined by the suite of lanthanides that include gadolinium, terbium, dysprosium, holmium, erbium, thulium, ytterbium, and lutetium, with those elements most used in high-performance permanent magnet applications highlighted with horseshoe magnet symbols in Figure 1 [1,2,3,4,5].
The REE resources can be broadly classified into primary and secondary sources. Primary REE sources are predominantly mineral ores produced by magmatic, hydrothermal, or metamorphic processes [6]. Commercial extraction of REEs is dominated by a few mineralogies, including bastnaesite, ion-adsorption clays, monazite, and xenotime, which account for over 95% of economic production. The LREEs are predominantly recovered from mineral concentrates of monazite and bastnaesite sourced from operations in China, the USA, Australia, India, and Madagascar, and to a lesser extent from the mineral loparite extracted in Russia. Most HREEs are produced from ion-adsorption clays and xenotime mineralization in Southern China, Myanmar, and Australia. Though xenotime is less abundant than other minerals, it is a primary source of HREEs, particularly dysprosium. Apatite has also been mined in its non-weathered state for REEs [1]. Table 1 shows a list of naturally occurring REE-containing ores. Bastnaesite, iimoriite, monazite, parisite, and xenotime are the ones that contain the maximum fraction of REEs [7].
Secondary REE sources, on the other hand, originate from industry byproducts that otherwise end up in landfills owing to their dilute concentrations [8,9]. Furthermore, secondary sources of REEs also include electronic wastes and several other useful commercial commodities. Thus, while the idea of primary REE sources is clear, the definition of secondary sources must be clarified for consistent referencing. In the present report, we thus broadly classify the secondary REE sources into two major categories, namely: (a) unprocessed and (b) processed sources. The unprocessed sources are REE-containing materials without any prior commercial use. They include mine tailings, coal ash, phosphogypsum, red mud, and marine sediments [10]. Conversely, the processed sources relate to REEs that have already been used for certain applications. These include REEs in catalysts, permanent magnets, polishing materials, fluorescent and LED lamps, metal hydride batteries, and electrical appliances and other applications [11].
REE sources are also categorized based on conventional and unconventional sources. While ores containing REEs have been strictly categorized as conventional sources, mine tailings and byproducts from coal ash, apatite, and phosphates have been labeled unconventional REE sources [12]. E-waste generation, continually rising over the years, is also seen as an unconventional REE source. However, with the fast depletion of conventional REE sources, unconventional sources may become the conventional and primary sources.
Furthermore, the primary REE sources are sparsely dispersed worldwide. A recent US Geological Survey report suggests that four countries possess more than 85% of the world’s natural REE resources [13] (Figure 2, Table 2). Figure 3 shows a global map of countries with REE reserves. Countries such as the United States are already mining their REEs at an unprecedented rate, which might bring forth the REE crisis in the coming decades [13] (Figure 2, Table 2). This has thus brought forth the idea of urban mining, which looks at recycling e-wastes to recover rare-earth resources as an alternative approach [14]. “Mining” REE from e-wastes can be more environmentally friendly than mining from virgin resources [15]. In the present report, we suggest demarcating REE reserves into two categories: primary and non-primary. The non-primary resources would include all the secondary and unconventional REE resources.
Several projects are currently assessing REE from various unconventional sources that will propel countries into more self-reliance and help alleviate their need to rely on imports of such materials. Table 3 briefly lists key activities from coal byproducts and acid mine drainage to extract REEs.
Concentrates of REE minerals are exclusively processed into refined products, either as mixed/semi-separated compounds or individual rare-earth compounds. Most refining occurs within China, Malaysia, Russia, and India, with small amounts separated in Vietnam, Norway, and Australia. Refined REEs have a range of end uses, but increasingly, high-powered permanent magnet applications dominate the sales value of REEs. Today, the REEs used in permanent magnets (Nd, Pr, Dy, and Tb) account for over 90% of the value of all REEs processed [6].
Over 50% of REE demand has historically come from China, other Asian countries, Australia, and North America. Under this context, global demand for REEs increased from below 157 kt of rare-earth oxide (REO) in 2017 to 256 kt REO in 2020, expecting to reach 305 kt REO by 2025 [29]. Due to the 2019 COVID-19 pandemic and its effects, the REE market contracted by 1% in 2020; however, its overall increase in demand has seen REE consumption reach record highs, enhanced mainly by rapid growth in permanent magnet applications [30,31]. In 2022 and early 2023, China still accounted for 70% of REE production and 90% of processing [32]. In general, REEs’ demand will be increasingly driven by their use in permanent magnets for the electrification of transport and the transition to renewable energy generation. This is evident from the more than 40% increase in refined REEs for permanent magnets between 2020 and early 2023 [32]. Simultaneously, the transitioning away from fossil fuels in transport and energy generation will reduce the REE demand from catalyst applications in both petroleum processing and emissions control systems for passenger and commercial vehicles [33].
Permanent magnet applications play a critical role in supporting the energy transition. Transport, energy-efficient equipment, and energy generation applications will cause significant growth in the demand for key magnet elements, with Nd and Pr being the most impacted, as they are the main rare-earth constituents of Nd-Fe-B alloys by volume. Short-term demand growth for Tb and Dy will accelerate in the period to 2030 before slowing, as the development of new magnet designs and production methods reduce the required HREE intensity in high-performance magnets. Replacement of some Nd-Pr by La and Ce tends to reduce the performance of the magnets, which may be unsuitable for the highest growth markets, such as automotive drivetrains. Using La and Ce in such mid- to low-quality magnets can increase their demand in the future.
Once mined, RE minerals are processed into concentrates containing multiple REEs, which need to be separated from each other (Figure 4), typically by solvent extraction. REEs are chemically very similar, so separation often requires a series of extractions using multiple solvents to separate desired individual or compound REEs. This step in the process involves large amounts of acid, water, and radioactive byproducts, so obtaining adequate solvents and treatment of waste are significant cost drivers. REOs must be further refined or reduced to metal before they can be used for magnet production (Figure 4). Electrowinning is the most common process for converting REOs into their metallic state, while ensuring low-impurity contents, particularly oxygen, nitrogen, and carbon [33,34].

2. Rare-Earth Permanent Magnets and Applications

The special technological importance of permanent magnets derives from their ability to produce a magnetic field, making them suitable for various transportation, industrial, residential/commercial, consumer electronics, defense, information technology, power generation, and medical applications. Unlike electromagnets that require a continuous electrical current to be supplied to generate a magnetic field and function as a source of magnetic flux [31,35,36], permanent magnets provide a magnetic flux with no external energy input.
Not all permanent magnets utilize REEs. There are four general classifications of permanent magnets, each with a range of performance metrics and typical end-uses, and these include ferrite, alnico, samarium-cobalt (Sm-Co), and neodymium-iron-boron (Nd-Fe-B). Nd-Fe-B-based magnets are the backbone of expansion in alternative energies, although their performance at high operating temperatures is limited. This limitation is typically overcome by adding Dy or Tb, which increases the coercivity and performance at higher temperatures. Several extensive reviews are available that cover the history of REE permanent magnet development [37,38,39,40,41,42,43]. Nd-Fe-B magnets typically contain around 32 wt.% of REEs, principally, Nd and Pr, plus Dy and Tb for higher-temperature performance. Figure 5 shows the estimated global production of permanent magnets by material type [31].
Together, the ferrite- and Nd-Fe-B-based magnets account for over 90% of the global production of permanent magnets. Despite the significant difference in magnetic performance between the two materials, their respective price-to-magnetic performance ratio is extremely favorable compared to all other material options. The high energy density of Nd-Fe-B makes them the magnet of choice for many energy-efficient and renewable energy applications, e.g., EV drivetrain motors, wind power generators, and HVAC (heating, ventilation, and air conditioning) units. Consequently, the market for these magnets is forecast to grow from 200,000 tons per year in 2022 to 450,000 tons in 2030 [31]. The details, by major markets, are shown below in Figure 6.
As can be seen, the growth in demand is largely driven by renewable energy applications. Unfortunately, the major REEs used, namely Nd and Pr, and for higher-temperature applications, Dy and Tb, are forecast to be in short supply, considering these aggressive demand forecasts. For example, the Adamas Intelligence forecast indicates that by 2040, the undersupplies of Nd-Pr, Dy, and Tb oxides will reach 90 kt, 1.8 kt, and 0.45 kt, respectively [44]. Nd-Fe-B magnets are forecast to grow at a CAGR of over 10% throughout this decade, largely driven by the growth in electric vehicles of all types. Several market studies are predicting that combined production of electric vehicles and hybrid electric vehicles (EVs/HEVs) will be in the range of 50 to 100 million units produced annually by 2030. Current EV/HEV motors contain, on average, 1.8 kg of Nd-Fe-B-based magnet material per vehicle [45]. This equates to over 90,000 mt of Nd-Fe-B magnet material at the low end of the forecast consumption for this one application in 2030. The preferred motor type used in EV traction drives is the interior permanent magnet (IPM) design. In this design, the magnets are installed into slots in the laminated rotor [45].
Figure 6. Major growth markets for Nd-Fe-B magnets through 2030. Adapted from [45].
Figure 6. Major growth markets for Nd-Fe-B magnets through 2030. Adapted from [45].
Sustainability 15 14901 g006
It is generally accepted that the magnetic parameter, maximum-energy product (BH)max, is the best all-round single indicator of permanent magnet performance. From the first principles, it can be shown that for a magnetic circuit containing an airgap, the energy stored in the field in the air gap is directly proportional to the product of flux density, B, and the corresponding field strength, H, at any point on the second-quadrant normal demagnetization curve, as shown in Figure 7 below [46]. The maximum value of this product, i.e., (BH)max, or the maximum-energy product, can be directly related to the maximum energy that can be generated with a permanent magnet [47].
The following chart (Figure 8) shows the historical development and commercialization of permanent magnets based on their maximum-energy product, (BH)max. This chart clearly demonstrates the improvement in magnetic performance, beginning with magnet steels in the early 1900s, to alnico’s, and finally, hard ferrites and rare-earth magnets. It is now over 40 years since the US announcement of Nd-Fe-B magnets at the 29th MMM conference held in Pittsburgh, PA, in November 1983 [22]. It is interesting to note that since the discovery and introduction of Nd-Fe-B magnets, no major new sintered permanent magnet material has been introduced.

3. Sintered Nd-Fe-B-Based Magnet Processing

Figure 9 outlines the basic steps for the powder metallurgical processing of sintered Nd-Fe-B-based magnets [49,50]. The metallization step involves the production of the REE metals through the reduction (mostly electrolytically) of RE2O3 (RE = Nd, Pr, and Dy, or their mixtures). Electrolytic reduction is accomplished by passing electric current through electrodes in a molten mixture of the oxide and a fluxing agent. Keeping the oxygen, nitrogen, sulfur, and carbon levels low is an important consideration for the subsequent permanent magnet production because removing any impurity later in the process is impractical.
The next step is the preparation of the master alloy using Nd, Pr, and Dy metals (or their alloys), along with iron, cobalt, ferro-boron, and in many cases, reclaimed and recycled alloy material. Other additive elements (Al, Cu, Ga, etc.) are included to enhance grain boundary refinement. The master alloy is subsequently melted and rapidly cooled into flakes, preferably via the strip-casting process (Figure 10). This rapid cooling results in the right phase and microstructure for developing the optimum magnetic properties in the finished magnet. This is critically important because slow cooling favors the formation of αFe, which makes crushing and grinding the ingot into a powder much more difficult and deteriorates the magnetic performance of a permanent magnet [51]. How to eliminate the formation of αFe is part of the lessons learnt in the early stages of the development of sintered Nd-Fe-B magnets. At the time, conventional metal molds were used to cast the alloys, resulting in relatively slow cooling rates and significant αFe formation. The solution was the development of the strip-casting process for Nd-Fe-B alloys.
The next step is the crushing of the strip-cast alloy to a coarse powder. Today, this typically uses the hydrogen decrepitation process, which is shown diagrammatically in Figure 11 [49]. In this step, the alloy is exposed to hydrogen gas pressure, during which it absorbs hydrogen in the grain boundaries, resulting in decrepitation into coarse powder. The coarse alloy powder is afterwards subjected to jet milling, during which it is comminuted to a fine powder with an average particle size of 5 µm (Figure 12). The objective of jet milling is to produce a powder consisting of single-crystal particles with a narrow particle size distribution. A typical jet milling system used for Nd-Fe-B magnets is schematically represented in Figure 13 [52].
Most commercially available sintered magnets are anisotropic, i.e., the magnetic easy axes of the powder particles (ideally single crystalline) are aligned in one direction. This helps to achieve a high remanent magnetization (hence, high (BH)max) and a more coherent rotation of the magnetic moments during demagnetization. For sintered magnets, this is accomplished by applying a magnetic field and compacting the powders in the magnetic field, such that the easy axes of magnetization are parallel. Figure 14 depicts a typical pulsed magnetizer equipment. The green body thus prepared is then ready for densification by sintering.
The sintering of Nd-Fe-B permanent magnets is typically undertaken in a vacuum furnace under a partial pressure of an inert gas such as argon. The vacuum furnace may need to be purged with an inert gas to minimize any oxygen and moisture contents. It is important to maintain a constant and well-controlled sintering temperature to ensure the magnet has no open porosity, as well as to prevent oxidation and aging during use. This requires the sintered density of the Nd-Fe-B magnet to be greater than 98% of the theoretical density.
A final thermal treatment is applied to the sintered Nd-Fe-B magnets to develop the microstructure for good hard magnetic properties. The addition of additive elements such as Al, Cu, and Ga (or their alloys) during the Nd-Fe-B alloy preparation is critical for this stage. For this, the Nd-Fe-B magnet alloys are prepared to be RE-rich, i.e., RE2+xFe14B1+y, such that the excess RE combines with the additive elements to form low melting (typically eutectic) phases in the grain boundary. The final thermal treatment leverages the low melting temperature (500–700 °C, depending on the constituent REs) of the phases to optimize the grain boundary for an enhanced hard magnetic performance, without significant grain growth. Grain growth deteriorates the coercivity of Nd-Fe-B, especially if non-uniform grain growth occurs. The shrinkage that occurs as part of the densification during sintering results in some magnet size variation, and therefore, final machining operations are necessary to achieve the required dimensional size and tolerances. Most times, magnets of smaller dimensions are machined from a larger sintered block.
Despite this commercial success, Nd-Fe-B magnets suffer from significant problems, including a high thermal coefficient of coercivity, β, high thermal coefficient of remanent magnetization, α, and poor corrosion resistance. β and α represent the percentage change in coercivity and remanence, respectively, with temperature, and are critical performance parameters. For Nd-Fe-B magnets, β can be up to −0.65%/°C, compared to −0.3%/°C in Sm-Co magnets and −0.01%/°C in alnico magnets. α in Nd-Fe-B magnets is also higher (−0.1%/°C), compared to −0.05%/°C in Sm-Co magnets and −0.02%/°C in alnico magnets. The negative sign is used to indicate a decrease in the properties; hence, the lower the absolute value, the better. The higher absolute values of β and α limit the maximum operating temperature of the Nd-Fe-B in application. Combining improved processing, microstructure refinement, grain boundary modification, alloy chemistry, and hermetic coatings has virtually eliminated the corrosion resistance issue. However, it is interesting to note that today’s solution for increasing the operating temperature of a Nd-Fe-B magnet was discovered in 1984; namely, the addition of HREE elements and, in particular, Dy and Tb (see Figure 15). As can be seen, a small addition of Dy significantly increases the intrinsic coercivity, HcJ, compared to the standard magnet composition [53]. The increase helps to ensure that there is sufficient coercivity for elevated-temperature applications.
Figure 16 shows an example of the demagnetization behavior of sintered Nd-Fe-B grade N55M (no Dy) at different temperatures. Note that both remanence and coercivity decreased with the increasing temperature. This grade loses around 50% of the room temperature coercivity at 100 °C, a major concern in any device operating at elevated temperature. Therefore, it has become common practice in the Nd-Fe-B magnet industry to add Dy or Tb to increase the intrinsic coercivity. Figure 17 shows the typical Nd-Fe-B grade designations developed based on Dy content and the corresponding maximum recommended operating temperature. It shows that increasing the Dy content increases the maximum operating temperature from around 80 °C to 220 °C.
Following the severe disruptions to the REE supply chain between 2011 and 2012, extensive R&D efforts were undertaken to find processing methods for minimizing the Dy content for higher-temperature and high-coercivity Nd-Fe-B grades. A major advancement is to diffuse Dy from the surface of a sintered magnet, instead of alloying it in the entire magnet volume (Figure 18). The goal is to increase the anisotropy field, and hence coercivity, by enriching the regions near the grain boundary of the main Nd2Fe14B phase with Dy. One of many variants is the TDK HAL (high anisotropy field layer) process [56]. In the HAL process, the Dy source is located on the surfaces of the sintered magnets and a low-temperature heat treatment under vacuum is applied, resulting in the solid-state diffusion of Dy around the grain boundary regions of the sintered magnet. This is shown diagrammatically in Figure 15. TDK claims the HAL process results in a 20% to 50% reduction in Dy content, with a 3% to 5% increase in Br. Today, it is common to find grain-boundary-diffused grades offered by all the major Nd-Fe-B manufacturers.

4. Introduction to REE Recycling

Some amount of recycling has been part of manufacturing Nd-Fe-B since its commercialization in the mid-1980s. Going forward, the key will be the development of economically viable EOL magnet recovery and recycling processes. Today, this is a minor contribution to the total REE supply, but it needs to grow rapidly throughout this decade to become an important source of REEs.
Figure 19 shows the electric motor in a 2022 BMW X5 PHEV, which can be seen to be embedded in the drive train and surrounded by an aluminum alloy housing [57]. Economically removing the motor and extracting the magnets embedded in the rotor structure is not a trivial matter. There are some innovative technologies that can cost-effectively recover the REE compounds in magnets. Extraction of the magnets from EOL for reuse (as opposed to the recovery of the REE compounds) has not received enough attention because it is still labor-intensive and expensive.
Figure 20 is a schematic of the current automotive scrap shredding and separation methods. It includes processes such as magnetic and eddy current separation, flotation, electrostatic separation, etc. The complexity of the current process would limit the recovery of REE in the Nd-Fe-B content of the ferrous stream. The black box with a question mark in Figure 17 represents the lack of processes that can enable the recovery of the Nd-Fe-B component of the scrap ferrous materials’ stream for further processing to REE compounds.

5. Recycling of Permanent Magnets

Significant amounts of scrap magnets in magnet manufacturing plants are recycled as part of subsequent production steps. However, grinding swarf generated during the machining of permanent magnets are typically contaminated with cutting media and fluids, as well as being oxidized. Such materials are generally suitable for recycling approaches that enable the recovery of the constituent REEs, typically as REOs, that will be subsequently reinserted in the supply chain.
Since the permanent magnets would eventually be used in applications, they mostly reach the point of being recycled when the devices and systems that contain them reach their end-of-life. Thus, the magnets eventually become part of a wider collection of waste, commonly referred to as electronic waste, or e-waste. This presents some challenges, depending on how the different types of e-waste are decommissioned. It also presents opportunities to establish e-waste as a clean energy resource. These challenges and opportunities are further discussed below.

5.1. Challenges and Opportunities in Recycling of REEs from E-Waste

Mitigating the negative health and environmental impacts of the toxic chemical contents of e-waste due to improper disposal is yet another challenge. Challenge and opportunity can be considered two sides of the same coin. For example, the REEs obtained from mining are heavily diluted in concentration, whereas the materials coming from e-waste, such as HDDs and EVs, were previously refined, and hence may require fewer processing steps.
(i)
Collection
The challenges in collecting e-waste for value recovery remain among the greatest primary bottlenecks. For example, while an estimated 53.6 metric tons (Mt) of e-waste was produced globally (7.3 kg per capita) in 2019, only 9.3 Mt (~17%) of the e-waste was documented as properly recycled in the same year. That leaves nearly 44.3 Mt of undocumented e-waste that might have been land-filled. The land-filled e-waste was valued at USD 47.6B in 2019, which is expected to increase in the years to come. The global e-waste generation is projected to rise to 74.7 Mt in 2030 and to 110 Mt in 2050 (Figure 21) [58,59,60].
In addition to the loss of embodied value, if e-waste constituents are not properly recycled, they pose different levels of threats, owing but not limited to toxic elements such as lead, mercury, cadmium, manganese, nickel, and chromium. Other toxic organic compounds include, but are not limited to, polychlorinated bisphenols, flame retardants, perchlorates, thiocyanates, phthalates, bisphenols, and organophosphates [61]. Improper disposal of e-waste could also lead to emissions of 98 Mt of CO2 equivalent, 71 kt of brominated flame retardants, and 50 t of mercury emissions by 2050. Consequently, a proper e-waste collection mechanism is needed to ensure that there is little to no contamination of the environment. Most of the recycling needs to be better managed, as can be observed from the waste generated and the waste managed appropriately for the different continents (Figure 22) [59]. Another challenge is the availability of e-waste and the diversity in its distribution across various states in one country. For example, in 2019, the EV waste generated in California was nearly half of that of the other combined 50 states of the US. Of those, 24 states had less than 1% EV waste produced (Figure 23) [62].
The situation with waste HDDs is more organized across the various states of the US: 19 of the 50 states had less than 1% of HDD waste produced, while California produced a maximum share of 12.3% in 2019 [62] (Figure 24). However, the global demand in recent years has decreased due to solid-state drives (SSDs), as shown in Figure 25a. The HDD demand for the USA is computed based on the ratio of the world and USA GDPs and the global HDD demand in [63].
Two applications, namely, e-mobility and wind turbines, would have the highest recycling potentials, which show sharp growth after 2030 [58,59]. In line with achieving the zero-emission target in the transportation sector, EV sales have been rapidly increasing globally in recent years (Figure 25b) [64]. A demand forecast for EV units in the US shows a rapid increase: 4.72 million unit sales by 2030 [65]. Permanent magnet motors are the most widely used motors in transportation due to their high efficiency and power density. Therefore, in the future, high demand for REE magnets can be expected for traction motors. Additionally, EVs have a significant number of auxiliary motors along with the main traction motors, and some of these motors might be equipped with Nd-Fe-B magnets with various magnet masses. An average magnet mass of 175 g/vehicle was assumed to consider these auxiliary motors in recycling [66,67]. A significant amount of feedstock materials for recycling from EVs will become available in 10–12 years [68]. In addition, the infrastructure for collecting, transporting, and subsequent value recovery from the EOL EVs need to be established. Moreover, the processing, including the removal of magnets from the EOL EVs, is also a complicated procedure, as manual disassembly can be too expensive and a time-consuming process [69,70].
In the USA, only 10.3% of the total electric energy came from wind energy in 2022 (Figure 26a), most of which was from the onshore plants, which generally use non-PM generators [71]. However, offshore wind turbines with a power rating of more than 10 MW are considered cost-effective by improving installation cost per kW. Therefore, to reduce the weight and size of the offshore wind generators, rare-earth permanent magnets are used. Moreover, offshore plants also use the more reliable direct-drive technology to avoid induction generations with gearboxes that are prone to failure [72].
Currently, the USA has two operating plants: the Block Island Wind (50 MW) and the coastal Virginia offshore plant (12 MW). However, recently, the US Department of Energy reported an ambitious target to achieve 22 GW by 2030, followed by 86 GW by 2050 [73]. We project that the demand for REE magnets, especially Nd-Fe-B magnets, will increase up to 675 Mt/year until 2050 for this offshore plant expansion in the USA. This demand also includes magnets for expansion and replacement. Materials for recycling can be expected after 2034, as some of the early installed turbines will reach their EOL [74].
Figure 26. (a) Contribution from different sources for the USA electric generation in 2022. (b) Planned offshore plant generation capacity in the USA. Data adapted from [72,74].
Figure 26. (a) Contribution from different sources for the USA electric generation in 2022. (b) Planned offshore plant generation capacity in the USA. Data adapted from [72,74].
Sustainability 15 14901 g026
Another challenge is the huge disparity in the weight of magnets used in different applications. Barring four applications, namely, electric vehicles, conventional vehicles, and onshore and offshore wind turbines, there is less than 0.35 kg of magnet usage per appliance in most other applications (Figure 27). Hence, only 30% of the magnet weight comprises REEs, i.e., only 100 mg of REEs is used in most applications. The situation is aggravated further by the fact that the four applications that can yield the largest stockpile of waste magnets are also the ones that require a longer time (>20 years) prior to being available for REE recovery (Figure 28) [75].
Figure 29 demonstrates the visual depiction of the usage of magnets in various applications and the vast disparity in the usage. For example, magnets obtained from three HDDs are equivalent to that obtained from one refrigerator, and magnets obtained from four electric bikes are equivalent to that obtained from one electric vehicle. In contrast, magnets obtained from four hundred electric vehicles can make up for one magnet of the wind turbine, and ten such magnets from wind turbines can make up the magnets required in a magnetic resonance imaging (MRI) unit (Figure 29) [76].
There are a few big projects that have investigated pre-processing of e-waste for REE recovery. For example, the VALOMAG project, a strategy for recovering REE from HDDs, was developed in Europe. To make the process sustainable and economically feasible, the project adopted an automation strategy at every step (Figure 30) [77]. The United States Department of Energy’s Critical Materials Institute team of researchers from Oak Ridge National Laboratory developed a process via which magnets, circuit boards, and other valuable materials can be disassembled using automated technology. This process opens two possibilities: direct reuse of the magnets in applications and/or recovery of critical REE materials from the magnet [78].
(ii)
Pre-processing
Pre-processing refers to the important prerequisites before a permanent magnet is recovered from the e-waste for direct reuse, indirect reuse, or recovery. Different e-waste would require different pre-processing techniques because the size and mass of permanent magnets vary from each source. Additionally, the design of the systems housing the permanent magnets varies. For example, unlike hard disk drives, permanent magnets recycled from electric vehicles and wind turbines would require different approaches. While the former is preferably shredded, and hence typically requires further processing prior to REEs’ recovery, the latter may require disassembly to extract the permanent magnets, thereby enhancing the possibilities of direct reuse and reducing contamination of the magnets.
Li et al. suggest that the recycling of EVs in the future should advance from a one-stage approach to a three-stage approach to decrease the contamination of the valuable materials that can be obtained (Figure 31). In the one-stage approach, the EVs are sent directly to the production line to separate and shred valuable materials, followed by recovery. In the three-stage approach, they propose that the EVs would at first be disassembled into the components, and these would further be segregated into sub-assemblies, followed by shredding and recovery to decrease material contamination [79].

5.2. Recycling of Magnets—Terminologies

There are various terminologies and conventions used to define recycling processes. There still needs to be a clearer definition, as at present, it varies from one publication to the next. For example, Li et al. have referred to magnets cut into small pieces and used again in a different application as direct reuse [80]. Gandha et al. have cryogenically crushed AM-generated polymer-bonded magnet waste and reused the powder to produce bonded magnets. The authors state that the same process can be applied to EOL-bonded magnets and referred to this process as recycling [81]. It is worth noting that the actual magnetic property, which is restricted to the individual particle, remains intact even after cryogenic crushing. Although one might argue that the property might be much lower than sintered magnets, the crushed powder still possess hard magnetic properties typical of permanent magnets. Thus, the question that arises is, what constitutes a direct reuse and what should the converse be called? In the present report, we categorize the entire recycling into two major divisions, namely, direct recycling and extended recycling (Figure 32). Instead of referring to the converse of direct recycling as indirect recycling, we have chosen extended recycling to avoid confusion. The direct recycling would include minimal damage to the original form of the magnets, although the definition of “minimal” is still subjective. Extended recycling, on the other hand, would include processes that prolong the reinsertion of the magnet or values recovered from them into an application. These processes typically require additional changes to the original form of the magnets, i.e., crushing, grinding, dissolution (hydrometallurgical), melting (pyrometallurgical), etc. Direct and extended recycling will be further discussed in the subsequent sections.

5.3. Direct Recycling and Extended Recycling

Lansink’s ladder can be used to assess the life cycle of any material. It comprises five major steps: prevention, reuse, recycling, recovery, and disposal. The shape of the inverted pyramid can be used to comprehend the importance of each step in the end-of-life value of rare-earth materials. It can also be used to gauge the extent of environmental degradation, highest at the top, and lowest at the bottom. We can expand the hierarchy in terms of magnet recycling strategies. The first step of the ladder represents prevention (Figure 33). Harper et al. describes this as eliminating the use of critical elements [80]. The key approach for prevention is the development of substitutes with no or fewer critical materials for the same applications. Attempts have been reported on the reduced use of critical REE materials in PM motors and generators, which are used for traction/generation purposes [82,83,84,85,86]. Solid-state drives (SSDs) are now replacing HDDs and they do not have critical REE permanent magnets. Therefore, it can be expected that HDD sales will decline due to the new SSD technology [63]. Nevertheless, this trend may be limited by advancements in HDDs with the same speed as SSDs, since HDDs offer lower storage costs. Such an advancement was recently announced by Seagate [87]. Moreover, research is ongoing for high-temperature superconducting (HTS) generators for wind turbines [88]. Reuse is the next most viable option if environmental impacts are considered.
The direct recycling of REEs in permanent magnets can reduce the cost, time, and energy of reprocessing used magnets by dissolution, metallization, and magnet fabrication. Hogberg et al. have demonstrated that segmented poles can substitute a bread-loafed-based solid-pole magnet in a direct-drive wind turbine generator or electric vehicle motor [90]. It was demonstrated that using such segmented poles, referred to as the Lego design, instead of a single piece of magnet, a reduction in Eddy current losses was achieved, which helped in achieving comparable values of torque, cogging torque, and torque ripple [90]. The Lego design can efficiently aid in future usage of these segmented magnets, based on the shape and size requirements (Figure 34) [80]. The rotor is subjected to heating in a furnace to easily separate the demagnetized magnets from the rotor core, and a shield can be used to avoid any damage to the furnace or magnets. Magnet coatings with epoxy or metals, such as Cu, Ni, and Zn, also affect the magnetic properties in the thermal demagnetization process. Thus, this process can be used for extracting and reusing permanent magnets obtained from motors [91]. The process of segmenting the magnets will generate some waste that will need to be recycled.
One of the biggest challenges in the direct recycling of magnets is the ability to qualify the magnets as defect-free. Cui et al. have demonstrated the usage of ultrasonic testing to detect the possible presence of internal defects, such as microcracks, porosities, and inclusions [92]. This technique can predominantly aid manufacturers in better understanding the conditions that lead to defect-free samples. Additionally, this method can also help in the identification of end-of-life magnets that can be directly recycled to new magnets.
Magnets that fail to be directly recycled due to poor mechanical properties can be converted to powder and subjected to extended recycling. The usage of hydrogen as a tool to aid in the conversion of sintered magnets to a fine powder with a much finer grain size has gained momentum over the last two decades. The two methods include hydrogen decrepitation (HD) and hydrogenation disproportionation desorption recombination (HDDR) processes [93,94]. Reports suggest that the extended recycling of the fine powders obtained from magnets that were subjected to the HD and HDDR processes, followed by sintering, has helped achieve comparable results to sintered magnets prepared from virgin resources [95,96]. The fine powder obtained using HDDR processing can also be used for bonded magnets by combining with a polymer. The particle morphology, size, ratio of polymer to magnetic material, etc., play a role in the performance of bonded magnets [97]. Gandha et al. have demonstrated that the end-of-life-bonded magnets can be reprocessed using cryo-milling followed by isostatic pressing to achieve energy density values similar to those of starting additively manufactured bonded magnets (Figure 35) [81]. There still needs to be clarity about when a magnet can be sent for reprocessing by HD/HDDR and re-sintered or made into a bonded magnet and when it should be sent for the recovery of constituent REEs.

5.4. Recovery—A Subset of Extended Recycling

The suitability of scrap magnets for direct and extended recycling can be assessed based on their conditions. For example, if the scrap magnet is significantly oxidized and mechanically damaged, then the recycling can be performed through the recovery of the REEs. Moreover, the recovery process might be the best option when the magnets are not easy to disassemble, especially when they are destroyed together with the device that housed them in the application, e.g., shredded HDDs. Hydrometallurgical and pyrometallurgical processes are typically applied for the recovery of REEs from magnets.
(i)
Hydrometallurgical recycling
In the traditional hydrometallurgical recycling process, the EOL magnets are initially demagnetized and subsequently crushed into a fine powder. The powder is then typically oxidized and dissolved with a mineral acid. At this stage, a minimum amount of iron (which now exists as Fe2O3) is dissolved, and the undissolved iron oxide is separated by filtration. To the acidic leachate solution containing RE3+, oxalic acid is added, and the RE2(C2O4)3 residue is filtered and separated. The RE2(C2O4)3 residue is subsequently annealed to obtain RE2O3 [98]. A hydrometallurgical process was presented in which processing one metric ton of permanent magnet required 13.6 metric tons of water, 250 kg of HCl, and 257 kg of oxalic acid [99]. The process yields a product in excess of 300 kg of REE oxides based on the effectiveness of the recovery process. The electricity input required to process one metric ton of permanent magnets via an example of the traditional hydrometallurgical process is demonstrated in Figure 36. The oxidation process typically consumes the maximum electricity input of 369 kWh [99].
In addition to a large amount of acid usage (typically H2SO4, H2C2O4, and HF to precipitate the REEs as double sulfates oxalates, or fluorides) and a large amount of wastewater generation, hydrometallurgical processing typically requires many steps to recover all the REEs [100,101,102,103]. Furthermore, separating iron, nickel, and boron to obtain a high-purity RE salt remains challenging. In contrast to Nd-Fe-B magnets, recycling of Sm-Co magnets is simpler. This is because: (a) Sm-Co magnets typically contain only Gd as another REE, thereby decreasing the complexity of separation, and (b) Sm-Co magnets crystallize as 1:5 and 2:17 stoichiometries, where the latter typically contain Fe, Co, Cu, and Zr [104]. We remark that recycling feedstock materials that contain a mixture of Nd-Fe-B and Sm-Co magnets, such as swarf generated during magnet processing, presents elevated-level recycling difficulties. The difficulties owe to the fact that most hydrometallurgical processes co-recover the REEs as a mixture in a product, requiring further separation of the REEs for subsequent permanent magnets’ development.
Venkatesan et al. have demonstrated the electrolytic dissolution of the Nd-Fe-B magnet at room temperature using NH4Cl as the electrolyte. The Nd-Fe-B magnet was used as the anode and a copper rod was used as the cathode. Upon dissolution of the Nd-Fe-B, the Fe2+ was oxidized to Fe3+ at the inert electrode (Figure 37). The RE3+ was subsequently separated from a mixture of RE(OH)3 and Fe(OH)3 by using 0.14(M) HCl solution. This is a significant improvement owing to the diluted concentration of mineral acid compared to conventional hydrometallurgical processes [105].
There is a concerted effort to depart from the traditional hydrometallurgical routes that require corrosive mineral acids for magnet dissolution. Such efforts in recent research use inorganic salts for the dissolution. Maat et al. have demonstrated the dissolution of Nd from Nd-Fe-B magnets by hydrothermal reaction. The usage of NaCl improves the oxidation capabilities of water. By employing a high temperature (250 °C) and high pressure in an autoclave, atomic hydrogen produced in the process adsorbs on the surface of the magnets and produces a lattice expansion, and thereby, disintegration of the magnets occurs. Subsequently, the Nd converts to Nd(OH)3, and Fe3O4 is magnetically separated. Furthermore, they demonstrate that the process can also be carried out using other oxidative salts, such as Na2S2O4 and Na2S2O5 [106].
Prodius et al. have demonstrated the dissolution of Nd-Fe-B obtained from magnet swarf, decrepitated magnets, and shredded e-wastes using CuSO4 solution at room temperature (Figure 38). Additionally, Sm-Co magnets have also been shown to dissolve via the same procedure. In the presence of aqueous Cu2+ and aided by aerial oxidation, the RE3+ and Fe2+ are converted to RE(OH)3 and Fe(OH)3. The Boron is precipitated as copper borate [Cu3(BO3)2], and excess copper precipitates as Cu2O and metallic Cu. The RE3+ is extracted as rare-earth oxalate from the leachate. The unique feature of acid-free REE dissolution has brought forth several awards to the technology and the team and has reached pilot-scale commercialization [107,108,109,110,111].
In addition to the dissolution, the separation of rare earth also involves large amounts of chemicals. Furthermore, the process is time-consuming and involves the usage of large separation setups. Prodius et al. demonstrated that using organic salt (typically made by combining an organic base, such as alkyl-imidazolium or alkyl-pyrrolidinium, and an organic acid, such as oxalic acid) dissolved in water to produce neutral pH, the heavy rare-earth oxalates (Gd to Lu) can be separated from the light rare-earth oxalates (La to Sm) in fifteen minutes. The process can be used on feedstock derived from ores and recycling and can also be used to extract heavy rare-earth elements from dilute feedstock (~5%). This opens the possibility of the separation of REEs at neutral pH without the usage of any corrosive mineral acids [112,113,114].
(ii)
Pyrometallurgical recycling
Pyrometallurgical processes do not generate large amounts of contaminated wastewater; however, they require highly corrosive gases, use large amounts of energy, and generate large amounts of solid wastes [115]. In pyrometallurgical processes, liquid metal extraction [116,117,118,119] and electroslag refining [120,121] have been predominantly developed and studied [122].
Pyrometallurgical recycling is a very energy-intensive process. For example, Bian et al. state that the reduction step in recovering one metric ton of permanent magnets typically consumes 6688 kWh (Figure 39) and 0.15 kg of argon. The pyrometallurgical recycling of one metric ton of permanent magnets produces 615 kg of Fe and 455 kg of REOs [97].
In contrast to processes that eliminate toxic and volatile waste products, several reports have been provided on research directions that address this issue using extended processes. Bian et al. have proposed a combination of vacuum induction melting followed by hydrolysis and magnetic separation (VIM-HMS) (Figure 40). In the VIM process, the Nd-Fe-B ingots are mixed with carbon powder and uniformly melted at 1400 °C. Once the melt pool has cooled, they form rare-earth and iron carbides, which are subsequently crushed into a fine powder. The crushed magnet powder is transferred to a flask, and distilled water is added, giving rise to hydrocarbon products in addition to rare-earth hydroxide and Fe3O4. The latter product is magnetically separated, which gives rise to 99.7% RE(OH)3 [123].
Yue et al. have demonstrated the re-fabrication of Nd-Fe-B magnetic powder from magnetic sludge. The sludge is treated with acetone to initially remove the organic impurities. Subsequently, the cleaned and dried sludge material is treated with calcium granules at 1050–1150 °C for 3 h to directly convert the sludge into Nd-Fe-B. The separation of CaO was carried out by dissolution using dilute acetic acid [124].
Saito et al. have demonstrated the complete recovery of Nd from Nd-Fe-B magnets using the flux method. Upon mixing Nd-Fe-B magnets with B2O3 and melting the contents, it was demonstrated that complete conversion to Nd2O3 was possible. Although the process brings a mandatory separation step of Fe2B and α-Fe, the process is an alternate pyrometallurgical pathway [125].
Hua et al. propose the usage of a molten salt mixture of KCl and MgCl2 to extract Nd from Nd-Fe-B scrap granules. Upon heat-treating the contents, the Nd oxidizes to NdCl3, and the boron separates as a solid Fe-B residue. The molten salt mixture is subsequently subjected to electrolysis, which yields the RE-Mg alloy. The proposed technique is an improvement in comparison to both hydrometallurgical as well as pyrometallurgical techniques. The molten salt lowers the initial reaction temperature for the formation of chloride and the electrolysis process ensures that no extra heat needs to be supplied into the system for the separation of the RE alloy. With near complete usage of Fe and REEs, no emission of toxic and corrosive chemicals, and a lower energy consumption, the proposed technology can have an advantage over the existing hydrometallurgical techniques, as well. Another advantage of this process is that it yields REEs in metallic forms, in contrast to oxides that would still require a reduction step [126].

6. Commercialization Efforts

Some efforts are ongoing concerning the commercialization of RE recycling and production to address the criticality issue of REEs present around the globe. One of the earliest commercialization efforts was the partnership developed between Umicore and Rhodia to recycle nickel metal hydride batteries. The team combined the ultra-high-temperature battery recycling process of Umicore and the refining of REEs expertise of Rhodia. A hybrid vehicle NMH battery typically contains about 2 kg of RE, predominantly La, Ce, and Nd. The project aimed to recycle batteries at Umicore, followed by producing REE concentrates in Rhodia’s plant in La Rochelle, France [127].
In 2012, AERC Recycling Solutions partnered with Global Tungsten and Powders Corporation to recycle spent fluorescent lamps. They are predominantly rich in yttrium, cerium, lanthanum, and europium, and the project’s goal was to recover the REEs from those [128]. Future availability of REEs’ recovery from fluorescent lamps will be affected by the significant advancement in the use of LED lamps.
USA Rare Earth LLC teamed up with Texas Mineral Resources Corp. in late 2019 to jointly start a plant in Wheat Ridge, Colorado, to aid in separating and purification of REEs from ores obtained from El Paso, Texas. The plant was their first venture outside China with complete domestication and aiding the US critical materials supply chain from extraction to processing and distribution. The team looked at ion exchange chromatography to process REEs and targeted an annual production of nearly 2313 tons of REO per year [129]. In the spring of 2020, USA Rare Earth LLC purchased the Nd-Fe-B-producing hub in North Carolina from Hitachi Metals America, Ltd. It was proposed that the plant could produce nearly 2000 tons of REE magnets per year [130].
Solvay announced in 2022 that their La Rochelle plant in France would start the manufacturing of separated REOs and proposed to use it for the manufacturing of traction motors, wind turbine generator motors, and other electrical appliances. Traditionally, Solvay had investigated REE separation, purification, recycling, and formulation. The long-term goal of the project was to produce the largest REE hub in Europe [131].
The US Department of Defense awarded USD 35 million to MP Materials for mining and separation of REEs from USA’s largest REE reserves in Mountain Pass, California. This would propel the US toward domestic resilience on REEs and thereby decrease their dependency on imports. The REEs separated by MP Materials were planned to be utilized by General Motors to advance their EV ventures [132,133].
Energy Fuels announced the acquisition of seventeen mines in Bahia, Brazil, for extracting Nd, Pr, Dy, and Tb from monazite sands. The province is historically significant for mineral sands containing REE deposits. It was proposed that when fully functional, it can deliver anywhere from 3000 to 10,000 metric tons of monazite concentrate to Energy Fuels. The deal was formally announced in the first quarter of 2023 [134].
A few efforts are looking at recycling end-of-life products to extract critical REEs. For example, the acid-free dissolution recycling of permanent magnets obtained from HDDs and magnet swarm was awarded a patent in 2018 to researchers from the Ames National Laboratory. The technology went commercial with TdVib LLC in 2021 and has received State and Federal funding supports to scale-up REE recycling from e-wastes [110].
In 2022, Noveon Magnetics, a Texas-based US company, announced that they started producing new magnets from end-of-life magnets. According to the company, their process would be 90% more energy-efficient and could reduce the CO2 emissions by 11 metric tons per ton of magnet production, compared to the production of magnets from virgin sources. Recently, the US DOD awarded them USD 28.8 million for magnet production [135].

7. Future Outlooks

The rising demand for EVs and the subsequent mandates proposed by several countries to achieve net-zero emissions in a certain timeframe warrants the need for research on better recycling and mining strategies. As of November 2022, nearly 140 countries have announced timelines for achieving net-zero emissions. While some countries have already achieved net-zero emissions, several countries are working toward it [136,137].
The announcement by several automotive companies to stop producing gasoline-powered vehicles serves as a major boost to aiding the e-waste supply chain and could open opportunities for research and development on improved EVs, in addition to recycling of the e-wastes subsequently produced. To date, Mercedes [138], General Motors [139], Ford [140], Stellantis [141], Volvo [142], and BMW [143,144] have made announcements on moving away from fuel-based vehicles by 2030–2040.
Global EV sales have been on the rise exponentially over the years, with a major share of this contributed by China and the USA (Figure 41). Among these, the battery-operated electric vehicles (BEVs) are projected to take up a major share. It is estimated that from 9 million BEV sales in 2022, the number is projected to rise to 52 million by 2030 [145].
In the drive for clean energy, the role of wind turbines is going to be decisive in several countries around the world. Most wind turbines are currently based on double-fed induction generators (DFIGs), instead of the more efficient direct-drive permanent magnet generators (DDPMGs). Although the initial investment cost of the DFIG is low, the use of coils in its rotor, compared to permanent magnets in DDPMGs, exposes DFIGs to several disadvantages. It limits their ability to connect and support power grids. They can disconnect from the grid during under-voltage grid fault conditions, which can also take other DFIGs offline, hence leading to periods of inoperability. Moreover, the DFIG is connected to the grid using a back-to-back converter that can convert only 25–30% of the power produced by the DFIG, which is very inefficient in comparison to DDPMGs, capable of full power conversion. These are a few of the disadvantages of DFIGs, which are poised to exacerbate as the turbines age. When the total costs of investments and operations are considered, DFIGs may become costlier than DDPMGs, even without considering the costs due to the downtimes of DFIGs. The availability of REEs for permanent magnets is a key limitation to transitioning from DFIGs to DDPMGs, and this opens avenues of research opportunities.

Author Contributions

Conceptualization, J.O., I.C.N., A.K. and D.P.; methodology, I.C.N., A.K. and J.O.; writing—original draft preparation, J.O., A.K., A.P.S.B.; writing—review and editing, J.O., A.K., A.P.S.B., D.P. and I.C.N.; visualization, I.C.N., A.K. and J.O.; supervision, I.C.N.; project administration, I.C.N. and D.P.; funding acquisition, I.C.N. and D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

This work was supported by the Critical Materials Innovation Hub funded by the US Department of Energy, Office of Energy Efficiency and Renewable Energy, Advanced Materials and Manufacturing Technologies Office. The work was performed, in part, at Ames National Laboratory, operated for the US Department of Energy by Iowa State University of Science and Technology under Contract No. DE-AC02-07CH11358. Authors thank Marilú Pérez García for assistance with figure preparation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rare Earth Elements: Frequently Asked Questions. Available online: https://www.woodmac.com/news/editorial/rare-earth-elements-frequently-asked-questions/ (accessed on 6 July 2023).
  2. Securing America’s Clean Energy Supply Chain. Available online: https://www.energy.gov/policy/securing-americas-clean-energy-supply-chain (accessed on 6 July 2023).
  3. Critical Materials for The Energy Transition: Rare Earth Elements. Available online: https://www.irena.org/Technical-Papers/Critical-Materials-For-The-Energy-Transition-Rare-Earth-elements (accessed on 6 July 2023).
  4. Binnemans, K.; Jones, P.T.; Müller, T.; Yurramendi, L. Rare Earths and the Balance Problem: How to Deal with Changing Markets? J. Sustain. Metall. 2018, 4, 126–146. [Google Scholar] [CrossRef]
  5. Rare Earth Elements. Available online: https://web.mit.edu/12.000/www/m2016/finalwebsite/elements/ree.html (accessed on 23 July 2023).
  6. Balaram, V. Rare Earth Elements: A Review of Applications, Occurrence, Exploration, Analysis, Recycling, and Environmental Impact. Geosci. Front. 2019, 10, 1285–1303. [Google Scholar] [CrossRef]
  7. Dostal, J. Rare Earth Element Deposits of Alkaline Igneous Rocks. Resources 2017, 6, 34. [Google Scholar] [CrossRef]
  8. Reck, B.K.; Graedel, T.E. Challenges in Metal Recycling. Science 2012, 337, 690–695. [Google Scholar] [CrossRef]
  9. Du, X.; Graedel, T.E. Uncovering the Global Life Cycles of the Rare Earth Elements. Sci. Rep. 2011, 1, 145. [Google Scholar] [CrossRef]
  10. Balaram, V. Potential Future Alternative Resources for Rare Earth Elements: Opportunities and Challenges. Minerals 2023, 13, 425. [Google Scholar] [CrossRef]
  11. Gaustad, G.; Williams, E.; Leader, A. Rare Earth Metals from Secondary Sources: Review of Potential Supply from Waste and Byproducts. Resour. Conserv. Recycl. 2021, 167, 105213. [Google Scholar] [CrossRef]
  12. Peiravi, M.; Dehghani, F.; Ackah, L.; Baharlouei, A.; Godbold, J.; Liu, J.; Mohanty, M.; Ghosh, T. A Review of Rare-Earth Elements Extraction with Emphasis on Non-Conventional Sources: Coal and Coal Byproducts, Iron Ore Tailings, Apatite, and Phosphate Byproducts. Min. Metall. Explor. 2021, 38, 1–26. [Google Scholar] [CrossRef]
  13. Mineral Commodity Summaries 2023. Available online: https://pubs.usgs.gov/publication/mcs2023 (accessed on 8 September 2023).
  14. Javed, A.; Singh, J. Process Intensification for Sustainable Extraction of Metals from E-Waste: Challenges and Opportunities. Environ. Sci. Pollut. Res. 2023, 1–34. [Google Scholar] [CrossRef]
  15. Zeng, X.; Mathews, J.A.; Li, J. Urban Mining of E-Waste Is Becoming More Cost-Effective Than Virgin Mining. Environ. Sci. Technol. 2018, 52, 4835–4841. [Google Scholar] [CrossRef]
  16. Illinois Basin Carbon Ore, Rare Earth, and Critical Minerals Initiative. Available online: https://netl.doe.gov/node/11905 (accessed on 6 July 2023).
  17. Critical Minerals in Coaly Strata of the Cherokee-Forest City Basin. Available online: https://netl.doe.gov/node/11894 (accessed on 6 July 2023).
  18. Manufacturing Valuable Coal-Derived Products in Southern Appalachia. Available online: https://netl.doe.gov/node/11892 (accessed on 6 July 2023).
  19. Assessment of Rare Earth Elements and Critical Minerals in Coal and Coal Ash in the U.S. Gulf Coast. Available online: https://netl.doe.gov/node/11895 (accessed on 6 July 2023).
  20. Bringing Alaska’s CORE-CM Potential into Perspective. Available online: https://netl.doe.gov/node/11890 (accessed on 6 July 2023).
  21. Clegg, S.M.; Alamos, L. Evaluation of Laser-Based Analysis of Rare Earth Elements in Coal-Related Materials. Available online: https://netl.doe.gov/sites/default/files/netl-file/20VPRREE_Clegg.pdf (accessed on 7 July 2023).
  22. Ziemkiewicz, P.; Noble, A.; Tech, V.; Quaranta, J.; Lian, W.; Lin, S.; Finklea, H. Development and Testing of an Integrated AMD/REE-CM Plant. Available online: https://netl.doe.gov/sites/default/files/netl-file/20VPRREE_Ziemkiewicz.pdf (accessed on 6 July 2023).
  23. Zhang, P. Technology Development, and Integration for Volume Production of High Purity Rare Earth Metals from Phosphate Processing DE-FE0032123. Available online: https://netl.doe.gov/sites/default/files/netl-file/22RS-26_Zhang.pdf (accessed on 7 July 2023).
  24. Ayora, C.; Macías, F.; Torres, E.; Lozano, A.; Carrero, S.; Nieto, J.M.; Pérez-López, R.; Fernández-Martínez, A.; Castillo-Michel, H. Recovery of Rare Earth Elements and Yttrium from Passive-Remediation Systems of Acid Mine Drainage. Environ. Sci. Technol. 2016, 50, 8255–8262. [Google Scholar] [CrossRef] [PubMed]
  25. Merten, D.; Geletneky, J.; Bergmann, H.; Haferburg, G.; Kothe, E.; Büchel, G. Rare Earth Element Patterns: A Tool for Understanding Processes in Remediation of Acid Mine Drainage. Geochemistry 2005, 65, 97–114. [Google Scholar] [CrossRef]
  26. Ferreira da Silva, E.; Bobos, I.; Xavier Matos, J.; Patinha, C.; Reis, A.P.; Cardoso Fonseca, E. Mineralogy and Geochemistry of Trace Metals and REE in Volcanic Massive Sulfide Host Rocks, Stream Sediments, Stream Waters and Acid Mine Drainage from the Lousal Mine Area (Iberian Pyrite Belt, Portugal). J. Appl. Geochem. 2009, 24, 383–401. [Google Scholar] [CrossRef]
  27. Sahoo, P.K.; Tripathy, S.; Equeenuddin, S.M.; Panigrahi, M.K. Geochemical Characteristics of Coal Mine Discharge Vis-à-Vis Behavior of Rare Earth Elements at Jaintia Hills Coalfield, Northeastern India. J. Geochem. Explor. 2012, 112, 235–243. [Google Scholar] [CrossRef]
  28. León, R.; Macías, F.; Cánovas, C.R.; Millán-Becerro, R.; Pérez-López, R.; Ayora, C.; Nieto, J.M. Evidence of Rare Earth Elements Origin in Acid Mine Drainage from the Iberian Pyrite Belt (SW Spain). Ore Geol. Rev. 2023, 154, 105336. [Google Scholar] [CrossRef]
  29. Rare Earth Oxide Demand Worldwide 2025. Available online: https://www.statista.com/statistics/1114638/global-rare-earth-oxide-demand/ (accessed on 6 September 2023).
  30. Rare Earth Magnet Market Outlook to 2035—Adamas Intelligence. Available online: https://www.adamasintel.com/rare-earth-magnet-market-outlook-to-2035/ (accessed on 6 July 2023).
  31. Global Permanent Magnet Industry Report. Available online: https://www.magnetreport.com/ (accessed on 6 July 2023).
  32. Critical Minerals Market Review 2023. Available online: https://www.iea.org/reports/critical-minerals-market-review-2023 (accessed on 6 September 2023).
  33. Metal Production by Salt Electrolysis. Available online: https://www.eurare.org/technologies/REE-metal-production.html (accessed on 6 July 2023).
  34. World Energy Transitions Outlook 2022. Available online: https://www.irena.org/Digital-Report/World-Energy-Transitions-Outlook-2022 (accessed on 6 July 2023).
  35. Croat, J.J.; Ormerod, J.G. Modern Permanent Magnets; Woodhead Publishing: Sawston, UK, 2022; ISBN 9780323886581/9780323886406. [Google Scholar]
  36. Rare Earth Magnets: Yesterday, Today and Tomorrow. Available online: https://www.slideshare.net/JohnOrmerod/2019-01-17-magnetics-2019 (accessed on 6 July 2023).
  37. Dushyantha, N.; Batapola, N.; Ilankoon, I.M.S.K.; Rohitha, S.; Premasiri, R.; Abeysinghe, B.; Ratnayake, N.; Dissanayake, K. The Story of Rare Earth Elements (REEs): Occurrences, Global Distribution, Genesis, Geology, Mineralogy and Global Production. Ore Geol. Rev. 2020, 122, 103521. [Google Scholar] [CrossRef]
  38. Nesbitt, E.A.; Wernick, J.H. Rare Earth Permanent Magnets; Academic Press: New York, NY, USA, 1973. [Google Scholar]
  39. Strnat, K.J.; Strnat, R.M.W. Rare Earth-Cobalt Permanent Magnets. J. Magn. Magn. Mater. 1991, 100, 38–56. [Google Scholar] [CrossRef]
  40. Menth, A.; Nagel, H.; Perkins, R.S. New high-performance permanent magnets based on rare earth-transition metal compounds. Annu. Rev. Mater. Res. 1978, 8, 21–47. [Google Scholar] [CrossRef]
  41. Livingston, J.D. Chalmers Anniversary Volume, 1st ed.; Christian, J.W., Haasen, P., Massalski, T.B., Eds.; Pergamon Press: Oxford, UK, 1981; pp. 243–268. [Google Scholar]
  42. Ormerod, J. The Physical Metallurgy and Processing of Sintered Rare Earth Permanent Magnets. J. Less-Common Met. 1985, 111, 49–69. [Google Scholar] [CrossRef]
  43. Strnat, K.J. Modern permanent magnets for applications in electro-technology. Proc. IEEE 1990, 78, 923–946. [Google Scholar] [CrossRef]
  44. Rare Earth Magnet Market Outlook to 2040. Available online: https://www.adamasintel.com/rare-earth-magnet-market-outlook-to-2040/ (accessed on 6 September 2023).
  45. Smith, B.J.; Riddle, M.E.; Earlam, M.R.; Iloeje, C.; Diamond, D. Rare Earth Permanent Magnets: Supply Chain Deep Dive Assessment; USDOE Office of Policy: Washington, DC, USA, 2022. [Google Scholar] [CrossRef]
  46. Cullity, B.D. Introduction to Magnetic Materials; Addison-Wesley Pub. Co.: Reading, MA, USA, 1972. [Google Scholar]
  47. Additive Manufacturing of Permanent Magnets. Available online: https://www.sigmaaldrich.com/US/en/technical-documents/technical-article/materials-science-and-engineering/nanoparticle-and-microparticle-synthesis/additive-manufacturing-of-permanent-magnets (accessed on 6 September 2023).
  48. Constantinides, S. The elements of magnetics. MRS Online Proc. Libr. Arch. 2013, 1492, 54–65. [Google Scholar] [CrossRef]
  49. Types of Rare Earth Magnets: Part 1. Available online: https://bunting-dubois.com/tech-briefs/types-of-rare-earth-magnets-part-1/ (accessed on 6 September 2023).
  50. Cui, J.; Ormerod, J.; Parker, D.; Ott, R.; Palasyuk, A.; Mccall, S.; Paranthaman, M.P.; Kesler, M.S.; McGuire, M.A.; Nlebedim, I.C.; et al. Manufacturing Processes for Permanent Magnets: Part I—Sintering and Casting. JOM 2022, 74, 1279–1295. [Google Scholar] [CrossRef]
  51. Croat, J.J. Rapidly Solidified Neodymium-Iron-Boron Permanent Magnets. In Rapidly Solidified Neodymium-Iron-Boron Permanent Magnets; Woodhead Publishing: Sawston, UK, 2017; pp. 1–374. [Google Scholar] [CrossRef]
  52. How to Assure Every Rare Earth (or Any) Magnet Meets Your Requirements 100 Percent of the Time? Available online: https://bunting-dubois.com/news/september-21-2016-how-to-assure-every-rare-earth-or-any-magnet-meets-your-requirements-100-percent-of-the-time/ (accessed on 6 September 2023).
  53. Pan, M.; Zhang, P.; Li, X.; Ge, H.; Wu, Q.; Jiao, Z.; Liu, T. Effect of Terbium Addition on the Coercivity of the Sintered Nd-Fe-B Magnets. J. Rare Earths 2010, 28, 399–402. [Google Scholar] [CrossRef]
  54. Bunting DuBois. Available online: https://bunting-dubois.com/ (accessed on 23 July 2023).
  55. Constantinides, S. The Demand for Rare Earth Materials in Permanent Magnets. In Proceedings of the 51st Annual Conference of Metallurgists COM Niagara Falls, Niagara Falls, NY, USA, 30 September–3 October 2012. [Google Scholar]
  56. Tech Library. Available online: https://product.tdk.com/en/techlibrary/index.html (accessed on 6 July 2023).
  57. BMW X5 Plug-In Hybrid Prototype: We Drive Future Electric SUV. Available online: https://www.greencarreports.com/news/1091471_bmw-x5-plug-in-hybrid-prototype-we-drive-future-electric-suv (accessed on 6 July 2023).
  58. Parajuly, K.; Kuehr, R.; Awasthi, A.K.; Fitzpatrick, C.; Lepawsky, J.; Smith, E.; Widmer, R.; Zeng, X. Future E-Waste Scenarios. Available online: https://www.unep.org/ietc/resources/publication/future-e-waste-scenarios (accessed on 7 July 2023).
  59. Adrian, S.; Drisse, M.B.; Cheng, Y.; Devia, L.; Deubzer, O.; Goldizen, F.; Gorman, J.; Herat, S.; Honda, S.; Iattoni, G.; et al. Quantities, Flows, and the Circular Economy Potential. The Global E-Waste Monitor. 2020. Available online: https://ewastemonitor.info/wp-content/uploads/2020/11/GEM_2020_def_july1_low.pdf (accessed on 7 July 2023).
  60. Baldé, C.P.; Angelo, E.D.; Luda, V.; Deubzer, O.; Kuehr, R. Global Transboundary E-Waste Flows Monitor 2022. Available online: https://ewastemonitor.info/wp-content/uploads/2022/06/Global-TBM_webversion_june_2_pages.pdf (accessed on 7 July 2023).
  61. Gorman, J.F.; Parvez, S.M.; Jahan, F.; Brune, M.-N.; Gorman, J.F.; Rahman, J.; Carpenter, D.; Islam, Z.; Rahman, M. Health Consequences of Exposure to E-Waste: An Updated Systematic Review. Lancet Glob. Health 2021, 5, 905–925. [Google Scholar] [CrossRef]
  62. Jin, H.; Song, B.D.; Yih, Y.; Sutherland, J.W. Sustainable Value Recovery of Nd-Fe-B Magnets: A Multi-Objective Network Design and Genetic Algorithm. ACS Sustain. Chem. Eng. 2018, 6, 4767–4775. [Google Scholar] [CrossRef]
  63. Global Hard Disk Drive (HDD) Shipments 1976–2022. Available online: https://www.statista.com/statistics/398951/global-shipment-figures-for-hard-disk-drives/ (accessed on 7 July 2023).
  64. Global Sales and Sales Market Share of Electric Cars, 2010–2021. Available online: https://www.iea.org/data-and-statistics/charts/global-sales-and-sales-market-share-of-electric-cars-2010-2021 (accessed on 7 July 2023).
  65. EV Sales Forecasts. Available online: https://evadoption.com/ev-sales/ev-sales-forecasts/ (accessed on 7 July 2023).
  66. Akatsu, K.; Matsui, N. New Trend of Motor Technology for Automobiles—Introduction and Overview. In Proceedings of the 5th IEEE Annual International Energy Conversion Congress and Exhibition (ECCE) Asia DownUnder Conference, Melbourne, Australia, 3–6 June 2013; pp. 130–135. [Google Scholar] [CrossRef]
  67. Reimer, M.V.; Schenk-Mathes, H.Y.; Hoffmann, M.F.; Elwert, T. Recycling Decisions in 2020, 2030, and 2040—When Can Substantial Nd-Fe-B Extraction Be Expected in the EU? Metals 2018, 8, 867. [Google Scholar] [CrossRef]
  68. Maani, T.; Mathur, N.; Singh, S.; Rong, C.; Sutherland, J.W. Potential for Nd, and Dy Recovery from End-of-Life Products to Meet Future Electric Vehicle Demand in the U.S. Procedia CIRP 2021, 98, 109–114. [Google Scholar] [CrossRef]
  69. Klier, T.; Risch, F.; Franke, J. Disassembly, Recycling, and Reuse of Magnet Material of Electric Drives. In Proceedings of the 2013 IEEE International Symposium on Assembly and Manufacturing (ISAM), Xi’an, China, 30 July–2 August 2013; pp. 88–90. [Google Scholar] [CrossRef]
  70. Elwert, T.; Goldmann, D.; Roemer, F.; Schwarz, S. Recycling of Nd-Fe-B Magnets from Electric Drive Motors of (Hybrid) Electric Vehicles. J. Sus. Meta. 2017, 3, 108–121. [Google Scholar] [CrossRef]
  71. Electricity Data Browser—Net Generation for All Sectors. Available online: https://www.eia.gov/electricity/data/browser/ (accessed on 7 July 2023).
  72. Abrahamsen, A.B.; Magnusson, N.; Jensen, B.B.; Runde, M. Large Superconducting Wind Turbine Generators. Energy Procedia 2012, 24, 60–67. [Google Scholar] [CrossRef]
  73. Wind Vision. Available online: https://www.energy.gov/eere/wind/wind-vision-1 (accessed on 7 July 2023).
  74. Fishman, T.; Graedel, T.E. Impact of the Establishment of US Offshore Wind Power on Neodymium Flows. Nat. Sustain. 2019, 2, 332–338. [Google Scholar] [CrossRef]
  75. Rizos, V.; Righetti, E.; Kassab, A. Developing a Supply Chain for Recycled Rare Earth Permanent Magnets in the EU Ceps In-Depth Analysis Challenges and Opportunities. Available online: https://www.ceps.eu/wp-content/uploads/2022/12/CEPS-In-depth-analysis-2022-07_Supply-chain-for-recycled-rare-earth-permanent-magnets.pdf (accessed on 7 July 2023).
  76. Kumari, A.; Kumar Sahu, S. A Comprehensive Review on Recycling of Critical Raw Materials from Spent Neodymium Iron Boron (Nd-Fe-B) Magnet. Sep. Purif. Technol. 2023, 317, 123527. [Google Scholar] [CrossRef]
  77. Coelho, F.; Abrahami, S.; Yang, Y.; Sprecher, B.; Li, Z.; Menad, N.-E.; Bru, K.; Marcon, T.; Rado, C.; Saje, B.; et al. Upscaling of Permanent Magnet Dismantling and Recycling through VALOMAG Project. Mater. Proc. 2021, 5, 74. [Google Scholar] [CrossRef]
  78. ORNL Licenses Rare Earth Magnet Recycling Process to Momentum Technologies. Available online: https://www.ornl.gov/news/ornl-licenses-rare-earth-magnet-recycling-process-momentum-technologies (accessed on 7 July 2023).
  79. Li, J.; Barwood, M.; Rahimifard, S. Robotic Disassembly for Increased Recovery of Strategically Important Materials from Electrical Vehicles. Robot Comput. Integr. Manuf. 2018, 50, 203–212. [Google Scholar] [CrossRef]
  80. Li, Z.; Kedous-Lebouc, A.; Dubus, J.M.; Garbuio, L.; Personnaz, S. Direct Reuse Strategies of Rare Earth Permanent Magnets for PM Electrical Machines—An Overview Study. EPJ Appl. Phys. 2019, 86, 20901. [Google Scholar] [CrossRef]
  81. Gandha, K.; Ouyang, G.; Gupta, S.; Kunc, V.; Paranthaman, M.P.; Nlebedim, I.C. Recycling of Additively Printed Rare-Earth Bonded Magnets. Waste Manag. 2019, 90, 94–99. [Google Scholar] [CrossRef]
  82. Pavel, C.C.; Thiel, C.; Degreif, S.; Blagoeva, D.; Buchert, M.; Schüler, D.; Tzimas, E. Role of Substitution in Mitigating the Supply Pressure of Rare Earths in Electric Road Transport Applications. Sustain. Mater. Technol. 2017, 12, 62–72. [Google Scholar] [CrossRef]
  83. Pavel, C.C.; Lacal-Arántegui, R.; Marmier, A.; Schüler, D.; Tzimas, E.; Buchert, M.; Jenseit, W.; Blagoeva, D. Substitution Strategies for Reducing the Use of Rare Earths in Wind Turbines. Res. Pol. 2017, 52, 349–357. [Google Scholar] [CrossRef]
  84. Baghel, A.P.S.; Nlebedim, I.C. A Hybrid Rotor Design with Reduced Rare Earth Magnets for Traction Motors. JOM 2023, 75, 557–565. [Google Scholar] [CrossRef]
  85. El-Refaie, A.; Raminosoa, T.; Reddy, P.; Galioto, S.; Pan, D.; Grace, K.; Alexander, J.; Huh, K.K. Comparison of Traction Motors That Reduce or Eliminate Rare-Earth Materials. In Proceedings of the ECCE 2016—IEEE Energy Conversion Congress and Exposition, Milwaukee, WI, USA, 18–22 September 2016; pp. 1–88. [Google Scholar] [CrossRef]
  86. Jeong, C.L.; Hur, J. A Novel Proposal to Improve Reliability of Spoke-Type BLDC Motor Using Ferrite Permanent Magnet. IEEE Trans. Ind. Appl. 2016, 52, 3814–3821. [Google Scholar] [CrossRef]
  87. Seagate Introduces HDDs as Fast as SSDs. Available online: https://www.networkworld.com/article/3681656/seagate-introduces-hdds-as-fast-as-ssds.html (accessed on 7 July 2023).
  88. Kolchanova, I.; Poltavets, V. Superconducting Generators for Wind Turbines. In Proceedings of the International Conference on Electrotechnical Complexes and Systems, Ufa, Russia, 26–29 October 2021; pp. 529–533. [Google Scholar]
  89. Harper, G.; Sommerville, R.; Kendrick, E.; Driscoll, L.; Slater, P.; Stolkin, R.; Walton, A.; Christensen, P.; Heidrich, O.; Lambert, S.; et al. Recycling Lithium-Ion Batteries from Electric Vehicles. Nature 2019, 575, 75–86. [Google Scholar] [CrossRef]
  90. Hogberg, S.; Pedersen, T.S.; Bendixen, F.B.; Mijatovic, N.; Jensen, B.B.; Holboll, J. Direct Reuse of Rare Earth Permanent Magnets—Wind Turbine Generator Case Study. In Proceedings of the 2016 22nd International Conference on Electrical Machines, ICEM 2016, Lausanne, Switzerland, 4–7 September 2016; pp. 1625–1629. [Google Scholar] [CrossRef]
  91. Hogberg, S.; Holboll, J.; Mijatovic, N.; Jensen, B.B.; Bendixen, F.B. Direct Reuse of Rare Earth Permanent Magnets—Coating Integrity. IEEE Trans. Magn. 2017, 53, 1–9. [Google Scholar] [CrossRef]
  92. Cui, B.; Cui, J.; Barnard, D.J.; Bond, L.J. Internal Defect Detection and Characterization of Samarium-Cobalt Sintered Magnets by Ultrasonic Testing Technique. J. Magn. Magn. Mater. 2023, 570, 170524. [Google Scholar] [CrossRef]
  93. Liu, X.B.; Kesler, M.S.; Besser, M.F.; Kramer, M.J.; McGuire, M.A.; Nlebedim, I.C. Effect of Processing Hydrogen Pressure on Magnetic Properties of HDDR Nd-Fe-B Magnet. IEEE Trans. Magn. 2021, 57, 2100604. [Google Scholar] [CrossRef]
  94. Sheridan, R.S.; Sillitoe, R.; Zakotnik, M.; Harris, I.R.; Williams, A.J. Anisotropic Powder from Sintered Nd-Fe-B Magnets by the HDDR Processing Route. J. Magn. Magn. Mater. 2012, 324, 63–67. [Google Scholar] [CrossRef]
  95. Zakotnik, M.; Harris, I.R.; Williams, A.J. Multiple Recycling of Nd-Fe-B-Type Sintered Magnets. J. Alloys Compd. 2009, 469, 314–321. [Google Scholar] [CrossRef]
  96. Harris, I.R.; Williams, A.; Walton, A.; Speight, J. Magnet Recycling. U.S. Patent 8734714B2, 27 May 2011. [Google Scholar]
  97. Cui, J.; Ormerod, J.; Parker, D.S.; Ott, R.; Palasyuk, A.; McCall, S.; Paranthaman, M.P.; Kesler, M.S.; McGuire, M.A.; Nlebedim, C.; et al. Manufacturing Processes for Permanent Magnets: Part II—Bonding and Emerging Methods. JOM 2022, 74, 2492–2506. [Google Scholar] [CrossRef]
  98. Kumari, A.; Sinha, M.K.; Pramanik, S.; Sahu, S.K. Recovery of Rare Earths from Spent Nd-Fe-B Magnets of Wind Turbine: Leaching and Kinetic Aspects. Waste Manag. 2018, 75, 486–498. [Google Scholar] [CrossRef]
  99. Becci, A.; Beolchini, F.; Amato, A.; Ippolito, M.; Innocenzi, V. Sustainable Strategies for the Exploitation of End-of-Life Permanent Magnets. Processes 2021, 9, 857. [Google Scholar] [CrossRef]
  100. Yang, Y.; Lan, C.; Wang, Y.; Zhao, Z.; Li, B. Recycling of Ultrafine Nd-Fe-B Waste by the Selective Precipitation of Rare Earth, and the Electrodeposition of Iron in Hydrofluoric Acid. Sep. Purif. Technol. 2020, 230, 115870. [Google Scholar] [CrossRef]
  101. Vander Hoogerstraete, T.; Blanpain, B.; Van Gerven, T.; Binnemans, K. From Nd-Fe-B Magnets towards the Rare-Earth Oxides: A Recycling Process Consuming Only Oxalic Acid. RSC Adv. 2014, 4, 64099–64111. [Google Scholar] [CrossRef]
  102. Abrahami, S.T.; Xiao, Y.; Yang, Y. Rare-Earth Elements Recovery from Post-Consumer Hard-Disc Drives. Miner. Process. Extr. Metall. Trans. Inst. Min. Metall. 2014, 124, 106–115. [Google Scholar] [CrossRef]
  103. Liu, F.; Porvali, A.; Wang, J.L.; Wang, H.; Peng, C.; Wilson, B.P.; Lundström, M. Recovery and Separation of Rare Earths and Boron from Spent Nd-Fe-B Magnets. Miner. Eng. 2020, 145, 106097. [Google Scholar] [CrossRef]
  104. Binnemans, K.; Jones, P.T.; Blanpain, B.; Van Gerven, T.; Yang, Y.; Walton, A.; Buchert, M. Recycling of Rare Earths: A Critical Review. J. Clean Prod. 2013, 51, 1–22. [Google Scholar] [CrossRef]
  105. Venkatesan, P.; Vander Hoogerstraete, T.; Binnemans, K.; Sun, Z.; Sietsma, J.; Yang, Y. Selective Extraction of Rare-Earth Elements from Nd-Fe-B Magnets by a Room temperature Electrolysis Pretreatment Step. ACS Sustain. Chem. Eng. 2018, 6, 9375–9382. [Google Scholar] [CrossRef]
  106. Maât, N.; Nachbaur, V.; Lardé, R.; Juraszek, J.; Le Breton, J.M. An Innovative Process Using Only Water and Sodium Chloride for Recovering Rare Earth Elements from Nd-Fe-B Permanent Magnets Found in the Waste of Electrical and Electronic Equipment. ACS Sustain. Chem. Eng. 2016, 4, 6455–6462. [Google Scholar] [CrossRef]
  107. Mudring, A.-V.; Prodius, D.; Nlebedim, C.I. Dissolution and Separation of Rare Earth Metals. U.S. Patent 10648063B2, 12 May 2018. [Google Scholar]
  108. Green Rare-Earth Recycling Goes Commercial in the US. Available online: https://www.ameslab.gov/news/green-rare-earth-recycling-goes-commercial-in-the-us (accessed on 7 July 2023).
  109. Rare-Earth Magnet Recycling Wins FLC Innovation Award. Available online: https://www.ameslab.gov/news/rare-earth-magnet-recycling-wins-flc-innovation-award (accessed on 7 July 2023).
  110. Acid Free Dissolution Recycling: A Second Chance for Rare Earth Elements. Available online: https://www.energy.gov/eere/ammto/articles/acid-free-dissolution-recycling-second-chance-rare-earth-elements (accessed on 7 July 2023).
  111. Prodius, D.; Gandha, K.; Mudring, A.V.; Nlebedim, I.C. Sustainable Urban Mining of Critical Elements from Magnet and Electronic Wastes. ACS Sustain. Chem. Eng. 2020, 8, 1455–1463. [Google Scholar] [CrossRef]
  112. Prodius, D.; Klocke, M.; Smetana, V.; Alammar, T.; Perez Garcia, M.; Windus, T.L.; Nlebedim, I.C.; Mudring, A.V. Rationally Designed Rare Earth Separation by Selective Oxalate Solubilization. Chem. Comm. 2020, 56, 11386–11389. [Google Scholar] [CrossRef] [PubMed]
  113. Prodius, D.; Nlebedim, C.A.; Mudring, A.-V. Separating Rare Earth Metal Oxalates. U.S. Patent 11090579B2, 17 August 2021. [Google Scholar]
  114. Rationally Designed Rare Earth Separation by Selective Oxalate Solubilization. Available online: https://www.ameslab.gov/cmi/research-highlights/rationally-designed-rare-earth-separation-by-selective-oxalate (accessed on 7 July 2023).
  115. Laputka, M.; Xie, W. A review of recent advances in pyrometallurgical process measurement and modelling and their applications to process improvement. Mining Metall. Explor. 2021, 38, 1135–1165. [Google Scholar] [CrossRef]
  116. Ellis, T.W.; Schmidt, F.A. Recycling of Rare Earth Metals from Rare Earth-Transition Metal Alloy Scrap by Liquid Metal Extraction. U.S. Patent 5437709A, 1 August 1995. [Google Scholar]
  117. Ott, R.T.; McCallum, R.W. Recovering Rare Earth Metals from Magnet Scrap. U.S. Patent 10323299B2, 18 June 2019. [Google Scholar]
  118. Ott, R.T.; McCallum, R.W.; Jones, L.L. Recovering Heavy Rare Earth Metals from Magnet Scrap. U.S. Patent 9725788B2, 28 January 2016. [Google Scholar]
  119. Xu, Y.; Chumbley, L.S.; Laabs, F.C. Liquid metal extraction of Nd from NdFeB magnet scrap. J. Mater. Res. 2000, 15, 2296–2304. [Google Scholar] [CrossRef]
  120. Schmidt, F.; Peterson, D.; Wheelock, J.; Jones, L. Method for Treating Rare Earth-Transition Metal Scrap. U.S. Patent 5174811A, 29 December 1992. [Google Scholar]
  121. Ellis, T.W.; Schmidt, F.A.; Jones, L.L. Methods and opportunities in the recycling of rare earth based materials. In Symposium on Metals and Materials Waste Reduction: Recovery and Remediation; Liddell, K.C., Bautista, R.G., Orth, R.J., Eds.; The Minerals, Metals & Materials Society: Warrendale, PA, USA, 1994; pp. 199–206. [Google Scholar]
  122. Nlebedim, I.C.; King, A.H. Addressing Criticality in Rare Earth Elements via Permanent Magnets Recycling. JOM 2018, 70, 115–123. [Google Scholar] [CrossRef]
  123. Bian, Y.; Guo, S.; Jiang, L.; Liu, J.; Tang, K.; Ding, W. Recovery of Rare Earth Elements from Nd-Fe-B Magnet by VIM-HMS Method. ACS Sustain. Chem. Eng. 2016, 4, 810–818. [Google Scholar] [CrossRef]
  124. Yue, M.; Yin, X.; Li, X.; Li, M.; Li, X.; Liu, W.; Wu, Y.; Zhang, D.; Chen, J.; Yi, X.; et al. Recycling of Nd-Fe-B Sintered Magnets Sludge via the Reduction-Diffusion Route to Produce Sintered Magnets with Strong Energy Density. ACS Sustain. Chem. Eng. 2018, 6, 6547–6553. [Google Scholar] [CrossRef]
  125. Saito, T.; Sato, H.; Ozawa, S.; Yu, J.; Motegi, T. The Extraction of Nd from Waste Nd–Fe–B Alloys by the Glass Slag Method. J. Alloys Compd. 2003, 353, 189–193. [Google Scholar] [CrossRef]
  126. Hua, Z.; Wang, J.; Wang, L.; Zhao, Z.; Li, X.; Xiao, Y.; Yang, Y. Selective Extraction of Rare Earth Elements from Nd-Fe-B Scrap by Molten Chlorides. ACS Sustain. Chem. Eng. 2014, 2, 2536–2543. [Google Scholar] [CrossRef]
  127. Rhodia, Umicore Jointly Develop Process Recycling Rare Earth Elements. Available online: https://www.worldofchemicals.com/media/rhodia-umicore-jointly-develop-process-for-recycling-rare-earth-elements/875.html (accessed on 7 July 2023).
  128. AERC Recycling Solutions Partners with Global Tungsten and Powders Corporation to Reclaim Rare Earth Metals from Spent Fluorescent Lamps. Available online: https://www.prweb.com/releases/2012/9/prweb9851073.htm (accessed on 7 July 2023).
  129. USA Rare Earth and Texas Mineral Resources Corp Announce Opening of First Heavy and Light Rare Earths Processing Facility Outside of China. Available online: https://www.prnewswire.com/news-releases/usa-rare-earth-and-texas-mineral-resources-corp-announce-opening-of-first-heavy-and-light-rare-earths-processing-facility-outside-of-china-300975088.html (accessed on 7 July 2023).
  130. USA Rare Earth Acquires US Rare Earth Permanent Magnet Manufacturing Capability from Hitachi; Mine-to-Magnet—Green Car Congress. Available online: https://www.greencarcongress.com/2020/04/20200415-rareearth.html (accessed on 7 July 2023).
  131. Solvay to Develop Major Hub for Rare Earth Magnets in Europe. Available online: https://www.solvay.com/en/press-release/solvay-develop-major-hub-rare-earth-magnets-europe (accessed on 7 July 2023).
  132. MP Materials—MP Materials Awarded Department of Defense Heavy Rare Earth Processing Contract. Available online: https://investors.mpmaterials.com/investor-news/news-details/2022/MP-Materials-Awarded-Department-of-Defense-Heavy-Rare-Earth-Processing-Contract/default.aspx (accessed on 7 July 2023).
  133. DoD Awards $35 Million to MP Materials to Build U.S. Heavy Rare Earth Separation Capacity. Available online: https://www.defense.gov/News/Releases/Release/Article/2941793/dod-awards-35-million-to-mp-materials-to-build-us-heavy-rare-earth-separation-c/ (accessed on 7 July 2023).
  134. Energy Fuels Completes Acquisition of Rare Earth and Heavy Mineral Project in Brazil. Available online: https://www.prnewswire.com/news-releases/energy-fuels-completes-acquisition-of-rare-earth-and-heavy-mineral-project-in-brazil-301744927.html (accessed on 7 July 2023).
  135. Noveon Targets Rare Earth Magnet Production with Recycled Inputs. Available online: https://www.recyclingtoday.com/news/noveon-targets-rare-earth-magnet-manufacturing-using-recycled-material/ (accessed on 7 July 2023).
  136. CAT Net Zero Target Evaluations. Available online: https://climateactiontracker.org/global/cat-net-zero-target-evaluations/ (accessed on 7 July 2023).
  137. Chart: The Road to Net Zero. Available online: https://www.statista.com/chart/26053/countries-with-laws-policy-documents-or-timed-pledges-for-carbon-neutrality/ (accessed on 7 July 2023).
  138. Mercedes-Benz Says It Will Go All-Electric in 2030, but with a Major Caveat. Available online: https://www.theverge.com/2021/7/22/22588587/mercedes-benz-electric-ev-2030-caveat (accessed on 8 July 2023).
  139. General Motors, Maker of Giant Trucks and SUVs, Vows to Be Carbon Neutral by 2040. Available online: https://www.theverge.com/2021/1/28/22254116/general-motors-carbon-neutral-vow-2040-tailpipe-emissions-ev (accessed on 8 July 2023).
  140. Ford Says It Will Go All-Electric in Europe by 2030. Available online: https://www.theverge.com/2021/2/17/22287284/ford-electric-vehicles-ev-europe-2030 (accessed on 8 July 2023).
  141. Stellantis, Parent Company of Dodge and Jeep, Had an ‘EV Day’ and It Was Extremely Weird. Available online: https://www.theverge.com/2021/7/8/22568625/stellantis-jeep-dodge-ev-day-shark-drone-autonomous (accessed on 8 July 2023).
  142. Volvo Says It Will Only Sell Electric Cars by 2030. Available online: https://www.theverge.com/2021/3/2/22308851/volvo-electric-cars-only-2030-online-sales (accessed on 8 July 2023).
  143. Here Are the Biggest Announcements from Volkswagen’s Battery Event. Available online: https://www.theverge.com/2021/3/15/22325813/vw-volkswagen-power-day-battery-electric-car-announcement (accessed on 8 July 2023).
  144. BMW’s I4 Sedan Is the Electric 4 Series You’ve Been Waiting for. Available online: https://www.theverge.com/2021/6/1/22462504/bmw-i4-ev-price-specs-photos-release-date (accessed on 8 July 2023).
  145. Who Will Drive Electric Cars to the Tipping Point? Available online: https://www.bcg.com/publications/2020/drive-electric-cars-to-the-tipping-point (accessed on 8 July 2023).
  146. Alsauskas, O.; Dasgupta, A.; Connelly, E.; Gouy, A.; Le Marois, J.-B.; Mcdonagh, S.; Fernandez Pales, A.; Petropoulos, A.; Teter, J. Global EV Outlook 2023. Available online: https://www.iea.org/reports/global-ev-outlook-2023 (accessed on 8 July 2023).
Figure 1. Classification of REEs based on their application in magnets.
Figure 1. Classification of REEs based on their application in magnets.
Sustainability 15 14901 g001
Figure 2. Share of (a) REE reserves and (b) mined REE worldwide. Data for the plot were extracted from [13].
Figure 2. Share of (a) REE reserves and (b) mined REE worldwide. Data for the plot were extracted from [13].
Sustainability 15 14901 g002
Figure 3. World map showing countries with REE reserves as of 2023. Non-grayscale colors have no other significance than to highlight the countries indicated.
Figure 3. World map showing countries with REE reserves as of 2023. Non-grayscale colors have no other significance than to highlight the countries indicated.
Sustainability 15 14901 g003
Figure 4. Steps in the processing and refinement of REEs.
Figure 4. Steps in the processing and refinement of REEs.
Sustainability 15 14901 g004
Figure 5. Global permanent magnet production by value in 2022 [31].
Figure 5. Global permanent magnet production by value in 2022 [31].
Sustainability 15 14901 g005
Figure 7. Second−quadrant demagnetization curves [46].
Figure 7. Second−quadrant demagnetization curves [46].
Sustainability 15 14901 g007
Figure 8. Historical development of permanent magnets [48].
Figure 8. Historical development of permanent magnets [48].
Sustainability 15 14901 g008
Figure 9. Powder metallurgical processing of Nd-Fe-B-based magnets [49,50]. Grain boundary is abbreviated as GB.
Figure 9. Powder metallurgical processing of Nd-Fe-B-based magnets [49,50]. Grain boundary is abbreviated as GB.
Sustainability 15 14901 g009
Figure 10. Image of typical alloying and strip-casting equipment.
Figure 10. Image of typical alloying and strip-casting equipment.
Sustainability 15 14901 g010
Figure 11. Schematic of hydrogen decrepitation of strip-cast Nd-Fe-B alloy [52].
Figure 11. Schematic of hydrogen decrepitation of strip-cast Nd-Fe-B alloy [52].
Sustainability 15 14901 g011
Figure 12. Image of typical jet-milling equipment.
Figure 12. Image of typical jet-milling equipment.
Sustainability 15 14901 g012
Figure 13. Schematic of a jet milling process [52].
Figure 13. Schematic of a jet milling process [52].
Sustainability 15 14901 g013
Figure 14. Image of typical pulsed magnetizer equipment.
Figure 14. Image of typical pulsed magnetizer equipment.
Sustainability 15 14901 g014
Figure 15. Effect of HREE (Tb) substitution on the coercivity of Nd-Fe-B magnets as a function of temperature. Demagnetization curves of Nd-Fe-B and (Nd,Tb)-Fe-B at room temperature (a); Temperature dependence of the demagnetization plots of Nd-Fe-B (b) and (Nd,Tb)-Fe-B (c). Reprinted (adapted) with permission from [53]. Copyright 2010 Elsevier.
Figure 15. Effect of HREE (Tb) substitution on the coercivity of Nd-Fe-B magnets as a function of temperature. Demagnetization curves of Nd-Fe-B and (Nd,Tb)-Fe-B at room temperature (a); Temperature dependence of the demagnetization plots of Nd-Fe-B (b) and (Nd,Tb)-Fe-B (c). Reprinted (adapted) with permission from [53]. Copyright 2010 Elsevier.
Sustainability 15 14901 g015
Figure 16. Demagnetization behavior of a N55-grade Nd-Fe-B magnet [54].
Figure 16. Demagnetization behavior of a N55-grade Nd-Fe-B magnet [54].
Sustainability 15 14901 g016
Figure 17. Nd-Fe-B grades and corresponding HREE content [55].
Figure 17. Nd-Fe-B grades and corresponding HREE content [55].
Sustainability 15 14901 g017
Figure 18. Schematic of the grain boundary Dy diffusion process [49,50].
Figure 18. Schematic of the grain boundary Dy diffusion process [49,50].
Sustainability 15 14901 g018
Figure 19. Powertrain of 2022 BMW X5 PHEV [57].
Figure 19. Powertrain of 2022 BMW X5 PHEV [57].
Sustainability 15 14901 g019
Figure 20. Typical automotive shredding and recovery process.
Figure 20. Typical automotive shredding and recovery process.
Sustainability 15 14901 g020
Figure 21. The rising projection in e-waste over the coming years. The plot was made by extracting data from [59].
Figure 21. The rising projection in e-waste over the coming years. The plot was made by extracting data from [59].
Sustainability 15 14901 g021
Figure 22. Statistics of e-waste production and its sound management in the different continents in 2019. The plot was made by extracting data from [59].
Figure 22. Statistics of e-waste production and its sound management in the different continents in 2019. The plot was made by extracting data from [59].
Sustainability 15 14901 g022
Figure 23. Collection of EV motors from different states of the United States of America in 2019. Data adapted from [62]. The pie data (a) shows the collection of EV motors from all states of USA. (b) The percentage share of states having less than 1% [grouped into others in (a)].
Figure 23. Collection of EV motors from different states of the United States of America in 2019. Data adapted from [62]. The pie data (a) shows the collection of EV motors from all states of USA. (b) The percentage share of states having less than 1% [grouped into others in (a)].
Sustainability 15 14901 g023
Figure 24. (a) Collection of HDDs from different states of the United States of America in 2019. Data adapted from [62]. The pie data (a) shows the collection of HDDs from all states of USA. (b) The percentage share of states having less than 1% [grouped into others in (a)].
Figure 24. (a) Collection of HDDs from different states of the United States of America in 2019. Data adapted from [62]. The pie data (a) shows the collection of HDDs from all states of USA. (b) The percentage share of states having less than 1% [grouped into others in (a)].
Sustainability 15 14901 g024
Figure 25. Global/USA demands for (a) HDDs and (b) EVs.
Figure 25. Global/USA demands for (a) HDDs and (b) EVs.
Sustainability 15 14901 g025
Figure 27. Average magnet weight in different applications. The plot was made by extracting data from [75]. CV: conventional vehicle, EB: electric bike, DC: desktop computer, LC: laptop computer, SP: smartphone, P: printer, DC: digital camera, LS: loudspeaker, WM: washing machine and dryer, AC: airconditioner, RG: refrigerator, IP: industrial pump, EV: electric vehicle, IR: industrial robot, WT: wind turbine (onshore), WT*: wind turbine (offshore).
Figure 27. Average magnet weight in different applications. The plot was made by extracting data from [75]. CV: conventional vehicle, EB: electric bike, DC: desktop computer, LC: laptop computer, SP: smartphone, P: printer, DC: digital camera, LS: loudspeaker, WM: washing machine and dryer, AC: airconditioner, RG: refrigerator, IP: industrial pump, EV: electric vehicle, IR: industrial robot, WT: wind turbine (onshore), WT*: wind turbine (offshore).
Sustainability 15 14901 g027
Figure 28. Average magnet lifespan in different applications. The plot was made by extracting data from [75]. CV: conventional vehicle, EB: electric bike, DC: desktop computer, LC: laptop computer, SP: smartphone, P: printer, DC: digital camera, LS: loudspeaker, WM: washing machine and dryer, AC: airconditioner, RG: refrigerator, IP: industrial pump, EV: electric vehicle, IR: industrial robot, WT: wind turbine (onshore), WT*: wind turbine (offshore).
Figure 28. Average magnet lifespan in different applications. The plot was made by extracting data from [75]. CV: conventional vehicle, EB: electric bike, DC: desktop computer, LC: laptop computer, SP: smartphone, P: printer, DC: digital camera, LS: loudspeaker, WM: washing machine and dryer, AC: airconditioner, RG: refrigerator, IP: industrial pump, EV: electric vehicle, IR: industrial robot, WT: wind turbine (onshore), WT*: wind turbine (offshore).
Sustainability 15 14901 g028
Figure 29. Visual depiction of magnet usages and their proportions in different applications. The plot was made by extracting data from [76].
Figure 29. Visual depiction of magnet usages and their proportions in different applications. The plot was made by extracting data from [76].
Sustainability 15 14901 g029
Figure 30. Proposed material reuse and recovery in the VALOMAG project. Adapted from [77].
Figure 30. Proposed material reuse and recovery in the VALOMAG project. Adapted from [77].
Sustainability 15 14901 g030
Figure 31. Proposed robotic disassembly approaches to accelerate recycling and minimize critical material contamination. Adapted from [79].
Figure 31. Proposed robotic disassembly approaches to accelerate recycling and minimize critical material contamination. Adapted from [79].
Sustainability 15 14901 g031
Figure 32. Schematic of direct and extended recycling of end-of-life magnets.
Figure 32. Schematic of direct and extended recycling of end-of-life magnets.
Sustainability 15 14901 g032
Figure 33. Waste management hierarchy utilized in magnet recycling. Adapted from [89].
Figure 33. Waste management hierarchy utilized in magnet recycling. Adapted from [89].
Sustainability 15 14901 g033
Figure 34. Adopting the Lego design to produce a segmented pole to replace a solid pole: an example of direct reuse. Adapted from [80].
Figure 34. Adopting the Lego design to produce a segmented pole to replace a solid pole: an example of direct reuse. Adapted from [80].
Sustainability 15 14901 g034
Figure 35. Schematic of the indirect reuse of bonded magnets. Reprinted (adapted) with permission from [81]. Copyright 2019 Elsevier.
Figure 35. Schematic of the indirect reuse of bonded magnets. Reprinted (adapted) with permission from [81]. Copyright 2019 Elsevier.
Sustainability 15 14901 g035
Figure 36. Electricity usage to process one metric ton of permanent magnet using hydrometallurgical processing. The plot has been constructed by adapting data from [99].
Figure 36. Electricity usage to process one metric ton of permanent magnet using hydrometallurgical processing. The plot has been constructed by adapting data from [99].
Sustainability 15 14901 g036
Figure 37. Schematic representation of electrolytic dissolution of Nd-Fe-B magnets. Reprinted (adapted) with permission from [105]. Copyright 2018 American Chemical Society.
Figure 37. Schematic representation of electrolytic dissolution of Nd-Fe-B magnets. Reprinted (adapted) with permission from [105]. Copyright 2018 American Chemical Society.
Sustainability 15 14901 g037
Figure 38. Acid-free dissolution process to make magnets from magnet swarf and shredded HDDs. Reprinted (adapted) with permission from [111]. Copyright 2020 American Chemical Society.
Figure 38. Acid-free dissolution process to make magnets from magnet swarf and shredded HDDs. Reprinted (adapted) with permission from [111]. Copyright 2020 American Chemical Society.
Sustainability 15 14901 g038
Figure 39. Electricity usage to process one metric ton of permanent magnets using pyrometallurgical processing. The plot has been constructed by adapting data from [99].
Figure 39. Electricity usage to process one metric ton of permanent magnets using pyrometallurgical processing. The plot has been constructed by adapting data from [99].
Sustainability 15 14901 g039
Figure 40. VIM-HMS process to extract RE hydroxides. Reprinted (adapted) with permission from [123]. Copyright 2016 American Chemical Society.
Figure 40. VIM-HMS process to extract RE hydroxides. Reprinted (adapted) with permission from [123]. Copyright 2016 American Chemical Society.
Sustainability 15 14901 g040
Figure 41. Global EV sales over the years. Replotted by extracting data from [146].
Figure 41. Global EV sales over the years. Replotted by extracting data from [146].
Sustainability 15 14901 g041
Table 1. REEs containing natural ores [7].
Table 1. REEs containing natural ores [7].
MineralFormulaREO (wt. %)
Allanite (Y,Ln,Ca)2(Al,Fe3+)3(SiO4)3(OH)39
Apatite (Ca,Ln)5(PO4)3(F,Cl,OH)19
Bastnaesite(Ln,Y)(CO3)F75
Eudialyte Na4(Ca,Ln)2(Fe2+,Mn2+,Y)ZrSi8O22(OH,Cl)29
Fergusonite(Ln, Y)NbO453
Iimoriite Y2(SiO4)(CO3)68
KainositeCa2(Y,Ln)2Si4O12(CO3).H2O38
Loparite (Ln,Na,Ca)(Ti,Nb)O3 30
Monazite(Ln,Th)PO465
Mosandrite(Na,Ca)3Ca3Ln(Ti,Nb,Zr)(Si2O7)2(O,OH,F)433
ParisiteCa(Ln)2(CO3)3F261
Rinkite(Ca,Ln)4Na(Na,Ca)2Ti(Si2O7)2(O,F)220
Steenstrupine Na14Ln6Mn2Fe2(Zr,Th)(Si6O18)2(PO4)7.3H2O31
Synchysite Ca(Ln)(CO3)2F51
Xenotime YPO4 61
Zircon (Zr,Ln)SiO4 4
Table 2. REE reserves for different countries as of 2022 [13].
Table 2. REE reserves for different countries as of 2022 [13].
CountryMine Production
(Tons)
Total Reserves
(Tons)
% of Reserves
Mined
Share of
World
Reserves
United States43,0002,300,0001.871.83
Australia18,0004,200,0000.433.34
Brazil8021,000,0000.000416.71
Myanmar12,000---
Canada-830,000-0.66
China210,00044,000,0000.4835.01
Greenland-1,500,000-1.19
India29006,900,0000.045.49
Madagascar960---
Russia260021,000,0000.0116.71
South Africa-790,000-0.63
Tanzania-890,000-0.71
Thailand7100---
Vietnam430022,000,0000.0217.50
Other countries80280,0000.030.22
Table 3. Key activities investigating the extraction of REE from coal-based products and acid mine drainage.
Table 3. Key activities investigating the extraction of REE from coal-based products and acid mine drainage.
S. No.InstitutionKey Goals and Findings
1University of Illinois, USAA database of carbon ore, rare earth, and critical minerals (CORE-CM) has been developed by collecting datasets from several thousand samples along the Illinois Basin.
Development of technology to aid in separating RE materials operating at or near the mine-face.
[16]
2University of Kansas Center for Research, USAThe commissioning of a Geotek core scanner will assess CMs in the Cherokee forest City Basin. This includes regions of Kansas, Iowa, Missouri, Nebraska, and Oklahoma, and Osage Nation.[17]
3Collaborative Composite Solutions Corporation, USARevitalize the coal production for CM assessment in the Southern Appalachian Basin.
Parameters needed to determine REE security costs for the US in the coming years have been investigated.
[18]
4University of Texas, Austin, USAInvestigation of REE from coal mines and power plants in the US Gulf Coast Basin.
Investigations found significant REE in Gibbons Creek mine (TREE: 1000–8000 ppm) and San Miguel mine (TREE: 300–900 ppm).
[19]
5NETL, USAInvestigation of CORE-CM in Usibelli Coal Mine and Graphite Creek in Alaska.
Preliminary studies have indicated an encouraging amount of REE minerals in coal samples from the mines.
[20]
6LANL, USAQuantitative investigation of REEs in coal samples using laser-induced breakdown spectroscopy (LIBS).
Construction of LIBS in a backpack to quantitatively analyze the presence of REE in coal.
[21]
7West Virginia University, USADevelopment of pilot-scale plant for treating acid mine drainage to produce RE and CMs.
Extracted and separated samples demonstrate the presence of >67% HREE+CM.
807 tons of TREE production per year waste product is non-hazardous and can be disposed of on-site.
[22]
8Florida Polytechnic University, USAExtract REE from phosphoric acid sludge.
Demonstrated extraction of 90% REE and 100% heavy REEs.
Planned production capacity of 900 to 1100 tons of REM per year.
[23]
9Institute of Environmental Assessment and Water Research, Spain7.9 and 3.5 mg/L REE and Y concentrations obtained from Monte Romero and Almagrera acid mine drainage.
REE includes predominantly La and Ce.
[24]
10Institute of Earth Sciences, Germany1.8 and 2.5 mg/L concentrations of REE detected in Giessenbach Creek.
Higher concentrations of middle and heavy REEs were detected.
[25]
11Universidade de Aveiro, Portugal110, 120, and 124 mg/kg of REE located in Lousal mine area in the Iberian Pyrite Belt, Portugal.
14–20 mg/kg of HREE detected.
[26]
12IIT Kharagpur, India0.71 mg/L of REE detected in mine drainage from Jaintia Hills coalfields, India.[27]
13University of Huelva, SpainOrigin of REE traced in acid mine drainage from the Iberian Pyrite Belt (SW Spain).
20–30 mg/kg of REE detected in Felsic volcanic from Poderosa mine.
20–30 mg/kg of REE detected in Shales from the Perrunal mine.
[28]
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

Ormerod, J.; Karati, A.; Baghel, A.P.S.; Prodius, D.; Nlebedim, I.C. Sourcing, Refining and Recycling of Rare-Earth Magnets. Sustainability 2023, 15, 14901. https://doi.org/10.3390/su152014901

AMA Style

Ormerod J, Karati A, Baghel APS, Prodius D, Nlebedim IC. Sourcing, Refining and Recycling of Rare-Earth Magnets. Sustainability. 2023; 15(20):14901. https://doi.org/10.3390/su152014901

Chicago/Turabian Style

Ormerod, John, Anirudha Karati, Ajay Pal Singh Baghel, Denis Prodius, and Ikenna C. Nlebedim. 2023. "Sourcing, Refining and Recycling of Rare-Earth Magnets" Sustainability 15, no. 20: 14901. https://doi.org/10.3390/su152014901

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