Next Article in Journal
Nanoparticle Metrology of Silicates Using Time-Resolved Multiplexed Dye Fluorescence Anisotropy, Small Angle X-ray Scattering, and Molecular Dynamics Simulations
Previous Article in Journal
Processing of Thin Films Based on Cellulose Nanocrystals and Biodegradable Polymers by Space-Confined Solvent Vapor Annealing and Morphological Characteristics
Previous Article in Special Issue
Hybrid Plasmas for Materials Processing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Plasma Technologies in Recycling Processes

by
Reinosuke Kusano
1 and
Yukihiro Kusano
2,*
1
School of Physics and Astronomy, University of St Andrews, North Haugh, St Andrews KY16 9SS, UK
2
Department of Marine Resources and Energy, Tokyo University of Marine Science and Technology, Tokyo 108-8477, Japan
*
Author to whom correspondence should be addressed.
Materials 2024, 17(7), 1687; https://doi.org/10.3390/ma17071687
Submission received: 26 February 2024 / Revised: 24 March 2024 / Accepted: 2 April 2024 / Published: 7 April 2024

Abstract

:
Plasmas are reactive ionised gases, which enable the creation of unique reaction fields. This allows plasmas to be widely used for a variety of chemical processes for materials, recycling among others. Because of the increase in urgency to find more sustainable methods of waste management, plasmas have been enthusiastically applied to recycling processes. This review presents recent developments of plasma technologies for recycling linked to economical models of circular economy and waste management hierarchies, exemplifying the thermal decomposition of organic components or substances, the recovery of inorganic materials like metals, the treatment of paper, wind turbine waste, and electronic waste. It is discovered that thermal plasmas are most applicable to thermal processes, whereas nonthermal plasmas are often applied in different contexts which utilise their chemical selectivity. Most applications of plasmas in recycling are successful, but there is room for advancements in applications. Additionally, further perspectives are discussed.

1. Introduction

As billions of tons of industrial waste gradually gnaw at the health of the environment each year, the need to streamline and ameliorate existing recycling processes has become growingly more critical [1].
There are various types and kinds of waste emitted all over the world. They can be classified by the states of substances, uses, and materials. Taking Europe as an example, approximately 25.8 Mt of plastic wastes is generated every year, among which less than 30% is collected for recycling each year [2]. The majority of these are considered to be processed by primary recycling. Recycling processes for metals, which often use thermal plasmas, generally include sorting, squeezing/compacting, and melting in a furnace [3]. In 2018, 94 Mt of steel scrap was recycled. Other commonly recycled metals include aluminium (4.9 Mt), copper (1.6 Mt), lead, zinc, nickel, titanium, cobalt, chromium and specialty and precious metals (gold, silver, and platinum) [4].
Plasma, a reactive ionised gas known as the fourth state of matter [5], has proved to be very useful in the realm of waste treatment and recycling processes. Through the lens of recycling models, classifications, and economical models [6], this review discusses cutting-edge advancements of plasma technologies in recycling processes and evaluates their successes and shortcomings.

2. Classifications of Recycling

There are four main types of recycling: primary, secondary, tertiary, and quaternary recycling [7,8,9]. These are classified for plastics, but for the sake of analysis these definitions will be applied to general waste as well. There are different definitions in different literature, but the most commonly used definitions will be presented here.
Primary recycling is a closed-loop process, where the waste is only broken down to the level where its original chemical structure is not significantly modified [8,9]. It is a simple and cost-effective process. Recycling of polyethylene terephthalate (PET) bottles is a good example of primary recycling.
Secondary recycling involves downgrading the waste into less valuable materials [8,9]. Secondary recycling may be less desirable than primary, tertiary, or quaternary recycling, but is often otherwise inevitable. Secondary recycling is often employed to recycle inorganic substances, which are then mixed with concrete for reinforcement, for example.
Primary and secondary recycling are often collectively referred to as mechanical recycling [8,9]. However, this term may cause confusion or misunderstanding, since mechanical recycling generally involves not only mechanical processes but also heating and mixing (which may include melting), as well as chemical reactions. Hence, since the term is not rigorously defined in literature, this expression is not used in the present review.
Tertiary recycling refers to chemical recycling and thermal recycling. Thermal recycling, in which the waste is broken down to a molecular level [9] is different from combustion. Quaternary recycling involves the thermal combustion of the waste and aims to use the energy generated from this process [8,9]. Tertiary or quaternary recycling is often employed for wastes which are difficult to be recycled in a simple manner, often due to their complex structures or contents.
Different types of waste disposal models are also explored. First, there exists a model known as the waste management hierarchy model, as shown in Figure 1. This model ranks the desirability of different types of waste management, raking the most desirable at the top and the least desirable at the bottom. Taking plastic as an example and starting from the most preferred method:
  • Reduction (reducing the amount of plastic material produced in the first place);
  • Reuse (reusing the used plastic);
  • Recycling (breaking down the used plastic to its constituent components, corresponding to primary recycling);
  • Recovery (breaking down the used plastic to its molecular components or using them for energy production, corresponding to tertiary recycling, as well as the generation of energy from quaternary recycling);
  • Disposal (throwing away the used plastic).
Essentially, the purpose of this model is not too dissimilar to the circular economy model, which is discussed later: the fundamental concept is to minimise the amount of processing i.e., the amount of change required to make the waste reusable again, at least to some capacity.
What is recycled and what is discarded as waste depends on what kind of waste is being processed. For example, with plastics, one may keep the organic components as syngas (a gas mixture of H2 and CO), while in the recovery of metal waste one often discards organic components.
Additionally, a circular economy, as depicted in Figure 2, is considered. A circular economy is a circular model of production and consumption, aiming at extending the life cycle of products as much as possible [10]. Specifically, waste should be reduced to a minimum, while end-of-life products should be kept within the economy by recycling to create further value. The concept of circular economy helps increase the amount of reused or recycled plastic in the overall cycle [11], to ensure that one prevents the needless discarding of plastics as much as possible. This is achieved by creating closed loops within the waste management process, and “looping back” with a given piece of waste to a given point on the progression, such that as much of the waste goes through the cycle as possible. As can be seen, secondary, tertiary, and quaternary recycling do not form a closed circular loop.
Something to consider with regards to a circular economy is that smaller loop sizes correspond to higher efficiencies. This is because as the size of the loop increases, this inevitably leads to an increase in the energy used in the processes needed to break down the plastic to its constituent components.
In this model, extra materials or defective materials produced during the manufacturing process may also be taken into consideration. This is important; to start with raw materials that are initially unused makes reprocessing easier. Additionally, defective materials found during manufacturing can be very easily processed for recycling, since their components are already known by the manufacturer. Environmentally speaking, this results in a net positive; while the production of waste is inevitable and new materials have to be added to the system constantly, using materials of known origin boosts their reusability. With regards to the classifications of different recyclable products, it is also worth mentioning that the circular economy is applicable to organic and inorganic substances.
The four-level classification (primary, secondary, tertiary, and quaternary), waste management hierarchy, and circular economy are different in a few ways. The four-level classification groups different types of recycling in terms of the process utilised, while the waste management hierarchy groups recycling processes in terms of their environmental friendliness. On the other hand, the circular economy attempts to describe the dynamic flow of recycling processes. Therefore, the processes described in the four-level classification and waste management hierarchy appear in the circular economy model. Combining these different recycling models helps classify a given recycling process in terms of effectiveness, accessibility, and need.
One major drawback of the circular economy model is that there are many cases where a loop is not actually formed in the recycling process. Therefore, it must be understood that the circular economy model with a closed loop is an idealistic one.

3. An Overview of Plasma

Plasma is often known as the fourth state of matter that contains electrons, ions, and radicals [5,12], which in nature can be seen in, e.g., stellar structures [13,14] and the northern lights [15].
In the waste management community among others, plasmas are generally classified into thermal and nonthermal plasmas. Thermal plasmas, also called high-temperature plasmas, have high energy densities and are very productive in terms of processing speed. In these types of plasmas, the gas temperature and the electron temperature are high, and the plasma is at an equilibrium state or is almost at an equilibrium state. Additionally, the gas molecules inside a thermal plasma are fully or mostly ionised. Meanwhile, nonthermal plasmas, otherwise known as low-temperature plasmas, cold plasmas, or non-equilibrium plasmas [16], are only weakly or partially ionised. An important characteristic of the nonthermal plasma is that the electron temperature is much higher than those of ions and neutrals, and in fact, in some cases can be even higher than the electron temperature in a thermal plasma. The electron temperature in a nonthermal plasma can be, for example, 104–105 K, corresponding to ca. 1–10 eV, while the temperatures of the ions and neutrals can be as low as room temperature, or often up to several hundred K [17].
In the following two subsections, the different advantages and disadvantages of thermal and nonthermal plasmas are discussed in detail. In short, thermal plasmas are suitable for high-speed processes, while nonthermal plasmas are favourably used for selective chemical processes [18].

3.1. Thermal Plasmas

Thermal plasmas can be operated with very high power, and the reaction process is also very fast, since the voltage required to sustain them is very low. Plasma torches using thermal plasmas are commonly used for waste handling processes. The main disadvantage of thermal plasma processes is the fact that such a process is essentially a thermal process, and therefore is has low chemical selectivity; if given the opportunity to do so, it will react indiscriminately with all target particles. Different configurations for the generation of thermal plasmas are shown in Figure 3, Figure 4 and Figure 5. Figure 3 is the design for a direct current (DC) plasma torch, Figure 4 is the design for an inductively coupled radiofrequency (RF) plasma torch, and Figure 5 is the design for a microwave (MW) plasma torch. In all cases, thermal plasmas are generated by introducing electrical energy and are extended by a gas flow. Waste is then exposed to the extended plasma.
Spark plasma is also a thermal plasma that is similar to an arc plasma. These technical terms are often interchangeably used. However, arcs are continuous while sparks are momentary. Spark plasma is generated by applying a pulsed electric field between electrodes.
An overview of the types of plasma that can be used in recycling processes are given below in Table 1.

3.2. Nonthermal Plasmas

The main advantage of nonthermal plasma is its selectivity due to its high electron temperature [5]. This means that the desired reaction paths for a given process can be easily designed using this type of plasma. Since the gas temperature is low, nonthermal plasmas typically do not affect the bulk properties of solids or liquids but interact only with their surfaces [5,12,28,29]. Therefore, nonthermal plasmas are preferably applied to surface [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49] or gas processing [50,51,52,53,54], and their comparative safety to thermal plasmas also make them applicable to medical and oncological processes [55,56,57,58,59,60]. In fact, recent progress in plasma treatment of cancer may be considered a worthwhile addition to more popularly addressed cancer treatments like single-cell RNA sequencing [61].
The main shortcoming, however, is that nonthermal plasmas are unable to take a high energy density; hence, they are not as productive as thermal plasmas in terms of processing speed. Additionally, nonthermal plasmas are often operated at low gas pressure at which it is easy to generate such types of plasmas; the downside to this is that low gas pressures require expensive vacuum systems. Although plasma at atmospheric pressure is desirable, if one attempts to increase the gas pressure, the temperature of the gas and ions in the plasma can easily increase. This is because the increase in gas pressure leads to an increase in the frequency of inelastic collisions between the electrons and the other particles in the plasma. In order to sustain the plasma, the input power has to be increased. Therefore, at higher pressures, the heating of the gas is accelerated and the nonthermal plasma evolves to a thermal plasma [12] at a very high likelihood, although there are several techniques with which nonthermal plasma can be stably operated at a high gas pressure [5,12,28,29,62].
Nonthermal plasmas generally used for processing materials include low-pressure plasmas and atmospheric-pressure plasmas. Low-pressure plasmas are generally operated by DC, RF, or MW. Atmospheric-pressure plasmas, which do not require an expensive vacuum system, is preferred for recycling processes. Atmospheric-pressure plasmas may be generated using DC (which include pulsed operation), high frequencies up to but less than a megahertz, RF, or MW [62].
Different types of atmospheric-pressure plasmas have been developed, which include corona discharge, dielectric barrier discharge (DBD), cold plasma torches, and gliding arcs [28,62].
A corona discharge is generated in a significantly non-uniform electric field, usually appearing in the vicinity of sharp edges of electrodes, as can be seen in Figure 6. The discharge current must be limited to avoid the plasma becoming an arc. This is done by applying pulsed high voltages, resulting in an expensive power supply and low speed processing.
DBD is generated between electrodes by applying an alternating current (AC) voltage. At least one dielectric insulator is inserted between the electrodes to block DC. DBD is the most widely used atmospheric-pressure plasma in materials processing [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,51,63]. Two configuration of DBDs are known, as illustrated in Figure 7. As in Figure 7a, a volume discharge is generated at the gap between the electrodes. The gap is typically limited to several millimetres. A surface discharge is generated at a dielectric material surface with an electrode on each side, or with all electrodes embedded in the dielectric material, like in Figure 7b. Alternatively, two electrodes are attached on one side of the dielectric material, while the surface discharge is generated on its reverse side. In this case, a surface discharge is achieved by feeding a specific gas such as helium or argon to the reverse side. The frequency required to operate a DBD is typically between 500 Hz and 500 kHz, but DBDs at 50 Hz are also reported [42]. As the frequency is lowered, capacitive impedance increases at a fixed gap, and at a fixed applied voltage the power is lowered. Therefore, in general, a higher frequency is preferred for high-speed processing.
In most cases, DBDs consist of non-uniform filamentary discharges, and subsequently, the plasma treatment is not uniform. However, it is possible to generate uniform DBDs, and such a discharge is called a glow discharge. It is reported that a filamentary discharge can be avoided by using a noble gas such as argon or helium [28,29]. It may also be avoided using short pulses or by operating DBDs using RF. When RF plasmas are used, the dielectric is unnecessary. However, the operation of DBDs using RF requires an impedance-matching network to deliver RF energy to the plasma, usually limiting high voltage outputs. Due to the limitation of the high-voltage application, the electrode gap must be small enough for the stable operation of the discharge. In this respect, atmospheric-pressure RF plasma is rather preferably developed for cold plasma torches (depicted in Figure 8), which do not require the insertion of a specimen between the electrodes [64,65].
DBD and other plasmas, which are typically capacitively coupled plasmas, can be extended by a gas flow of argon or helium into an ambient air to realise cold plasma torches. They are typically weakly ionised [12,28]. A cold plasma torch can be operated at a high frequency (kHz) or RF in a DBD [12], or without covering electrodes with the dielectrics [12].
It is desirable that atmospheric-pressure plasma processing demonstrates a hybrid condition, which simultaneously achieves high chemical selectivity and efficient productivity [12,34,62]. In most cases, neither thermal nor nonthermal plasmas are able to provide them simultaneously. However, a gliding arc can achieve such a hybrid condition [5,12,16,34,62]. A gliding arc, as seen in Figure 9, is generated between diverging electrodes, and it is extended and quenched by a gas flow [12,16,27,63,66,67,68,69,70,71,72,73,74,75,76,77]. Atmospheric-pressure MW plasmas can also be generated and extended in a similar way. When the gas flow is fast enough to quench the plasma, its properties are similar to those of the gliding arcs [12]. However, using MW is argued to be less economical [18,62]. An industrially available open-air technology called technology “Plasmatreat®” is also a plasma generated as an arc that is extended and quenched by a gas flow, and its performance is similar to general gliding arcs.

3.3. Terminologies

There are inconsistencies in the literature regarding the classification of plasmas. There are, for example, publications claiming that high-temperature plasmas are only seen in stellar structures and fusion reactors and that group both thermal and nonthermal plasmas as low-temperature plasmas [78]. However, this claim requires reconsideration on a few fronts and may be cause for confusion.
Firstly, the claim is semantically undesirable. If thermal plasmas were low-temperature plasmas, they should not be called thermal plasmas in the first place.
Secondly, this claim is fallacious where astrophysical phenomena are considered. There are stellar prominences known as slingshot prominences, which are effectively coronal plasma condensations on rapidly rotating stars [79]. For the coronae of stars with temperatures at around 106 to 107 K, the temperatures of slingshot prominences are set at around 8500 K [79,80]. It is true that the coronae are at very high temperatures, but the temperatures of the prominences themselves can be easily achieved using DC thermal plasma torches, as their maximum temperature is somewhere around 20,000 K [81]. This is to say, according to the aforementioned claim, a slingshot prominence at 8500 K may be considered a high-temperature plasma, while simultaneously a DC-generated thermal plasma at 20,000 K is considered a low-temperature plasma. These are incompatible, and hence claiming that thermal and nonthermal plasmas are both low temperature is inappropriate. If high-temperature plasmas correspond to an environment of fusion reactions, it is suggested that this kind of plasma is called a fusion plasma.
Another cause for possible confusion regarding plasmas is the fact that plasma arcs and plasma torches have been differentiated, and plasma arcs have been categorised into transferred arcs and non-transferred arcs [78]. However, it is known that transferred and non-transferred arcs are essentially plasma torches [81]. In fact, the plasma torch uses an arc which is extended by a gas flow. Furthermore, the technical term “plasma arc” should be avoided, since an “arc” is an arc plasma. Hence, the categorisation and differentiation of plasmas into plasma arcs and plasma torches is unnecessary.
The term “plasma arc” originated from “plasma arc welding”, which is essentially the use of plasma torches. It is speculated that “plasma arc” just isolated itself from this term, as it sounds attractive on its own. However, in the realm of plasma physics it is desirable to append the word “plasma” as a suffix rather than a prefix, such as in, e.g., arc plasma, DBD plasma, etc.
To avoid these confusions in the recent literature, the present review takes the following definitions for plasma:
  • Thermal plasmas refer to high-temperature plasmas, which are at an equilibrium state or almost at an equilibrium state;
  • Nonthermal plasmas refer to low-temperature plasmas which are at a non-equilibrium state, or more specifically, whose electron temperature is much higher than the temperature of the gas.
These definitions are not unique—they have already been established in other literature reviews [82]. With this background, different types of plasma processes in recycling are reviewed, in conjunction with different types of economic models.

4. Thermal Plasmas in Recycling Processes

Thermal plasmas are most extensively used in recycling. They include plasma torches and spark plasmas. In both cases, plasma is used to effectively heat up solid waste. Plasma torches are commonly employed in plasma gasification (see Section 4.1) and metal recovery (see Section 4.2), while spark plasmas are used to sinter ceramics (see Section 4.3). Plasma torches are further classified into three different types in terms of their operation frequencies. They are DC, RF, and MW plasma torches as shown in Figure 3, Figure 4 and Figure 5. Table 2 summarises the differences in these plasma torches.
These are essentially the same with some minor differences, which make them suitable for specific situations. Detailed descriptions are presented in Section 4.1 and Section 4.2.

4.1. Plasma Gasification

A commonly used recycling process that utilises thermal plasmas is the thermal gasification process. Plasma gasification produces a syngas (H2, CO) which can be used as a fuel. The process also produces raw materials from which valuable resources, such as hydrocarbons or ammonia, may be synthesised. It is evident that this process corresponds to the recovery process in Figure 1, and therefore this is a tertiary recovery process in Figure 2.
The plasma gasification process is a leading method for solid-waste treatment [87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133] and has been reviewed quite extensively over the past several years [82,94,95,96,99,134,135]. This process, which usually involves thermal plasmas, is considered very effective and environmentally friendly [109,110,134,136,137,138,139,140,141,142,143,144,145,146]. One important thing to note about the plasma gasification process is that it is a purely thermal process.
While there technically do exist gasification processes that do not utilise plasmas, many cases where the term “gasification” is used implicitly refer to plasma gasification.
Done at high temperatures, plasma gasification entails the thermal decomposition of waste. In the case of inorganic compounds, these are usually oxidised and turned into slag, which can either be disposed of [147] or reused in materials like concrete [148], while the organic compounds are turned into a syngas which has a double merit; not only is it a fuel, but it is also cleaner than its more popular counterparts [87].
Syngas is environmentally friendlier than what is produced in gasification processes that do not involve plasma, since hydrocarbons are produced for such processes. The burning of hydrocarbons results in the production of CO2, whereas burning the hydrogen gas in syngas from plasma gasification only produces H2O [107]. In the past, CO was also used as a fuel [149], but it has recently been avoided due to its toxicity [150]. It is also worth noting that while the components in the syngas may be used as a clean fuel, they can also be used to make, e.g., ammonia (NH3), methane (CH4), methanol (CH3OH), formaldehyde (HCHO), or formic acid (HCOOH).
A simplified schematic diagram of a typical plasma gasification process is shown in Figure 10.
Plasma gasification using thermal plasma involves a plasma torch, in which an electric arc is generated and extended by a gas flow to heat the waste. The gas used to extend the electric arc is often air, or oxygen-rich air. Sometimes, argon (Ar) is introduced as well so that the ignition becomes easier at a lower voltage [12].
As for the plasma itself, its temperature can generally reach around 13,000 °C but touches the waste in question at a much lower temperature, between 2700 to 4500 °C, which is more than enough to decompose the target waste [151].
In recent publications, one can often find the terms “plasma pyrolysis” and “plasma gasification”. Technically speaking, plasma pyrolysis is thermal decomposition of target waste in “the near absence of oxygen” [152], while the presence of oxygen is encouraged in plasma gasification. However, these two terms are often used as synonyms [81,82,135,153], and hence they are used synonymously in this review as well.
As illustrated in Figure 3, Figure 4 and Figure 5, there are three main types of plasma torches that are used for plasma gasification: DC plasma torches, RF plasma torches, and MW plasma torches [81].
Due to the availability of large-scale power supplies, DC plasmas are capable of generating a power output in the range of megawatts. Their main shortcoming is that since the temperature of their arc is so high, the electrodes start to erode. Hence, impurities are generated in the arc, and this affects the quality of the products which are processed by DC plasmas [81].
RF plasmas can be generated in two ways: it can either be generated as an inductively coupled plasma (ICP), or as a capacitively coupled plasma (CCP). In the ICP, the time-varying magnetic field induces an electric field, whereas in the CCP, the electric field is directly oscillated. The electron density can be higher in the ICP, whereas with the CCP, a DC self-bias can be generated with an appropriate impedance-matching network, such that the ions in the plasma are accelerated toward the powered electrode. This can be used for physical etching of the target material. One must consider here that in gasification, a rise in temperature is necessary; therefore, the ICP is more suitable in this case [154]. If it is an ICP, RF plasmas are operated without the use of an electrode, and hence do not suffer the disadvantage of generating impurities—the caveat being that they are less efficient compared to DC plasma, generating only a few hundred kilowatts [81,135]. Additionally, there are some issues with impedance matching [81,135].
MW plasmas can have a higher plasma density than RF plasmas, and their plasma zone can also be much larger [78]. Their other advantages include that they are highly effective and only require a “simple and robust” design for their reactor [82], while also being a low-voltage operation. Additionally, they do not have any issues with impedance matching, and they do not require electrodes either [135]. These are the reasons why some researchers in the field consider MW plasmas to be the most superior type of thermal plasmas [135]. However, MW plasmas do come with their own set of shortcomings. First, MW plasmas generate the least amount of power out of the three types; they are generally able to generate up to only 10 kW of power [81]. Furthermore, while they require the least amount of power in sustaining the arc, the initial ignition of the arc requires a comparatively high voltage [135] and thereby a separate ignition system, while neither DC nor RF plasma requires this.
It is interesting to see just how many areas of waste management plasma gasification is used in: the treatment of medical waste [141,155,156] and municipal solid waste (MSW) [142,157], as well as deriving fuel from enhanced landfill mining (ELM) [158] and from the conversion of coal [159] and biomass [160,161], as well as plenty more.
However, plasma gasification proves to be disadvantageous through the lens of economic viability [113,162]. With regards to this, it is highly debatable whether plasma gasification is necessarily the most efficient method of waste decomposition. As will be discussed in Section 5.1, although low-temperature plasma (i.e., nonthermal plasma) may also be used to gasify waste [91,96,118], the plasma gasification process primarily involves thermal plasma which, as stated previously, inevitably involves an extremely high processing temperature [96,118]. Additionally, it is not always the case that the waste product needs to be decomposed into its molecular constituents, at least not to the extent that plasma gasification allows. Here, one must take the waste management hierarchy model and the circular economy model into consideration. If waste was decomposed to its molecular components every time, the process is not as energy-efficient or as cost-effective as reusing a less-decomposed part of the original material.
In terms of the waste management hierarchy model [163] in Figure 1, plasma gasification corresponds to the recovery step, which is far from ideal. In other words, purely economically speaking, plasma gasification is not the most ideal method of recycling using plasmas. However, it is worth noting that the major advantage of plasma gasification renders the sorting of waste into specific categories unnecessary or at least reduces its need, since the waste is decomposed into its molecular constituents anyway.
Currently, a motivating factor of plasma gasification research seems to be to maintain the high gas temperature while decreasing the input power [135], as the high input power is a major setback of the thermal plasma. However, since this type of plasma is inherently characterised by a high input power and a high pressure, this is not an optimal solution. One can only hope that the theoretical framework of plasma gasification is further developed to fix its setbacks [82], and that the development of numerical models [142,156,157,161,164,165,166,167,168,169,170] will optimise the efficiency and effectiveness of this process in the future. For example, it may be advantageous for future numerical methods to simulate a situation in which a temperature limit is set for the plasma discharge, such that the target material is selectively thermally decomposed. For instance, selecting an appropriate temperature window may facilitate selectivity [142,156,157,161,164,165,166,167,168,169,170].

4.2. Recovery of Metals

Plasma processes are also commonly used in the recovery of inorganic materials. For example, as mentioned earlier, slag can be recovered in the gasification process for use in construction.
Thermal plasma may also be used in the recovery of metals from waste. Most of these cases involve thermal plasma torches. The recovery of metals also corresponds to the recovery process in Figure 1 and therefore corresponds to tertiary recycling in Figure 2. Pyrometallurgy and hydrometallurgy are commonly used techniques for recycling metal. Pyrometallurgy is a thermal recycling technique that extracts metals from electronic waste, often using thermal plasmas. In other words, if pyrometallurgy is employed, a thermal plasma process is often involved. Hydrometallurgy is a wet process using liquids. It does not require high temperatures but emits significant amounts of waste liquids.
For example, the recovery of metals from household uses, industrial applications such as steel, aluminium, and copper, and electronic waste is one of the most popular applications of thermal plasma [3,171,172,173,174,175,176,177], and it is commonly performed with secondary batteries and waste circuit boards. However, the recycling of secondary batteries is complicated, so it is discussed further in Section 5.3. For waste circuit boards, a plasma torch is used to ensure that the organic molecule macrostructures in the circuit board are destroyed to generate a syngas (plasma gasification), and the bonds connecting the metals and fibreglass boards are also broken efficiently [172]. However, the main purpose of this process is not in the generation of combustible gases but rather the recovery of metals from the waste. Indeed, the remaining metals exist in a mixture [175,177]. This mixture of metals can then promptly be coupled with acid leaching and a depolariser to yield desirable metals like copper, aluminium, and iron [177]. DC plasma torches are most often used for this process, as the high temperature improves the separation of the slag metal [175]. In all cases, thermal plasma torches are used, similarly to the plasma gasification process.
There are of course other applications of plasmas in the recovery of metals from different types of waste. For example, another application of plasmas in this area is the recovery of pig iron from red mud [178]. Red mud is a hazardous residue that is separated from sodium aluminates during the Bayer process [178,179]. A DC plasma torch is used for this process [178].
Similarly, electroplating sludges usually come from electroplating plants and consist of a large quantity of heavy metals [180,181]. While metals like chromium and nickel among others can be obtained from this waste using vitrification [180], in-flight treatment of the waste with a DC plasma torch has also been used to obtain these same metals [182].
For galvanic sludge, a byproduct of industrial metal surface treatment, a DC plasma torch is used for the decomposition [1,175].
Lastly, aluminium dross (or slag), which is considered an odorous hazardous waste [1,175,183], may be treated using plasmas. This may achieved by using a plasma torch to recover aluminium at a very high efficiency [175,184], or by applying an MW plasma torch to produce calcinated alumina [183].
For a comprehensive review of these processes, consult the review by Changming [175].
It is important to note, however, that some also consider the decomposition of zircon to fall under this category [175]. This is not the case. It is indeed true that zircon refractories may be thermally decomposed using plasmas [185,186], but these processes do not belong in this waste decomposition category as zircon is not waste. Zircon is the source for zirconia, a mineral with many uses [187], and additionally has been considered as one of the leading hosts for radioactive waste [188,189]. Hence, zircon is not waste, and should not be considered as such.

4.3. Spark Plasma Sintering

Spark plasma is used to sinter ceramics. The process is called spark plasma sintering, which is also called plasma pressure compaction or pulsed electric current sintering, and it is additionally referred to as a field-assisted sintering technique [190,191,192,193,194,195,196,197,198,199,200,201,202,203,204]. Spark plasma sintering is a pressure-assisted sintering process, in which a material is subjected to a simultaneous uniaxial pressure and high-intensity pulsed current induced by a spark. It is similar to hot pressing, but in spark plasma sintering, thermal energy is directly provided to a material by a current flow in a plasma. The original concept was proposed in 1912 [205], but its practical use started in the 1980s after the first laboratory-scale systems were developed [206]. The development of commercial spark-plasma-sintering systems induced a wide range of use for sintering.
Spark plasma sintering is also applied to waste processes. Spark plasma sintering is proposed as a solid-state recycling technique [207], exemplifying the consolidation of aluminium alloy scraps [208]. Calcareous waste concrete powder is processed by spark plasma sintering [209], improving the flexural strength and modulus of the sintered bodies. It is reported that rice straw agricultural waste can be processed by spark plasma sintering to produce aluminosilicate matrices for caesium immobilization [210]. Calcium silicate matrices are manufactured from boric acid production waste using spark plasma sintering for removing and immobilizing 60Co [211].

5. Nonthermal Plasmas in Recycling Processes

Nonthermal plasmas are characterised by their high electron temperatures and low gas temperatures, as listed in Table 1. The electron temperature is high enough to affect the chemical bonding of molecules, while the gas temperature may be as low as room temperature. They are therefore suitable for surface chemical processes and gas-phase chemical processes. Gas-phase reactions in nonthermal plasmas typically involve several hundred reactions [212]. On the other hand, since the gas temperature is low, nonthermal plasmas do not exhibit optimal performance if a thermal process is desired for solids or liquids.
Nonthermal plasmas may also be applied to waste recycling processes. A number of their applications are listed below, and are discussed in detail in the following subsections:
  • Plasma reforming;
  • Wind turbine blade recycling;
  • Battery recycling;
  • Paper recycling.
In short, plasma reforming is an example of gas-phase reactions. In wind turbine blade recycling, there has been an attempt to use nonthermal plasma for a moderate thermal process with less success. Recycling processes for lithium-ion batteries and papers employ plasma surface modifications.

5.1. Plasma Reforming

Plasma reforming is a very broad term that entails the production of usable fuels and raw materials. Reforming is not only used for waste handling, but also the conversion of organic compounds like methane or other hydrocarbons and alcohols to hydrogen, syngas, or valuable organic molecules [213,214,215].
Plasma reforming using nonthermal plasmas, at least in the literature, considers only gas-phase reactions, since bulk properties of solids or liquids are not significantly affected by the nonthermal plasma. Plasma reforming encompasses both (nonthermal) plasma gasification and plasma liquefaction; these processes are possible using nonthermal plasmas, by tuning the selectivity of reactions. The major difference between nonthermal plasma gasification and plasma liquefaction is the state of the main product at room temperature, whether it be gas or liquid; otherwise, they are essentially the same. The term “plasma gasification” is mostly used for the thermal plasma process. To avoid confusion, this review collectively refers to both nonthermal plasma gasification and nonthermal plasma liquefaction as plasma reforming.
Nonthermal plasmas are often used in the production of hydrogen from waste plastics [216,217,218,219], or for the conversion of methane [220,221,222,223] or other organic materials to hydrocarbons [224,225]. A high selectivity is needed for such a process, and therefore, nonthermal plasmas are desirable.
It is interesting to see that plasma reforming is often used for the treatment of organic materials [224,225,226,227,228,229]. Namely, gasification or liquefaction via nonthermal plasmas is often used in combination with biomass [230,231]. For example, this process is used to combine biorefinery with the chemical recycling of plastics [232], to upgrade biomass vapours so that biofuels may be produced [230], or for the refinery of biomass by removing tar from it [231]. The last process particularly benefits from the plasma reforming process, as the level of tar produced from this process is less than 0.1% of the level of tar in usual plasma gasification processes [96].
Although it is reported that thermal gasification is also used in processes like the decomposition of municipal waste [96], it is in fact an example of cold plasma pyrolysis, a thermal process operated at around 600 °C that decomposes plastic waste with the help of biological catalysts [91,233]. At this range of temperatures, this plasma process is regarded as a nonthermal plasma process.
While admittedly the energy required for this process is significantly lower than if one were to use thermal plasmas, the same shortcomings regarding economic viability in terms of the two models still apply. This is to say, it still corresponds to the recovery step in the waste management hierarchy model in Figure 1. There is also an additional shortcoming regarding the efficiency of gasification; since gasification is a purely thermal process, the low temperature of the nonthermal plasma is not advantageous to the speed of this procedure. Hence, gas production using nonthermal plasmas is less efficient than with thermal plasmas.
It is important to understand that plasma gasification is favourable if syngas or hydrogen are desirable products. However, some processes may require the use of syngas as a raw material in an organic compound. Since plasma reforming allows for the direct formation of favourable compounds, this is the more appropriate process for such instances.
While plasma gasification produces a syngas, the corresponding main product produced in plasma liquefaction is liquid fuel. It is evident that this process, much like plasma gasification, also corresponds to the recovery process in Figure 1 and is therefore also a tertiary recovery process. The production of liquids is not possible using thermal plasmas, since the temperatures achieved by thermal plasmas are too high for the target product fuel to remain as a liquid that can still be further decomposed.
Plasma liquefaction using nonthermal plasma is a process by which solid biomass is transformed into a liquid form, with the help of catalysts and solvents. This is an attractive process, since producing liquid forms of fuels is advantageous [234]. Note the two following reasons. Firstly, in applications other than in fuel, there is a preference to keep biomass in a form where their molecules are decently large: this usually corresponds to a liquid form. Secondly, transportation becomes easier when biomass is in a liquid form; the transport of gaseous fuels requires containers that can sustain high gas pressures.

5.2. Wind Turbine Blade Recycling

There are two major challenges in the recycling of wind turbine blades. The first challenge is that wind turbine blades have a complicated structure, and the separation of components can prove to be quite difficult [235]; therefore, primary recycling of these blades is challenging. The second challenge is that wind turbine blades are partially made from fibre-reinforced polymers (FRP), typically with high contents of inexpensive glass fibres.
FRPs are widely used in aerospace and aeronautics [236,237], construction [238,239,240], transportation [241,242], and of course, wind energy, among many other areas. This is due to their light weight, environmental friendliness, mouldability, resistance to corrosion, and general mechanical properties [242]. FRPs are mainly composed of two materials: 1. fibres that are made of glass, carbon, or natural fibre [243] among others and 2. polymer matrices, which are either thermoplastics or thermosets [244]. The fact that there are various combinations for FRPs provides different challenges in their decomposition and separation. Recycling by way of heating proves to be particularly difficult when thermosets are present in the FRP, as thermosets remain hard when exposed to heat [245]. In wind turbine blades, thermosets are generally used.
There are two main ways to separate glass fibres: first, by retaining the original shape of the fibre, e.g., recycling glass fibres by using nonthermal plasma, or second, by melting it at a high temperature [244]. In the former case, the physical properties of the fibre are severely degraded, while the latter may only be performed at the laboratory level [241]. Additionally, the fact that glass fibre is not so expensive in the first place decreases the merit of melting it, especially considering the cost associated with the recycling process. The volume fraction of polymers in wind turbine blades is not very large. Although polymer matrices in wind turbine blades are much more valuable for recycling than glass fibres, this means that the plasma gasification of wind turbine blades produces significant amounts of degraded glass while also producing a limited amount of syngas, even in comparison to the regular plasma gasification process, which is unfavourable. Therefore, the wind energy community considers that it is cheapest to directly dispose of FRPs in landfills, and this is what is done in the majority of cases [246,247,248,249]. In this respect, it may be considered that wind energy is not an environmentally friendly green technology. Admittedly, FRPs are also used in electronic parts including circuit boards, whose wastes are generally handled by thermal plasma as discussed before. The important difference when recycling electronic wastes is that it primarily aims at recycling valuable metal components.
From this, it is not surprising to see that plasmas are not commonly used to recycle FRPs in wind turbine blades. There have been attempts to upcycle the carbon fibre in FRPs via pyrolysis, but the problem is that the upcycled carbon fibre is still degraded in comparison to virgin carbon fibres. Even though the use of thermal plasma somewhat mitigates the amount of graphitization that is usually seen [250], the previous issues regarding heat degradation still stand. However, the recovery of carbon fibres from carbon FRPs via chemical swelling and plasma treatment has been performed with success; compared to the original fibres, the recycled carbon fibres had 93.55% tensile strength and 118.76% interfacial shear strength [251].
Recycling entire wind turbine blades has also been attempted [252,253], and while researchers seem to be hopeful despite the degraded recycled products, the fact that most wind turbine waste is disposed of in landfills suggests that research in this field has proved to be unsuccessful so far.

5.3. Recycling Lithium-Ion Batteries

Lithium-ion batteries (LIBs) are referred to as secondary batteries, and are widely used in cars, mobile phones, and the like.
Plasmas are commonly used in the production processes of LIBs [6,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277]. Both thermal plasmas [278,279,280] and nonthermal plasmas [281,282,283,284,285] are used for these production processes. The recyclability of LIBs is limited, and there are also many challenges that such a process would entail, which include the variety of battery designs and the limitations of current recycling technologies [286,287]. There are many expensive components in a LIB that are preferably recovered rather than recycled, which complicates the recycling process further [286]. Therefore, components are selectively recycled, and for LIBs, this is the case for metal components.
Broadly speaking, there are two different methods of recycling LIBs:
  • A heat-based smelting process (pyrometallurgy);
  • A liquid-based leaching process (hydrometallurgy).
Of these, smelting is a thermal process for which thermal plasmas are often used [288]. However, nonthermal plasmas have also been used to recycle LIBs, albeit very briefly: for example, to heat and combust the byproduct gas in an LIB recycling process [287], and the used graphite anode in the LIB has been shown to be fully functional after reactivation using a plasma jet [255]. According to the corresponding patent, the temperature of the plasma torch is 1150 °C (ca. 900 K) [289]. Such a temperature can be easily achieved using a quenched gliding arc. So, in this review, it is assumed to be a nonthermal plasma. This process can be broadly considered as a part of recycling techniques, as properties of substances or materials under the subjects are somehow recovered or improved so that they are ready to use and or reuse. On a tangent, it is very important to consider that techniques which have previously been considered out of the scope of recycling may potentially be applied in relevant recycling processes.
The latter process would be an example of reuse, as in Figure 1 and Figure 2.

5.4. Paper Recycling

De-inking used paper is a useful and environmentally conscious process in recycling [290,291] and may be considered a tertiary recycling process that corresponds to the recovery and recycling processes in Figure 1 and Figure 2. While biological catalysts like cutinases are commonly used [290,291], nonthermal plasmas, like atmospheric-pressure helium DBD plasmas, are also employed to speed up these catalytic reactions in recycling [292]. One may use MW low-pressure plasma to increase the hydrophilicity of the paper and decrease its disintegration time. This treatment decreases the energy consumption of the de-inking process [292,293].
In publications that discuss de-inking, references are sometimes made to “plasma jets”. This is synonymous to a cold plasma torch, or a nonthermal plasma torch. In recent years, it has been discovered that cold plasma torches that use helium are more efficient at increasing the hydrophilicity of polymethyl methacrylate (PMMA) [294] than those that use argon. This increase in wettability was found through surface characterization techniques, using the contact angle measurements of PMMA with water [294]. When surface modification plays an important factor in finding a solution in recycling processes, the concepts of surface energy and surface tension, as well as simple surface characterisation techniques such as contact angle measurements, are considered to be important [295,296,297,298,299,300,301,302,303,304,305,306].
These examples indirectly assist recycling processes. Enthusiastically considering these indirect processes as plasma recycling techniques may lead to novel applications.

6. A Critique of Applying Plasmas in Recycling Processes

In general, thermal plasmas are suitable for thermal processes, while nonthermal plasmas better suit applications in surface processes and gas-phase reactions. Since each advantage is distinctively different, one possible perspective to synergistically combine the processes of thermal and nonthermal plasmas in recycling can be to employ a thermal plasma to generate gases, which are afterwards treated by a nonthermal plasma. Alternatively, after inorganic solids are processed by a thermal plasma, their surfaces may be modified by a nonthermal plasma.
Both thermal and nonthermal plasma processes are known to be environmentally friendly, dry processes that do not use solvents. However, one cannot neglect the expenses induced by employing them into any specific application. The investment cost including a plasma generator and a chamber usually dominates the expense of the process. It means that if the plasma system used is not efficiently used for mass production, the investment cost may become high.
Another important issue is the use of electricity within the running cost. Plasmas used for materials processing are electric gas discharges, which inevitably consume a certain amount of electric energy. If the mass consumption of electrical energy is an issue, the use of plasma should be avoided.
It is often the case that costs are the major driving force for selecting a process, and not the effectiveness of the technologies themselves. Since this is also the case for recycling processes, it is difficult to foresee future trends. For example, owing to recent serious developments in the generation and storage of electricity, as well as technological advancements, the above concerns regarding expenses may become minor issues.
From the viewpoints of the recycling classifications discussed in Section 2, it is desirable not to destroy waste down to a molecular level, but to preserve bulk properties. In this sense, nonthermal plasmas will play a more important role in solutions of recycling. However, in order for nonthermal plasmas to play more important roles in recycling, an overall change is required with respect to the design, manufacture, and suitability of the treatment to a given process.

7. Conclusions

When thermal plasmas are used for waste management, these are generated in plasma torches in all cases excluding spark plasma sintering. While the terminology differs from publication to publication, it is seen that the processes are essentially all the same. It has also been noted that when a large quantity of waste is thermally processed, this is most efficiently achieved by using a thermal plasma; this also helps achieve the maximum temperature. This is because the electrical impedance of thermal plasmas is low, so a high current can be generated from a low voltage. This leads to a high-power process, in which a high-temperature environment can easily be achieved from the Joule heat. With regards to DC, RF, and MW plasmas, if the contamination from the electrodes melting is not a significant issue, DC plasmas are the most efficient for large-scale processes.
The shortcomings of thermal plasmas are that a high power is needed to achieve a high temperature, and that they need an environment or a container resistant to high temperature. It is advisable to consider these limitations. Additionally, the residue waste after gasification will also be exposed to high temperatures; for cases in which this is not appropriate, it is important to select a process that does not reach very high temperatures.
The number of attempts that use nonthermal plasmas in recycling processes are much less than for thermal plasmas. This is due to the fact that nonthermal plasmas are unsuited for the high-speed processing of bulk solids or liquids. However, considering their selectivity, gas-phase reactions, and their affinity for surface reactions with solids and liquids, there are applications that are able to utilise these advantages.
Alternatively, one may consider nonthermal plasmas in processes that assist the main recycling process. In these cases, nonthermal plasmas may be found to be increasingly more useful. In reality, nonthermal plasmas are already used in recycling processes, even if they are not explicitly defined as such. In order to improve recycling processes in the future, this may be something worth considering.

Author Contributions

Conceptualization, Y.K.; investigation, R.K. and Y.K.; data curation, R.K. and Y.K.; writing—original draft preparation, R.K.; writing—review and editing, R.K. and Y.K.; visualization, R.K.; supervision, Y.K. 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

Data sharing is not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MDPIMultidisciplinary Digital Publishing Institute
DCDirect current
ACAlternating current
RFRadiofrequency
MWMicrowave
ICPInductively coupled plasma
CCPCapacitively coupled plasma
MSWMunicipal solid waste
ELMEnhanced landfill mining
DBDDielectric barrier discharge
LIBLithium-ion battery
PMMAPolymethyl methacrylate
FRPFibre-reinforced polymer

References

  1. Heberlein, J.; Murphy, A.B. Thermal plasma waste treatment. J. Phys. D Appl. Phys. 2008, 41, 053001. [Google Scholar] [CrossRef]
  2. A European Strategy for Plastics in a Circular Economy. Available online: https://www.europarc.org/wp-content/uploads/2018/01/Eu-plastics-strategy-brochure.pdf (accessed on 22 March 2024).
  3. Burkhard, R.; Hoffelner, W.; Eschenbach, R.C. Recycling of metals from waste with thermal plasma. Resour. Conserv. Recycl. 1994, 10, 11–16. [Google Scholar] [CrossRef]
  4. Metal Recycling Factsheet. Available online: https://circulareconomy.europa.eu/platform/sites/default/files/euric_metal_recycling_factsheet.pdf (accessed on 22 March 2024).
  5. Fridman, A.; Kennedy, L.A. Plasma Physics and Engineering; Taylor & Francis: New York, NY, USA, 2004. [Google Scholar]
  6. Wang, C.; Wu, L.; Wang, H.; Zuo, W.; Li, Y.; Liu, J. Fabrication and shell optimization of synergistic TiO2-MoO3 core-shell nanowire array anode for high energy and power density lithium-ion batteries. Adv. Funct. Mater. 2015, 25, 3524–3533. [Google Scholar] [CrossRef]
  7. Kumar, R. Tertiary and quaternary recycling of thermoplastics by additive manufacturing approach for thermal sustainability. Mater. Today Proc. 2021, 37, 2382–2386. [Google Scholar] [CrossRef]
  8. Rajab, M.A.; Shaban, S.Y.; Hussen, L.J.H. Recycling and improving the environmental impact of plastic waste. Am. J. Eng. Res. 2018, 7, 131–134. [Google Scholar]
  9. Schyns, Z.O.G.; Shaver, M.P. Mechanical recycling of packaging plastics: A review. Macromol. Rapid Commun. 2020, 42, 2000415. [Google Scholar] [CrossRef] [PubMed]
  10. Circular Economy: Definition, Importance and Benefits. Available online: https://www.europarl.europa.eu/topics/en/article/20151201STO05603/circular-economy-definition-importance-and-benefits (accessed on 21 March 2024).
  11. Johansen, M.R.; Christensen, T.B.; Ramos, T.M.; Syberg, K. A review of the plastic value chain from a circular economy perspective. J. Environ. Manag. 2022, 302, 113975. [Google Scholar] [CrossRef] [PubMed]
  12. Kusano, Y. Atmospheric Pressure Plasmas for Polymer Surface Modification-Alternating Current Gliding Arcs and Ultrasound Enhanced Plasmas; Department of Wind Energy, Technical University of Denmark: Lyngby, Denmark, 2019. [Google Scholar]
  13. Daley-Yates, S.; Jardine, M.M.; Johnston, C.D. Heating and cooling in stellar coronae: Coronal rain on a young Sun. Mon. Not. R. Astron. Soc. 2023, 526, 1646–1656. [Google Scholar] [CrossRef]
  14. Jardine, M.; Cameron, A.C. Slingshot prominences: Nature’s wind gauges. Mon. Not. R. Astron. Soc. 2019, 482, 2853–2860. [Google Scholar] [CrossRef]
  15. Haerendel, G. My dealings with the aurora borealis. Front. Astron. Space Sci. 2022, 9, 1033542. [Google Scholar] [CrossRef]
  16. Kusano, R.; Kusano, Y. Hybrid plasmas for materials processing. Materials 2023, 16, 4013. [Google Scholar] [CrossRef] [PubMed]
  17. Lieberman, M.A.; Lichtenberg, A.J. Principles of Plasma Discharges and Materials Processing; Wiley: New York, NY, USA, 1994. [Google Scholar]
  18. Fridman, A.; Nester, S.; Kennedy, L.A.; Saveliev, A.; Mutaf-Yardimci, O. Gliding arc gas discharge. Prog. Energy Combust. Sci. 1999, 25, 211–231. [Google Scholar] [CrossRef]
  19. Prokisch, C.; Bilgic, A.M.; Voges, E.; Broekaert, J.A.C.; Jonkers, J.; van Sande, M.; van der Mullen, J.A.M. Photographic plasma images and electron number density as well as electron temperature mappings of a plasma sustained with a modified argon microwave plasma torch (MPT) measured by spatially resolved Thomson scattering. Spectrochim. Acta Part B At. Spectrosc. 1999, 54, 1253–1266. [Google Scholar] [CrossRef]
  20. Kuthi, A.; Liu, J.; Young, C.; Lee, L.C.; Gundersen, M. Chapter 3. Pseudospark-Based Pulse Generator for Corona-Assisted Combustion Experiments. In Combustion Processes in Propulsion—Control, Noise, and Pulse Detonation; Elsevier: Burlington, MA, USA, 2006; ISBN 978-0-12-369394-5. [Google Scholar] [CrossRef]
  21. Janda, M.; Martišovitš, V.; Hensel, K.; Dvonč, L.; Machala, Z. Measurement of the electron density in Transient Spark discharge. Plasma Sources Sci. Technol. 2014, 23, 065016. [Google Scholar] [CrossRef]
  22. Khedim, S.; Cheknane, A.; Benyoucef, B.; Hilal, H.S. Measurement of neutral gas temperatures in nitrogen-corona discharges. Ind. J. Phys. 2011, 85, 1433–1443. [Google Scholar] [CrossRef]
  23. Ramkumar, M.C.; Cools, P.; Arankumar, A.; de Geyter, N.; Morent, R.; Kumar, V.; Udaykumar, S.; Gopinath, P.; Jaganathan, S.K.; Pandiyaraj, K.N. Chapter 9—Polymer Coatings for Biocompatibility and Reduced Nonspecific Adsorption. In Functionalised Cardiovascular Stents; Woodhead Publishing: Duxford, UK, 2018; ISBN 978-0-08-100496-8. [Google Scholar] [CrossRef]
  24. Xu, X. Dielectric barrier discharge—properties and applications. Thin Solid Films 2001, 390, 237–242. [Google Scholar] [CrossRef]
  25. Lu, X.; Liu, D.; Xian, Y.; Nie, L.; Cau, Y.; He, G. Cold atmospheric-pressure air plasma jet: Physics and opportunities. Phys. Plasmas 2021, 28, 100501. [Google Scholar] [CrossRef]
  26. Li, H.; Xie, M. Plasma parameters of a gliding arc jet at atmospheric pressure obtained by a line-ratio method. Plasma Sci. Techol. 2013, 15, 776. [Google Scholar] [CrossRef]
  27. Zhu, J.J.; Ehn, A.; Gao, J.L.; Kong, C.D.; Aldén, M.; Larsson, A.; Salewski, M.; Leipold, F.; Kusano, Y.; Li, Z.S. Translational, rotational, vibrational and electron temperatures of a gliding arc discharge. Opt. Express. 2017, 25, 20243–20257. [Google Scholar] [CrossRef]
  28. Kusano, Y. Atmospheric pressure plasma processing for polymer adhesion: A review. J. Adhes. 2014, 90, 755–777. [Google Scholar] [CrossRef]
  29. Kogoma, M.; Kusano, M.; Kusano, Y. (Eds.) Generation and Applications of Atmospheric Pressure Plasma; Nova Science Publishers: New York, NY, USA, 2011. [Google Scholar]
  30. Kaplan, S.L.; Rose, P.W. Plasma surface treatment of plastics to enhance adhesion. Int. J. Adhes. Adhes. 1991, 11, 109–113. [Google Scholar] [CrossRef]
  31. Kusano, Y.; Mortensen, H.; Stenum, B.; Goutianos, S.; Mitra, S.; Ghanbari-Siahkali, A.; Kingshott, P.; Sørensen, B.F.; Bingslev, H. Atmospheric pressure plasma treatment of glassy carbon for adhesion improvement. Int. J. Adhes. Adhes. 2007, 27, 402–408. [Google Scholar] [CrossRef]
  32. Kusano, Y.; Mortensen, H.; Stenum, B.; Kingshott, P.; Andersen, T.L.; Brøndsted, P.; Bilde-Sørensen, J.B.; Sørensen, B.F.; Bingslev, H. Atmospheric pressure plasma treatment of glass fibre composite for adhesion improvement. Plasm. Process. Polym. 2007, 4, S455–S459. [Google Scholar] [CrossRef]
  33. Kusano, Y.; Andersen, T.L.; Michelsen, P.K. Atmospheric pressure plasma surface modification of carbon fibres. J. Phys. Conf. Ser. 2008, 100, 012002. [Google Scholar] [CrossRef]
  34. Kusano, Y. Plasma surface modification at atmospheric pressure. Surf. Eng. 2009, 25, 415–416. [Google Scholar] [CrossRef]
  35. Teodoru, S.; Kusano, Y.; Rozlosnik, N.; Michelsen, P.K. Continuous plasma treatment of ultra high molecular weight polyethylene (UHMWPE) fibres for adhesion improvement. Plasm. Process. Polym. 2009, 6, S375–S381. [Google Scholar] [CrossRef]
  36. Leipold, F.; Kusano, Y.; Hansen, F.; Jacobsen, T. Decontamination of a rotating cutting tool during operation by means of atmospheric pressure plasmas. Food Control 2010, 21, 1194–1198. [Google Scholar] [CrossRef]
  37. Kusano, Y.; Singh, S.V.; Bardenshtein, A.; Krebs, N.; Rozlosnik, N. Plasma surface modification of glass fibre reinforced polyester enhanced by ultrasonic irradiation. J. Adhes. Sci. Technol. 2010, 24, 1831–1839. [Google Scholar] [CrossRef]
  38. Kusano, Y.; Teodoru, S.; Hansen, C.M. The physical and chemical properties of plasma treated ultra-high-molecular-weight polyethylene fibers. Surf. Coat. Technol. 2011, 205, 2793–2798. [Google Scholar] [CrossRef]
  39. Kusano, Y.; Singh, S.V.; Norrman, K.; Leipold, F.; Drews, J.; Morgen, P.; Bardenshtein, A.; Krebs, N. Ultrasound enhanced plasma treatment of glass-fibre-reinforced polyester in atmospheric pressure air for adhesion improvement. J. Adhes. 2011, 87, 720–731. [Google Scholar] [CrossRef]
  40. Kusano, Y.; Singh, S.V.; Norrman, K.; Drews, J.; Leipold, F.; Rozlosnik, N.; Bardenshtein, A.; Krebs, N. Ultrasound enhanced plasma surface modification at atmospheric pressure. Surf. Eng. 2012, 28, 453–457. [Google Scholar] [CrossRef]
  41. Siró, I.; Kusano, Y.; Norrman, K.; Goutianos, S.; Plackett, D. Surface modification of nanofibrillated cellulose films by atmospheric pressure dielectric barrier discharge. J. Adhes. Sci. Technol. 2013, 27, 294–308. [Google Scholar] [CrossRef]
  42. Kusano, Y.; Norrman, K.; Singh, S.V.; Leipold, F.; Morgen, P.; Bardenshtein, A.; Krebs, N. Ultrasound enhanced 50 Hz plasma treatment of glass-fiber-reinforced polyester at atmospheric pressure. J. Adhes. Sci. Technol. 2013, 27, 825–833. [Google Scholar] [CrossRef]
  43. Kusano, Y.; Bardenshtein, A.; Morgen, P. Fluoropolymer coated alanine films treated by atmospheric pressure plasmas-in comparison with gamma irradiation. Plasma Process. Polym. 2018, 15, 1700131. [Google Scholar] [CrossRef]
  44. Kusano, Y.; Madsen, B.; Berglund, L.; Aitomäki, Y.; Oksman, K. Dielectric barrier discharge plasma treatment of cellulose nanofibre surfaces. Surf. Eng. 2018, 34, 825–831. [Google Scholar] [CrossRef]
  45. Kusano, Y.; Madsen, B.; Berglund, L.; Oksman, K. Modification of cellulose nanofiber surfaces by He/NH3 plasma at atmospheric pressure. Cellulose 2019, 26, 7185–7194. [Google Scholar] [CrossRef]
  46. Cederløf, D.J.H.; Kusano, Y.; Fæster, S. Fluorination of sized glass fibres for decreased wetting by atmospheric pressure plasma treatment in He/CF4. J. Adhes. 2020, 96, 2–12. [Google Scholar] [CrossRef]
  47. Kusano, Y.; Cederløf, D.J.H.; Fæster, S. Plasma surface modification of glass fibre sizing for manufacturing polymer composites. Key Eng. Mater. 2020, 843, 159–164. [Google Scholar] [CrossRef]
  48. Fang, C.; Cederløf, D.J.H.; Bardenshtein, A.; Kusano, Y. Air-to-air atmospheric pressure plasma treatment-perspective for composite manufacturing. IOP Conf. Ser. Mater. Sci. Eng. 2020, 942, 012030. [Google Scholar] [CrossRef]
  49. Fang, C.; Kusano, Y.; Bardenshtein, A. High-speed plasma treatment of polyethylene terephthalate films using ultrasound assisted dielectric barrier discharge. Packag. Technol. Sci. 2021, 35, 643–649. [Google Scholar] [CrossRef]
  50. Fateev, A.; Leipold, F.; Kusano, Y.; Stenum, B.; Tsakadze, E.; Bindslev, H. Plasma chemistry in an atmospheric pressure Ar/NH3 dielectric barrier discharge. Plasm. Process. Polym. 2005, 2, 193–200. [Google Scholar] [CrossRef]
  51. Kusano, Y.; Leipold, F.; Fateev, A.A.; Stenum, B.; Bindslev, H. Production of ammonia-derived radicals in a dielectric barrier discharge and their injection for denitrification. Surf. Coat. Technol. 2005, 200, 846–849. [Google Scholar] [CrossRef]
  52. Leipold, F.; Fateev, A.; Kusano, Y.; Stenum, B.; Bindslev, H. Reduction of NO in the exhaust gas by reaction with N radicals. Fuel 2006, 85, 1383–1388. [Google Scholar] [CrossRef]
  53. Drews, J.; Kusano, Y.; Leipold, F.; Bardenshtein, A.; Krebs, N. Ozone production in a dielectric barrier discharge with ultrasonic irradiation. Ozone Sci. Eng. 2011, 33, 483–488. [Google Scholar] [CrossRef]
  54. Kusano, Y.; Drews, J.; Leipold, F.; Fateev, A.; Bardenshtein, A.; Krebs, N. Influence of ultrasonic irradiation on ozone generation in a dielectric barrier discharge. J. Phys. Conf. Ser. 2012, 406, 012006. [Google Scholar] [CrossRef]
  55. Babaeva, N.Y.; Naidis, G.V. Modeling of plasmas for biomedicine. Trends Biotechnol. 2017, 36, 603–614. [Google Scholar] [CrossRef] [PubMed]
  56. Klein, A.L.; Yao, X.; Lin, L.; Keidar, M.; Sherman, J. Tumor-treating fields and cold atmospheric plasma treatment: A comparative study in the treatment of glioblastoma. Neurosurgery 2020, 67, 311. [Google Scholar] [CrossRef]
  57. Faramarzi, F.; Zafari, P.; Alimohammadi, M.; Moonesi, M.; Rafiei, A.; Bekeschus, S. Cold physical plasma in cancer therapy: Mechanisms, signaling, and immunity. Oxid. Med. Cell. Longev. 2021, 2021, 9916796. [Google Scholar] [CrossRef] [PubMed]
  58. Siu, A.; Volotskova, O.; Cheng, X.; Khalsa, S.S.; Bian, K.; Murad, F.; Keidar, M.; Sherman, J.H. Differential effects of cold atmospheric plasma in the treatment of malignant glioma. PLoS ONE 2015, 10, e0126313. [Google Scholar] [CrossRef]
  59. Privat-Maldonado, A.; Bogaerts, A. Plasma in cancer treatment. Cancers 2020, 12, 2617. [Google Scholar] [CrossRef]
  60. Jezeh, M.A.; Tayebi, T.; Khani, M.R.; Niknejad, H.; Shokri, B. Direct cold atmospheric plasma and plasma-activated medium effects on breast and cervix cancer cells. Plasma Process. Polym. 2020, 17, 1900241. [Google Scholar] [CrossRef]
  61. Tan, J.K.; Awuah, W.A.; Roy, S.; Ferreira, T.; Ahluwalia, A.; Guggilapu, S.; Javed, M.; Asyura, M.M.A.Z.; Adebusoye, F.T.; Ramamoorthy, K.; et al. Exploring the advances of single-cell RNA sequencing in thyroid cancer: A narrative review. Med. Oncol. 2024, 41, 27. [Google Scholar] [CrossRef] [PubMed]
  62. Fridman, A.; Chirokov, A.; Gutsol, A. Non-thermal atmospheric pressure discharges. J. Phys. D Appl. Phys. 2005, 38, R1. [Google Scholar] [CrossRef]
  63. Zhu, J.; Gao, J.; Li, Z.; Ehn, A.; Aldén, M.; Larsson, A.; Kusano, Y. Sustained diffusive alternating current gliding arc discharge in atmospheric pressure air. Appl. Phys. Lett. 2014, 105, 234102. [Google Scholar] [CrossRef]
  64. Lee, B.J.; Kusano, Y.; Kato, N.; Naito, K.; Horiuchi, T.; Koinuma, H.K.H. Oxygen plasma treatment of rubber surface by the atmospheric pressure cold plasma torch. Jpn. J. Appl. Phys. 1997, 36, 2888–2891. [Google Scholar] [CrossRef]
  65. Mortensen, H.; Kusano, Y.; Leipold, F.; Rozlosnik, N.; Kingshott, P.; Sørensen, B.F.; Stenum, B.; Bindslev, H. Modification of glassy carbon surfaces by an atmospheric pressure cold plasma torch. Jpn. J. Appl. Phys. 2006, 45, 8506–8511. [Google Scholar] [CrossRef]
  66. Kusano, Y.; Teodoru, S.; Leipold, F.; Andersen, T.L.; Sørensen, B.F.; Rozlosnik, N.; Michelsen, P.K. Gliding arc discharge-application for adhesion improvement of fibre reinforced polyester composites. Surf. Coat. Technol. 2008, 202, 5579–5582. [Google Scholar] [CrossRef]
  67. Kusano, Y.; Norrman, K.; Drews, J.; Leipold, F.; Singh, S.V.; Morgen, P.; Bardenshtein, A.; Krebs, N. Gliding arc surface treatment of glass fiber reinforced polyester enhanced by ultrasonic irradiation. Surf. Coat. Technol. 2011, 205, S490–S494. [Google Scholar] [CrossRef]
  68. Kusano, Y.; Sørensen, B.F.; Andersen, T.L.; Leipold, F. Adhesion improvement of glass-fibre-reinforced polyester composites by gliding arc discharge treatment. J. Adhes. 2013, 89, 433–459. [Google Scholar] [CrossRef]
  69. Sun, Z.W.; Zhu, J.J.; Li, Z.S.; Aldén, M.; Leipold, F.; Salewski, M.; Kusano, Y. Optical diagnostics of a gliding arc. Opt. Express. 2013, 21, 6028–6044. [Google Scholar] [CrossRef]
  70. Kusano, Y.; Sørensen, B.F.; Andersen, T.L.; Toftegaard, H.L.; Leipold, F.; Salewski, M.; Sun, Z.W.; Zhu, J.J.; Li, Z.S.; Aldén, M. Water-cooled non-thermal gliding arc for adhesion improvement of glass-fibre-reinforced polyester. J. Phys. D Appl. Phys. 2013, 46, 135203. [Google Scholar] [CrossRef]
  71. Kusano, Y.; Andersen, T.L.; Toftegaard, H.L.; Leipold, F.; Bardenshtein, A.; Krebs, N. Plasma treatment of carbon fibres and glass-fibre-reinforced polyesters at atmospheric pressure for adhesion improvement. Int. J. Mater. Eng. Innov. 2014, 5, 122–137. [Google Scholar] [CrossRef]
  72. Zhu, J.J.; Sun, Z.W.; Li, Z.S.; Ehn, A.; Aldén, M.; Salewski, M.; Leipold, F.; Kusano, Y. Dynamics, OH distributions and UV emission of a gliding arc at various flow-rates investigated by optical measurements. J. Phys. D Appl. Phys. 2014, 47, 295203. [Google Scholar] [CrossRef]
  73. Kusano, Y.; Salewski, M.; Leipold, F.; Zhu, J.J.; Ehn, A.; Li, Z.S.; Aldén, M. Stability of alternating current gliding arcs. Eur. Phys. J. D 2014, 68, 319. [Google Scholar] [CrossRef]
  74. Kusano, Y.; Zhu, J.J.; Ehn, A.; Li, Z.S.; Aldén, M.; Salewski, M.; Leipold, F.; Bardenshtein, A.; Krebs, N. Observation of gliding arc surface treatment. Surf. Eng. 2015, 31, 282–288. [Google Scholar] [CrossRef]
  75. Zhu, J.J.; Gao, J.L.; Ehn, A.; Aldén, M.; Li, Z.S.; Moseev, D.; Kusano, Y.; Salewski, M.; Alpers, A.; Gritzmann, P.; et al. Measurements of 3D slip velocities and plasma column lengths of a gliding arc discharge. Appl. Phys. Lett. 2015, 106, 044101. [Google Scholar] [CrossRef]
  76. Kusano, Y.; Berglund, L.; Aitomäki, Y.; Oksman, K.; Madsen, B. Gliding arc surface modification of carrot nanofibre coating-perspective for composite processing. IOP Conf. Ser. Mater. Sci. Eng. 2016, 139, 012027. [Google Scholar] [CrossRef]
  77. Zhu, J.J.; Gao, J.L.; Ehn, A.; Aldén, M.; Li, Z.S.; Larsson, A.; Kusano, Y. Spatiotemporally resolved characteristics of a gliding arc discharge in a turbulent air flow at atmospheric pressure. Phys. Plasmas 2017, 24, 013514. [Google Scholar] [CrossRef]
  78. Sikarwar, V.S.; Hrabovský, M.; van Oost, G.; Pohořelý, M.; Jeremiáš, M. Progress in waste utilization via thermal plasma. Prog. Energy Combust. Sci. 2020, 81, 100873. [Google Scholar] [CrossRef]
  79. Waugh, R.F.P.; Jardine, M.M.; Morin, J.; Donati, J.F. Slingshot prominences: A hidden mass loss mechanism. Mon. Not. R. Astron. Soc. 2021, 505, 5104–5116. [Google Scholar] [CrossRef]
  80. Cameron, A.C.; Duncan, D.K.; Ehrenfreund, P.; Foing, B.H.; Penston, M.V.; Robinson, R.D.; Soderblom, D.R. Fast spectroscopic variations on rapidly rotating, cool dwarfs - III. Masses of circumstellar absorbing clouds on AB Doradus. Mon. Not. R. Astron. Soc. 1990, 247, 415–438. [Google Scholar]
  81. Gabbar, H.A.; Darda, S.A.; Damideh, V.; Hassen, I.; Aboughaly, M.; Lisi, D. Comparative study of atmospheric pressure DC, RF, and microwave thermal plasma torches for waste to energy applications. Sustain. Energy Technol. Assess. 2021, 47, 101447. [Google Scholar] [CrossRef]
  82. Şanlisoy, A.; Çarpinlioǧlu, O.M. A review on plasma gasification for solid waste disposal. Int. J. Hydrog. Energy 2017, 42, 1361–1365. [Google Scholar] [CrossRef]
  83. Lee, H.G.; Park, H.W.; Choi, S.; Park, H.S.; Park, D.W. Production of synthesis gas from coal by DC non-transferred steam plasma gasification system. J. Chem. Eng. Jpn. 2014, 47, 334–339. [Google Scholar] [CrossRef]
  84. Uhm, H.S.; Na, Y.H.; Hong, Y.C.; Shin, D.H.; Cho, C.H. Production of hydrogen-rich synthetic gas from low-grade coals by microwave steam-plasmas. Int. J. Hydrog. Energy 2014, 39, 4351–4355. [Google Scholar] [CrossRef]
  85. Shie, J.L.; Tsou, F.J.; Lin, K.L.; Chang, C.Y. Bioenergy and products from thermal pyrolysis of rice straw using plasma torch. Bioresour. Technol. 2010, 101, 761–768. [Google Scholar] [CrossRef] [PubMed]
  86. Shie, J.L.; Chen, L.X.; Lin, K.L.; Chang, C.Y. Plasmatron gasification of biomass lignocellulosic waste materials derived from municipal solid waste. Energy 2014, 66, 82–89. [Google Scholar] [CrossRef]
  87. Janajreh, I.; Raza, S.S.; Valmundsson, A.S. Plasma gasification process: Modeling, simulation and comparison with conventional air gasification. Energy Convers. Manag. 2013, 65, 801–809. [Google Scholar] [CrossRef]
  88. Byun, Y.; Namkung, W.; Cho, M.; Chung, J.W.; Kim, Y.S.; Lee, J.H.; Lee, C.R.; Hwang, S.M. Demonstration of thermal plasma gasification/vitrification for municipal solid waste treatment. Environ. Sci. Technol. 2010, 44, 6680–6684. [Google Scholar] [CrossRef]
  89. Byun, Y.; Cho, M.; Chung, J.W.; Namkung, W.; Lee, H.D.; Jang, S.D.; Kim, Y.S.; Lee, J.H.; Lee, C.R.; Hwang, S.M. Hydrogen recovery from the thermal plasma gasification of solid waste. J. Hazard. Mater. 2011, 190, 317–323. [Google Scholar] [CrossRef]
  90. Cui, P.; Xu, Z.; Yao, D.; Qi, H.; Zhu, Z.; Wang, Y.; Li, X.; Liu, Z.; Yang, S. Life cycle water footprint and carbon footprint analysis of municipal sludge plasma gasification process. Energy 2022, 261, 125280. [Google Scholar] [CrossRef]
  91. Danthurebandara, M.; van Passel, S.; Vanderreydt, I.; van Acker, K. Environmental and economic performance of plasma gasification in Enhanced Landfill Mining. Waste Manag. 2015, 45, 458–467. [Google Scholar] [CrossRef]
  92. Georgiev, I.B.; Mihailov, B.I. Some general conclusions from the results of studies on solid fuel steam plasma gasification. Fuel 1991, 71, 895–901. [Google Scholar] [CrossRef]
  93. Jiang, P.; Parvez, A.M.; Meng, Y.; Dong, X.; Xu, M.; Luo, X.; Shi, K.; Wu, T. Novel two-stage fluidized bed-plasma gasification integrated with SOFC and chemical looping combustion for the high efficiency power generation from MSW: A thermodynamic investigation. Energy Convers. Manag. 2021, 236, 114066. [Google Scholar] [CrossRef]
  94. Kalinci, Y.; Hepbasli, A.; Dincer, I. Exergoeconomic analysis of hydrogen production from plasma gasification of sewage sludge using specific exergy cost method. Int. J. Hydrog. Energy 2011, 36, 11408–11417. [Google Scholar] [CrossRef]
  95. Kuo, P.C.; Illathukandy, B.; Wu, W.; Chang, J.S. Energy, exergy, and environmental analyses of renewable hydrogen production through plasma gasification of microalgal biomass. Energy 2021, 223, 120025. [Google Scholar] [CrossRef]
  96. Kwon, S.; Im, S.K. Feasibility of non-thermal plasma gasification for a waste-to-energy power plant. Energy Convers. Manag. 2022, 251, 114978. [Google Scholar] [CrossRef]
  97. Ma, Y.; Qi, H.; Zhang, J.; Cui, P.; Zhu, Z.; Wang, Y. Thermodynamic analysis of a carbon capture hydrogen production process for end-of-life tires using plasma gasification. J. Clean. Prod. 2023, 384, 135662. [Google Scholar] [CrossRef]
  98. Lemmens, B.; Elslander, H.; Vanderreydt, I.; Peys, K.; Diels, L.; Oosterlinck, M.; Joos, M. Assessment of plasma gasification of high caloric waste streams. Waste Manag. 2007, 27, 015003. [Google Scholar] [CrossRef]
  99. Li, J.; Wang, H.; Chen, H.; Wu, H.; Xu, G.; Dong, Y.; Zhao, Q.; Liu, T. Comparative thermodynamic and techno-economic analysis of various medical waste-to-hydrogen/methanol pathways based on plasma gasification. Appl. Therm. Eng. 2023, 221, 119762. [Google Scholar] [CrossRef]
  100. Mazzoni, L.; Almazrouei, M.; Ghenai, C.; Janajreh, I. A comparison of energy recovery from MSW through plasma gasification and entrained flow gasification. Energy Procedia 2017, 142, 3480–3485. [Google Scholar] [CrossRef]
  101. Mazzoni, L.; Janajreh, I. Plasma gasification of municipal solid waste with variable content of plastic solid waste for enhanced energy recovery. Int. J. Hydrog. Energy 2017, 42, 19446–19457. [Google Scholar] [CrossRef]
  102. Mazzoni, L.; Rizwan, A.; Janajreh, I. Plasma gasification of two waste streams: Municipal solid waste and hazardous waste from the oil and gas industry. Energy Procedia 2017, 105, 4159–4166. [Google Scholar] [CrossRef]
  103. Mallick, R.; Vairakannu, P. CO2 plasma gasification of bakelite-based electrical switch waste feedstock. J. Clean. Prod. 2023, 423, 138813. [Google Scholar] [CrossRef]
  104. Mallick, R.; Vairakannu, P. Experimental investigation of acrylonitrile butadiene styrene plastics plasma gasification. J. Environ. Manag. 2023, 345, 118655. [Google Scholar] [CrossRef] [PubMed]
  105. Materazzi, M.; Lettieri, P.; Mazzei, L.; Taylor, R.; Chapman, C. Tar evolution in a two stage fluid bed-plasma gasification process for waste valorization. Fuel Process. Technol. 2014, 128, 146–157. [Google Scholar] [CrossRef]
  106. Materazzi, M.; Lettieri, P.; Mazzei, L.; Taylor, R.; Chapman, C. Fate and behavior of inorganic constituents of RDF in a two stage fluid bed-plasma gasification plant. Fuel 2015, 150, 473–485. [Google Scholar] [CrossRef]
  107. Matveev, I.B.; Messerle, V.E.; Ustimenko, A.B. Plasma gasification of coal in different oxidants. IEEE Trans. Plasma Sci. 2008, 36, 2947–2954. [Google Scholar] [CrossRef]
  108. Messerle, V.E.; Mossé, A.L.; Ustimenko, A.B. Processing of biomedical waste in plasma gasifier. Waste Manag. 2018, 79, 791–799. [Google Scholar] [CrossRef] [PubMed]
  109. Messerle, V.E.; Ustimenko, A.B.; Lavrichshev, O.A. Comparative study of coal plasma gasification: Simulation and experiment. Fuel 2016, 164, 172–179. [Google Scholar] [CrossRef]
  110. Messerle, V.E.; Mossé, A.L.; Ustimenko, A.B. Plasma gasification of carbonaceous wastes: Thermodynamic analysis and experiment. Thermophys. Aeromechanics 2016, 23, 613–620. [Google Scholar] [CrossRef]
  111. Mehrpooya, M.; Hosseini, S.S. A novel integration of plasma gasification melting process with direct carbon fuel cell. Int. J. Hydrog. Energy 2024, 50, 388–401. [Google Scholar] [CrossRef]
  112. Morrin, S.; Lettieri, P.; Chapman, C.; Mazzei, L. Two stage fluid bed-plasma gasification process for solid waste valorisation: Technical review and preliminary thermodynamic modelling of sulphur emissions. Waste Manag. 2012, 32, 676–684. [Google Scholar] [CrossRef]
  113. Paulino, R.F.S.; Essiptchouk, A.M.; Silveira, J.L. The use of syngas from biomedical waste plasma gasification systems for electricity production in internal combustion: Thermodynamic and economic issues. Energy 2020, 199, 117419. [Google Scholar] [CrossRef]
  114. Paulino, R.F.S.; Essiptchouk, A.M.; Costa, L.P.C.; Silveira, J.L. Thermodynamic analysis of biomedical waste plasma gasification. Energy 2022, 244, 122600. [Google Scholar] [CrossRef]
  115. Rojas-Pérez, F.; Castillo-Benavides, J.A.; Richmond-Navarro, G.; Zamora, E. CFD modeling of plasma gasification reactor for municipal solid waste. IEEE Trans. Plasma Sci. 2018, 46, 2435–2444. [Google Scholar] [CrossRef]
  116. Perna, A.; Minutillo, M.; Lavadera, A.L.; Jannelli, E. Combining plasma gasification and solid oxide cell technologies in advanced power plants for waste to energy and electric energy storage applications. Waste Manag. 2018, 73, 424–438. [Google Scholar] [CrossRef] [PubMed]
  117. Qi, H.; Cui, P.; Liu, Z.; Xu, Z.; Yao, D.; Wang, Y.; Zhu, Z.; Yang, S. Conceptual design and comprehensive analysis for novel municipal sludge gasification-based hydrogen production via plasma gasifier. Energy Convers. Manag. 2021, 245, 114635. [Google Scholar] [CrossRef]
  118. Qin, L.; Li, O.L. Recent progress of low-temperature plasma technology in biorefning process. Nano Converg. 2023, 10, 38. [Google Scholar] [CrossRef]
  119. Ramos, A.; Teixeira, C.A.; Rouboa, A. Environmental assessment of municipal solid saste by two-stage plasma gasification. Energies 2018, 12, 137. [Google Scholar] [CrossRef]
  120. Ramos, A.; Berzosa, J.; Espí, J.; Clarens, F.; Rouboa, A. Life cycle costing for plasma gasification of municipal solid waste: A socioeconomic approach. Energy Convers. Manag. 2020, 209, 112508. [Google Scholar] [CrossRef]
  121. Ramos, A.; Rouboa, A. Life cycle thinking of plasma gasification as a waste-to-energy tool: Review on environmental, economic and social aspects. Renew. Sust. Energ. Rev. 2022, 153, 111762. [Google Scholar] [CrossRef]
  122. Rutberg, P.G.; Bratsev, A.N.; Kuznetsov, V.A.; Popov, V.E.; Ufimtsev, A.A.; Shtengel, S.V. On efficiency of plasma gasification of wood residues. Biomass Bioenergy 2011, 35, 495–504. [Google Scholar] [CrossRef]
  123. Sauve, G.; Esguerra, J.L.; Laner, D.; Johansson, J.; Svensson, N.; van Passel, S.; van Acker, K. Integrated early-stage environmental and economic assessment of emerging technologies and its applicability to the case of plasma gasification. J. Clean. Prod. 2023, 382, 134684. [Google Scholar] [CrossRef]
  124. Sturm, G.S.J.; Muñoz, A.N.; Aravind, P.V.; Stefanidis, G.D. Microwave-driven plasma gasification for biomass waste treatment at miniature scale. IEEE Trans. Plasma Sci. 2016, 44, 670–678. [Google Scholar] [CrossRef]
  125. Tamošiūnas, A.; Chouchène, A.; Valatkevicius, P.; Gimžauskaitė, D.; Aikas, M.; Uscila, R.; Ghorbel, M.; Jeguirim, M. The potential of thermal plasma gasification of olive pomace charcoal. Energies 2017, 10, 710. [Google Scholar] [CrossRef]
  126. Tamošiūnas, A.; Gimžauskaitė, D.; Uscila, R.; Aikas, M. Thermal arc plasma gasification of waste glycerol to syngas. Appl. Energy 2019, 251, 113306. [Google Scholar] [CrossRef]
  127. Vecten, S.; Wilkinson, M.; Bimbo, N.; Dawson, R.; Herbert, B.M. Hydrogen-rich syngas production from biomass in a steam microwave-induced plasma gasification reactor. Bioresour. Technol. 2021, 337, 125324. [Google Scholar] [CrossRef]
  128. Yin, K.; Zhang, R.; Yan, M.; Sun, L.; Ma, Y.; Cui, P.; Zhu, Z.; Wang, Y. Thermodynamic and economic analysis of a hydrogen production process from medical waste by plasma gasification. Process Saf. Environ. Prot. 2023, 178, 8–17. [Google Scholar] [CrossRef]
  129. Yoon, S.J.; Yun, Y.M.; Seo, M.W.; Kim, Y.K.; Ra, H.W.; Lee, J.G. Hydrogen and syngas production from glycerol through microwave plasma gasification. Int. J. Hydrog. Energy 2013, 38, 14559–14567. [Google Scholar] [CrossRef]
  130. Zhang, Q.; Dor, L.; Fenigshtein, D.; Yang, W.; Blasiak, W. Gasification of municipal solid waste in the Plasma Gasification Melting process. Appl. Energy 2012, 90, 106–112. [Google Scholar] [CrossRef]
  131. Zhang, Q.; Dor, L.; Zhang, L.; Yang, W.; Blasiak, W. Performance analysis of municipal solid waste gasification with steam in a Plasma Gasification Melting reactor. Appl. Energy 2012, 98, 219–229. [Google Scholar] [CrossRef]
  132. Zhang, Q.; Wu, Y.; Dor, L.; Yang, W.; Blasiak, W. A thermodynamic analysis of solid waste gasification in the Plasma Gasification Melting process. Appl. Energy 2013, 112, 405–413. [Google Scholar] [CrossRef]
  133. Zhang, Q.; Dor, L.; Biswas, A.K.; Yang, W.; Blasiak, W. Modeling of steam plasma gasification for municipal solid waste. Fuel Process. Technol. 2013, 106, 546–554. [Google Scholar] [CrossRef]
  134. Fabry, F.; Rehmet, C.; Rohani, V.; Fulcheri, L. Waste gasification by thermal plasma: A review. Waste Biomass Valorization 2013, 4, 421–439. [Google Scholar] [CrossRef]
  135. Oliveira, M.; Ramos, A.; Ismail, T.M.; Monteiro, E.; Rouboa, A. A review on plasma gasification of solid residues: Recent advances and developments. Energies 2022, 15, 1475. [Google Scholar] [CrossRef]
  136. Achinas, S.; Kapetanios, E. Efficiency evaluation of RDF plasma gasification process. Energy Environ. Res. 2013, 3, 150–157. [Google Scholar] [CrossRef]
  137. Agon, N.; Hrabovský, M.; Chumak, O.M.; Hlína, M.; Kopecký, V.; Mašláni, A.; Bosmans, A.; Helsen, L.M.; Skoblia, S.; van Oost, G.; et al. Plasma gasification of refuse derived fuel in a single-stage system using different gasifying agents. Waste Manag. 2015, 47, 246–255. [Google Scholar] [CrossRef] [PubMed]
  138. Montiel-Bohórquez, N.D.; Agudelo, A.F.; Pérez, J.F. Effect of origin and production rate of MSW on the exergoeconomic performance of an integrated plasma gasification combined cycle power plant. Energy Convers. Manag. 2021, 238, 114138. [Google Scholar] [CrossRef]
  139. Montiel-Bohórquez, N.D.; Saldarriaga-Loaiza, J.D.; Pérez, J.F. Analysis of investment incentives for power generation based on an integrated plasma gasification combined cycle power plant using municipal solid waste. Case Stud. Therm. Eng. 2022, 30, 101748. [Google Scholar] [CrossRef]
  140. Cudjoe, D.; Wang, H. Plasma gasification versus incineration of plastic waste: Energy, economic and environmental analysis. Fuel Process. Technol. 2022, 237, 107470. [Google Scholar] [CrossRef]
  141. Chen, H.; Li, J.; Li, T.; Xu, G.; Jin, X.; Wang, M.; Liu, T. Performance assessment of a novel medical-waste-to-energy design based on plasma gasification and integrated with a municipal solid waste incineration plant. Energy 2022, 254, 123156. [Google Scholar] [CrossRef]
  142. Chu, C.; Boré, A.; Liu, X.W.; Cui, J.; Wang, P.; Liu, X.; Chen, G.; Liu, B.; Ma, W.; Lou, Z.; et al. Modeling the impact of some independent parameters on the syngas characteristics during plasma gasification of municipal solid waste using artificial neural network and stepwise linear regression methods. Renew. Sust. Energ. Rev. 2022, 157, 112052. [Google Scholar] [CrossRef]
  143. Mountouris, A.; Voutsas, E.C.; Tassios, D.P. Solid waste plasma gasification: Equilibrium model development and exergy analysis. Energy Convers. Manag. 2006, 47, 1723–1737. [Google Scholar] [CrossRef]
  144. Mountouris, A.; Voutsas, E.C.; Tassios, D.P. Plasma gasification of sewage sludge: Process development and energy optimization. Energy Convers. Manag. 2008, 49, 2264–2271. [Google Scholar] [CrossRef]
  145. Moustakas, K.G.; Fatta-Kassinos, D.; Xydis, G.A.; Malamis, S.; Haralambous, K.J.; Loizidou, M.D. Demonstration plasma gasification/vitrification system for effective hazardous waste treatment. J. Hazard. Mater. 2005, 123, 120–126. [Google Scholar] [CrossRef] [PubMed]
  146. Moustakas, K.G.; Xydis, G.A.; Malamis, S.; Haralambous, K.J.; Loizidou, M.D. Analysis of results from the operation of a pilot plasma gasification/vitrification unit for optimizing its performance. J. Hazard. Mater. 2008, 151, 473–480. [Google Scholar] [CrossRef] [PubMed]
  147. Galeno, G.; Minutillo, M.; Perna, A. From waste to electricity through integrated plasma gasification/fuel cell (IPGFC) system. Int. J. Hydrog. Energy 2011, 36, 1692–1701. [Google Scholar] [CrossRef]
  148. Sanjaya, E.; Abbas, A. Plasma gasification as an alternative energy-from-waste (EFW) technology for the circular economy: An environmental review. Resour. Conserv. Recycl. 2023, 189, 106730. [Google Scholar] [CrossRef]
  149. Homel, M.; Gür, T.M.; Koh, J.H.; Virkar, A.V. Carbon monoxide-fueled solid oxide fuel cell. J. Power Sources 2010, 194, 6367–6372. [Google Scholar] [CrossRef]
  150. Trimm, D.L. Minimisation of carbon monoxide in a hydrogen stream for fuel cell application. Appl. Catal. A Gen. 2005, 296, 1–11. [Google Scholar] [CrossRef]
  151. Favas, J.; Monteiro, E.; Rouboa, A. Hydrogen production using plasma gasification with steam injection. Int. J. Hydrog. Energy 2017, 42, 10997–11005. [Google Scholar] [CrossRef]
  152. Bhatt, K.P.; Patel, S.; Upadhyay, D.S.; Patel, R.N. A critical review on solid waste treatment using plasma pyrolysis technology. Chem. Eng. Process. 2022, 177, 108989. [Google Scholar] [CrossRef]
  153. Huang, H.; Tang, L. Treatment of organic waste using thermal plasma pyrolysis technology. Energy Convers. Manag. 2007, 48, 1331–1337. [Google Scholar] [CrossRef]
  154. Lei, F.; Li, X.; Liu, Y.; Liu, D.; Yang, M.; Yu, Y. Simulation of a large size inductively coupled plasma generator and comparison with experimental data. AIP Adv. 2018, 8, 015003. [Google Scholar] [CrossRef]
  155. Erdogan, A.A.; Yilmazoglu, M.Z. Plasma gasification of the medical waste. Int. J. Hydrog. Energy 2021, 46, 29108–29125. [Google Scholar] [CrossRef] [PubMed]
  156. Erdogan, A.A.; Yilmazoglu, M.Z. Experimental and numerical investigation of medical waste disposal via plasma gasification. Appl. Energy 2024, 353, 122014. [Google Scholar] [CrossRef]
  157. Indrawan, N.; Mohammed, S.; Kumar, A.; Huhnke, R.L. Modeling low temperature plasma gasification of municipal solid waste. Environ. Technol. Innov. 2019, 15, 100412. [Google Scholar] [CrossRef]
  158. Diaz-Silvarrey, L.S.; Zhang, K.; Phan, A.N. Monomer recovery through advanced pyrolysis of waste high density polyethylene (HDPE). Green Chem. 2018, 20, 1813–1823. [Google Scholar] [CrossRef]
  159. Galvita, V.; Messerle, V.E.; Ustimenko, A.B. Hydrogen production by coal plasma gasification for fuel cell technology. Int. J. Hydrog. Energy 2007, 32, 3899–3906. [Google Scholar] [CrossRef]
  160. Hlína, M.; Hrabovsky, M.; Kopecký, V.; Konrad, M.; Kavka, T.; Skoblia, S. Plasma gasification of wood and production of gas with low content of tar. Czechoslov. J. Phys. 2006, 56, B1179–B1184. [Google Scholar] [CrossRef]
  161. Ismail, T.M.; Monteiro, E.; Ramos, A.; El-Salam, M.A.; Rouboa, A. An Eulerian model for forest residues gasification in a plasma gasifier. Energy 2019, 182, 1069–1083. [Google Scholar] [CrossRef]
  162. Munir, M.T.; Mardon, I.; Al-Zuhair, S.; Shawabkeh, A.; Saqib, N.U. Plasma gasification of municipal solid waste for waste-to-value processing. Renew. Sust. Energ. Rev. 2019, 116, 109461. [Google Scholar] [CrossRef]
  163. Ibarra, R.M. Chapter 20—Recycling of thermosets and their composites. In Thermosets: Structure, Properties, and Applications; Woodhead Publishing: Philadelphia, PA, USA, 2018; ISBN 978-0-08-101021-1. [Google Scholar] [CrossRef]
  164. Ibrahimoglu, B.; Cucen, A.; Yilmazoglu, M.Z. Numerical modeling of a downdraft plasma gasification reactor. Int. J. Hydrog. Energy 2017, 42, 2583–2591. [Google Scholar] [CrossRef]
  165. Minutillo, M.; Perna, A.; Di Bona, D. Modelling and performance analysis of an integrated plasma gasification combined cycle (IPGCC) power plant. Energy Convers. Manag. 2009, 50, 2837–2842. [Google Scholar] [CrossRef]
  166. Okati, A.; Khani, M.R.; Shokri, B.; Monteiro, E.; Rouboa, A. Numerical modeling of plasma gasification process of polychlorinated biphenyl wastes. Energy Rep. 2021, 7, 270–285. [Google Scholar] [CrossRef]
  167. Zitouni-Petrogianni, A.; Voutsas, E. Modeling, optimization and cost analysis of municipal solid waste treatment with plasma gasifcation. Environ. Process. 2021, 8, 747–767. [Google Scholar] [CrossRef]
  168. Sakhraji, M.; Ramos, A.; Monteiro, E.; Bouziane, K.; Rouboa, A. Plasma gasification process using computational fluid dynamics modeling. Energy Rep. 2022, 8, 1541–1549. [Google Scholar] [CrossRef]
  169. Tavares, R.; Ramos, A.; Rouboa, A. A theoretical study on municipal solid waste plasma gasification. Waste Manag. 2019, 90, 37–45. [Google Scholar] [CrossRef]
  170. Zhang, Q.; Dor, L.; Yang, W.; Blasiak, W. Eulerian model for municipal solid waste gasification in a fixed-bed plasma gasification melting reactor. Energy Fuels 2011, 25, 4129–4137. [Google Scholar] [CrossRef]
  171. Samal, S.; Blanco, I. An overview of thermal plasma arc systems for treatment of various wastes in recovery of metals. Materials 2022, 15, 683. [Google Scholar] [CrossRef]
  172. Bai, L.; Sun, W.; Yang, Z.; Ouyang, Y.; Wang, M.; Yuan, F. Laboratory research on design of three-phase AC arc plasma pyrolysis device for recycling of waste printed circuit boards. Processes 2022, 10, 1031. [Google Scholar] [CrossRef]
  173. Szałatkiewicz, J. Metals recovery from waste of printed circuit boards processed in plasmatron plasma reactor. IFAC Proc. Vol. 2013, 46, 478–483. [Google Scholar] [CrossRef]
  174. Mitrasinovic, A.; Pershin, L.; Wen, J.Z.; Mostaghimi, J. Recovery of Cu and valuable metals from E-waste using thermal plasma treatment. JOM 2011, 63, 24–28. [Google Scholar] [CrossRef]
  175. Changming, D.; Chao, S.; Gong, X.; Ting, W.; Xiange, W. Plasma methods for metals recovery from metal-containing waste. Waste Manag. 2018, 77, 373–387. [Google Scholar] [CrossRef]
  176. Wang, S.B.; Cheng, C.M.; Lan, W.; Zhang, X.H.; Liu, D.P.; Yang, S.Z. Experimental study of thermal plasma processing waste circuit boards. Adv. Mater. Res. 2013, 652–654, 1553–1561. [Google Scholar] [CrossRef]
  177. Rath, S.S.; Nayak, P.; Mukherjee, P.S.; Chaudhury, G.R.; Mishra, B.K. Treatment of electronic waste to recover metal values using thermal plasma coupled with acid leaching—A response surface modeling approach. Waste Manag. 2012, 32, 575–583. [Google Scholar] [CrossRef]
  178. Rath, S.S.; Jayasankar, K.; Bijoy, K.S.; Mishra, B.K.; Mukherjee, P.S. Kinetics and statistical behaviour of iron recovery from red mud using plasma arc furnace. High Temp. Mater. Process. 2011, 30, 211–215. [Google Scholar] [CrossRef]
  179. Tang, W.; Khavarian, M.; Yousefi, A. Chapter 14—Red Mud. In Sustainable Concrete Made with Ashes and Dust from Different Sources; Woodhead Publishing: Kidlington, UK, 2022; ISBN 978-0-12-824050-2. [Google Scholar] [CrossRef]
  180. Chou, I.C.; Kuo, Y.M.; Lin, C.; Wang, J.W.; Wang, C.T.; Chang-Chien, G.P. Electroplating sludge metal recovering with vitrification using mineral powder additive. Resour. Conserv. Recycl. 2012, 58, 45–49. [Google Scholar] [CrossRef]
  181. Kabtamu, D.M.; Wu, Y.n.; Chen, Q.; Zheng, L.; Otake, K.i.; Matović, L.; Li, F. Facile upcycling of hazardous Cr-containing electroplating sludge into value-added metal-organic frameworks (MOFs) for efficient adsorptive desulfurization. ACS Sustain. Chem. Eng. 2020, 8, 12443–12452. [Google Scholar] [CrossRef]
  182. Ramachandran, K.; Kikukawa, N. Plasma in-flight treatment of electroplating sludge. Vacuum 2000, 59, 244–251. [Google Scholar] [CrossRef]
  183. Lin, W.C.; Tsai, C.H.; Zhang, D.N.; Syu, S.S.; Kuo, Y.M. Recycling of aluminum dross for producing calcinated alumina by microwave plasma. Sustain. Environ. Res. 2022, 32, 50. [Google Scholar] [CrossRef]
  184. Drouet, M.G.; Meunier, J.; Laflamme, C.B.; Handfield, M.D.; Biscaro, A.; Lemire, C. A rotary arc furnace for aluminum dross processing. In Proceedings of the International Symposium on Environmental Technologies: Plasma Systems and Applications, Atlanta, GA, USA, 8–11 October 1995; Mayne, P.W., Mulholland, J.A., Eds.; Georgia Institute of Technology: Atlanta, GA, USA, 1995; Volume 2, pp. 539–548. [Google Scholar]
  185. McPherson, R.; Shafer, B.V. Spherulites and phase separation in plasma-dissociated zircon. J. Mater. Sci. 1984, 19, 2696–2704. [Google Scholar] [CrossRef]
  186. Pavlik, R.S., Jr.; Holland, H.J. Thermal decomposition of zircon refractories. J. Am. Ceram. Soc. 2001, 84, 2930–2936. [Google Scholar] [CrossRef]
  187. Yugeswaran, S.; Ananthapadmanabhan, P.V.; Lusvarghi, L. Zircon dissociation in air plasma through a low power transferred arc plasma torch. Ceram. Int. 2015, 41, 265–273. [Google Scholar] [CrossRef]
  188. Assessing Zircon’s Safety for Nuclear Waste Storage. Available online: https://www.gla.ac.uk/schools/ges/research/researchfacilities/isaac/casestudies/zircon/ (accessed on 2 January 2024).
  189. Montel, J.M. Minerals and design of new waste forms for conditioning nuclear waste-Les minéraux et la formulation de nouvelles matrices de stockage pour les déchets radioactifs. Comptes Rendus Geosci. 2011, 343, 230–236. [Google Scholar] [CrossRef]
  190. Guillon, O.; Gonzalez-Julian, J.; Dargatz, B.; Kessel, T.; Schierning, G.; Räthel, J.; Herrmann, M. Field-assisted sintering technology/spark plasma sintering: Mechanisms, materials, and technology developments. Adv. Eng. Mater. 2014, 16, 830–849. [Google Scholar] [CrossRef]
  191. Yucheng, W.; Zhengyi, F. Study of temperature field in spark plasma sintering. Mater. Sci. Eng. B 2002, 90, 34–37. [Google Scholar] [CrossRef]
  192. Gu, Y.W.; Loh, N.H.; Khor, K.A.; Tor, S.B.; Cheang, P. Spark plasma sintering of hydroxyapatite powders. Biomaterials 2002, 23, 37–43. [Google Scholar] [CrossRef]
  193. Chuvildeev, V.N.; Panov, D.V.; Boldin, M.S.; Nokhrin, A.V.; Blagoveshchensky, Y.V.; Sakharov, N.V.; Shotin, S.V.; Kotkov, D.N. Structure and properties of advanced materials obtained by Spark Plasma Sintering. Acta Astronaut. 2015, 109, 172–176. [Google Scholar] [CrossRef]
  194. Munir, Z.A.; Anselmi-Tamburini, U.; Ohyanagi, M. The effect of electric field and pressure on the synthesis and consolidation of materials: A review of the spark plasma sintering method. J. Mater. Sci. 2006, 41, 763–777. [Google Scholar] [CrossRef]
  195. Olevsky, E.A.; Bradbury, W.L.; Haines, C.D.; Martin, D.G.; Kapoor, D. Fundamental aspects of spark plasma sintering: I. Experimental analysis of scalability. J. Am. Ceram. Soc. 2012, 95, 2406–2413. [Google Scholar] [CrossRef]
  196. Song, S.X.; Wang, Z.; Shi, G.P. Heating mechanism of spark plasma sintering. Ceram. Int. 2013, 39, 1393–1396. [Google Scholar] [CrossRef]
  197. Hulbert, D.M.; Anders, A.; Andersson, J.; Lavernia, E.J.; Mukherjee, A.K. A discussion on the absence of plasma in spark plasma sintering. Scr. Mater. 2009, 60, 835–838. [Google Scholar] [CrossRef]
  198. Olevsky, E.A.; Garcia-Cardona, C.; Bradbury, W.L.; Haines, C.D.; Martin, D.G.; Kapoor, D. Fundamental aspects of spark plasma sintering: II. Finite element analysis of scalability. J. Am. Ceram. Soc. 2012, 95, 2414–2422. [Google Scholar] [CrossRef]
  199. Guo, S.Q.; Nishimura, T.; Kagawa, Y.; Yang, J.M. Spark plasma sintering of zirconium diborides. J. Am. Ceram. Soc. 2008, 91, 2848–2855. [Google Scholar] [CrossRef]
  200. Anselmi-Tamburini, U.; Garay, J.E.; Munir, Z.A. Fundamental investigations on the spark plasma sintering/synthesis process: III. Current effect on reactivity. Mater. Sci. Eng. A 2005, 407, 24–30. [Google Scholar] [CrossRef]
  201. Mamedov, V. Spark plasma sintering as advanced PM sintering method. Powder Metall. 2002, 45, 322–328. [Google Scholar] [CrossRef]
  202. Shen, Z.; Johnsson, M.; Zhao, Z.; Nygren, M. Spark plasma sintering of alumina. J. Am. Ceram. Soc. 2004, 85, 1921–1927. [Google Scholar] [CrossRef]
  203. Dudina, D.V.; Bokhonov, B.B.; Olevsky, E.A. Fabrication of porous materials by spark plasma sintering: A review. Materials 2019, 12, 541. [Google Scholar] [CrossRef]
  204. Hu, Z.Y.; Zhang, Z.H.; Cheng, X.W.; Wang, F.C.; Zhang, Y.F.; Li, S.L. A review of multi-physical fields induced phenomena and effects in spark plasma sintering: Fundamentals and applications. Mater. Des. 2020, 191, 108662. [Google Scholar] [CrossRef]
  205. Weintraub, G.; Rush, H. Improvements in and Relating to the Treatment of Refractory Materials and Apparatus Therefor. UK Patent 191,220,348 A, 8 December 1913. [Google Scholar]
  206. Biesuz, M.; Grasso, S.; Sglavo, V.M. What’s new in ceramics sintering? A short report on the latest trends and future prospects. Curr. Opin. Solid State Mater. Sci. 2020, 24, 100868. [Google Scholar] [CrossRef]
  207. Paraskevas, D.; Vanmeensel, K.; Vleugels, J.; Dewulf, W.; Deng, Y.; Duflou, J.R. Spark plasma sintering as a solid-state recycling technique: The case of aluminum alloy scrap consolidation. Materials 2014, 7, 5664–5687. [Google Scholar] [CrossRef]
  208. Vukšić, M.; Žmak, I.; Ćurković, L.; Kocjan, A. Spark plasma sintering of dense alumina ceramics from industrial waste scraps. Open Ceram. 2021, 5, 100076. [Google Scholar] [CrossRef]
  209. Kanda, Y. Spark plasma sintering using calcareous waste concrete powder. Constr. Build. Mater. 2022, 349, 128726. [Google Scholar] [CrossRef]
  210. Panasenko, A.E.; Shichalin, O.O.; Yarusova, S.B.; Ivanets, A.I.; Belov, A.A.; Dran’kov, A.N.; Azon, S.A.; Fedorets, A.N.; Buravlev, I.Y.; Mayorov, V.Y.; et al. A novel approach for rice straw agricultural waste utilization: Synthesis of solid aluminosilicate matrices for cesium immobilization. Nucl. Eng. Technol. 2022, 54, 3250–3259. [Google Scholar] [CrossRef]
  211. Shichalin, O.O.; Yarusova, S.B.; Ivanov, N.P.; Papynov, E.K.; Belov, A.A.; Azon, S.A.; Buravlev, I.Y.; Myagchilov, A.V.; Fedorets, A.N.; Rastorguev, V.L.; et al. Calcium silicate solid-state matrices from boric acid production waste for 60Co removal and immobilization by spark plasma sintering. J. Water Process Eng. 2024, 59, 105042. [Google Scholar] [CrossRef]
  212. Teodoru, S.; Kusano, Y.; Bogaerts, A. The effect of O2 in a humid O2/N2/NOx gas mixture on NOx and N2O remediation by an atmospheric pressure dielectric barrier discharge. Plasma Process. Polym. 2012, 9, 652–689. [Google Scholar] [CrossRef]
  213. Abiev, R.S.; Sladkovskiy, D.A.; Semikin, K.V.; Murzin, D.Y.; Rebrov, E.V. Non-thermal plasma for process and energy intensification in dry reforming of methane. Catalysts 2020, 10, 1358. [Google Scholar] [CrossRef]
  214. Budhraja, N.; Pal, A.; Mishra, R. Plasma reforming for hydrogen production: Pathways, reactors and storage. Int. J. Hydrog. Energy 2023, 48, 2467–2482. [Google Scholar] [CrossRef]
  215. Lausche, A.C.; Schaidle, J.A.; Schweitzer, N.M.; Thompson, L.T. Chapter 14. Nanoscale Carbide and Nitride Catalysts. In Comprehensive Inorganic Chemistry II: From Elements to Applications-Volume 7: Surface Inorganic Chemistry and Heterogeneous Catalysis; Elsevier: Waltham, MA, USA, 2013; ISBN 978-0-08-096529-1. [Google Scholar] [CrossRef]
  216. Aminu, I.; Nahil, M.A.; Williams, P.T. Hydrogen from waste plastics by two-stage pyrolysis/low temperature plasma catalytic processing. Energy Fuels. 2020, 34, 11679–11689. [Google Scholar] [CrossRef]
  217. Aminu, I.; Nahil, M.A.; Williams, P.T. Hydrogen production by pyrolysis-nonthermal plasma/catalytic reforming of waste plastic over different catalyst support materials. Energy Fuels 2022, 36, 3788–3801. [Google Scholar] [CrossRef]
  218. Blanquet, E.; Nahil, M.A.; Williams, P.T. Enhanced hydrogen-rich gas production from waste biomass using pyrolysis with non-thermal plasma-catalysis. Catal. Today 2019, 337, 216–224. [Google Scholar] [CrossRef]
  219. Blanquet, E.; Williams, P.T. Biomass pyrolysis coupled with non-thermal plasma/catalysis for hydrogen production: Influence of biomass components and catalyst properties. J. Anal. Appl. Pyrolysis. 2021, 159, 105325. [Google Scholar] [CrossRef]
  220. Li, X.S.; Zhu, A.M.; Wang, K.J.; Xu, Y.; Song, Z.M. Methane conversion to C2 hydrocarbons and hydrogen in atmospheric non-thermal plasmas generated by different electric discharge techniques. Catal. Today 2004, 98, 617–624. [Google Scholar] [CrossRef]
  221. Raja, R.B.; Sarathi, R.; Vinu, R. Selective production of hydrogen and solid carbon via methane pyrolysis using a swirl-induced point-plane non-thermal plasma reactor. Energy Fuels 2022, 36, 826–836. [Google Scholar] [CrossRef]
  222. Lee, D.H.; Song, Y.H.; Kim, K.T.; Lee, J.O. Comparative study of methane activation process by different plasma sources. Plasma Chem. Plasma Process. 2013, 33, 647–661. [Google Scholar] [CrossRef]
  223. Nozaki, T.; Okazaki, K. Non-thermal plasma catalysis of methane: Principles, energy efficiency, and applications. Catal. Today 2013, 211, 29–38. [Google Scholar] [CrossRef]
  224. Li, X.; Zhang, H.; Shao, S.; Lv, Z.; Ge, S.; Cai, Y. Direct non-thermal plasma regeneration of deactivated HZSM-5 for catalytic pyrolysis of rape straw. J. Anal. Appl. Pyrolysis. 2021, 157, 105209. [Google Scholar] [CrossRef]
  225. Li, X.; Yan, J.; Shao, S.; Ge, S.; Lv, Z.; Cai, Y. Synthetic HZSM-5/MCM-41 catalyst by Direct Non-thermal Plasma for promoted production of aromatic hydrocarbons by catalytic pyrolysis of rape straw. J. Anal. Appl. Pyrolysis. 2022, 161, 105365. [Google Scholar] [CrossRef]
  226. Cai, X.; Wei, X.; Wu, J.; Ding, J.; Du, C. Plasma pyrolysis and gasifcation of carambola leaves using non-thermal arc plasma. Waste Dispos. Sustain. Energy 2020, 2, 193–207. [Google Scholar] [CrossRef]
  227. Lim, M.T.; Tan, E.S.; Chai, Y.H.; Chin, B.L.F.; Juwono, F.H.; Hisham, D.H.B.; Anuar, M.A.M. Pilot-scale flue gas pyrolysis system for organic and plastic wastes with improved liquid properties in a non-thermal plasma reactor. J. Anal. Appl. Pyrolysis. 2023, 173, 106062. [Google Scholar] [CrossRef]
  228. Sun, D.L.; Wang, F.; Ruoyu, H.; Xie, C.R. Preparation of carbon black via arc discharge plasma enhanced by thermal pyrolysis. Diam. Relat. Mater. 2016, 61, 21–31. [Google Scholar] [CrossRef]
  229. Titov, E.Y.; Bodrikov, I.V.; Vasiliev, A.L.; Kurskii, Y.A.; Ivanova, A.G.; Golovin, A.L.; Shirokov, D.A.; Titov, D.Y.; Bodrikova, E.R. Non-thermal plasma pyrolysis of fuel oil in the liquid phase. Energies 2023, 16, 4017. [Google Scholar] [CrossRef]
  230. Fan, Y.; Zhao, W.; Shao, S.; Cai, Y.; Chen, Y.; Jin, L. Promotion of the vapors from biomass vacuum pyrolysis for biofuels under Non-thermal Plasma Synergistic Catalysis (NPSC) system. Energy 2018, 142, 462–472. [Google Scholar] [CrossRef]
  231. Saleem, F.; Harris, J.; Zhang, K.; Harvey, A. Non-thermal plasma as a promising route for the removal of tar from the product gas of biomass gasification—A critical review. Chem. Eng. J. 2020, 382, 122761. [Google Scholar] [CrossRef]
  232. Fan, Y.; Lu, D.; Han, Y.; Yang, J.; Qian, C.; Li, B. Production of light aromatics from biomass components co-pyrolyzed with polyethylene via non-thermal plasma synergistic upgrading. Energy 2023, 265, 126427. [Google Scholar] [CrossRef]
  233. Ismail, M.M.; Dincer, I. Development and evaluation of an integrated waste to energy system based on polyethylene plastic wastes pyrolysis for production of hydrogen fuel and other useful commodities. Fuel 2023, 334, 126409. [Google Scholar] [CrossRef]
  234. Snoeckx, R.; Rabinovich, A.; Dobrynin, D.; Bogaerts, A.; Fridman, A. Integrated early-stage environmental and economic assessment of emerging technologies and its applicability to the case of plasma gasification. Plasma Process Polym. 2016, 14, 1600115. [Google Scholar] [CrossRef]
  235. Mishnaevsky, L.; Branner, K.; Petersen, H.N.; Beauson, J.; McGugan, M.; Sørensen, B.F. Materials for wind turbine blades: An overview. Materials 2017, 10, 1285. [Google Scholar] [CrossRef]
  236. Anguita, J.V.; Smith, C.T.G.; Stute, T.; Funke, M.; Delkowski, M.; Silva, S.R.P. Dimensionally and environmentally ultra-stable polymer composites reinforced with carbon fibres. Nat. Mater. 2020, 19, 317–322. [Google Scholar] [CrossRef] [PubMed]
  237. Vieira, D.R.; Vieira, R.K.; Chain, M.C. Strategy and management for the recycling of carbon fiber-reinforced polymers (CFRPs) in the aircraft industry: A critical review. Int. J. Sustain. Dev. World Ecol. 2017, 24, 214–223. [Google Scholar] [CrossRef]
  238. Zhang, L.W.; Sojobi, A.O.; Liew, K.M. Sustainable CFRP-reinforced recycled concrete for cleaner eco-friendly construction. J. Clean. Prod. 2019, 233, 56–75. [Google Scholar] [CrossRef]
  239. Caggegi, C.; Pensée, V.; Fagone, M.; Cuomo, M.; Chevalier, L. Experimental global analysis of the efficiency of carbon fiber anchors applied over CFRP strengthened bricks. Constr. Build. Mater. 2014, 53, 203–212. [Google Scholar] [CrossRef]
  240. Chin, S.C.; Shafiq, N.; Nuruddin, M.F. Behaviour of RC beams with CFRP-strengthened openings. Struct. Concr. 2015, 17, 32–43. [Google Scholar] [CrossRef]
  241. Gonçalves, R.M.; Martinho, A.; Oliveira, J.P. Recycling of reinforced glass fibers waste: Current status. Materials 2022, 15, 1596. [Google Scholar] [CrossRef] [PubMed]
  242. Qureshi, J. A review of recycling methods for fibre reinforced polymer composites. Sustainability 2022, 14, 16855. [Google Scholar] [CrossRef]
  243. Thomas, L.; Ramachandra, M. Advanced materials for wind turbine blade - A review. Mater. Today Proc. 2018, 5 Pt 3, 2635–2640. [Google Scholar] [CrossRef]
  244. Jensen, J.P.; Skelton, K. Wind turbine blade recycling: Experiences, challenges and possibilities in a circular economy. Renew. Sust. Energ. Rev. 2018, 97, 165–176. [Google Scholar] [CrossRef]
  245. Sastri, V.R. Chapter 3—Materials Used in Medical Devices. In Plastics in Medical Devices: Properties, Requirements, and Applications; Elsevier: Waltham, MA, USA, 2022; ISBN 978-0-323-85126-8. [Google Scholar] [CrossRef]
  246. de Fazio, D.; Boccarusso, L.; Formisano, A.; Viscusi, A.; Durante, M. A review on the recycling technologies of fibre-reinforced plastic (FRP) materials used in industrial fields. J. Mar. Sci. Eng. 2023, 11, 851. [Google Scholar] [CrossRef]
  247. Khalil, Y.F. Comparative environmental and human health evaluations of thermolysis and solvolysis recycling technologies of carbon fiber reinforced polymer waste. Waste Manag. 2018, 76, 767–778. [Google Scholar] [CrossRef] [PubMed]
  248. Mamanpush, S.H.; Li, H.; Englund, K.; Tavousi Tabatabaei, A. Recycled wind turbine blades as a feedstock for second generation composites. Waste Manag. 2018, 76, 708–714. [Google Scholar] [CrossRef] [PubMed]
  249. Rybicka, J.; Tiwari, A.; Leeke, G.A. Technology readiness level assessment of composites recycling technologies. J. Clean. Prod. 2016, 112, 1001–1012. [Google Scholar] [CrossRef]
  250. Zhang, L.; Liu, W.; Jiang, H.; Zhang, X.; Shang, Y.; Jiang, C.; Wang, X.; Qi, G.; Li, B.; Xu, P.; et al. Upcycling of carbon fiber-reinforced polymer composites. Compos. Sci. Technol. 2023, 231, 109824. [Google Scholar] [CrossRef]
  251. Pei, C.; Chen, P.Y.; Kong, S.C.; Wu, J.; Zhu, J.H.; Xing, F. Recyclable separation and recovery of carbon fibers from CFRP composites: Optimization and mechanism. Sep. Purif. Technol. 2021, 278, 119591. [Google Scholar] [CrossRef]
  252. Kavaliauskas, Z.; Kėželis, R.; Grigaitienė, V.; Milieška, M.; Valincius, V.; Uscila, R.; Snapkauskienė, V.; Gimžauskaitė, D.; Baltušnikas, A. Recycling of wind turbine blades into microfiber using plasma technology. Materials 2023, 16, 3089. [Google Scholar] [CrossRef] [PubMed]
  253. Grigaitienė, V.; Uscila, R.; Valinčius, V.; Kėželis, R.; Milieška, M.; Gimžauskaitė, D.; Snapkauskienė, V.; Kavaliauskas, Z. Possibility of recycling wind turbine blades using plasma technology. WIT Trans. Ecol. Environ. 2022, 257, 65–71. [Google Scholar] [CrossRef]
  254. Beletskii, E.V.; Kamenskii, M.A.; Alekseeva, E.V.; Volkov, A.I.; Lukyanov, D.A.; Anishchenko, D.V.; Radomtseu, A.O.; Reveguk, A.A.; Glumov, O.V.; Levin, O.V. One-step atmospheric plasma-assisted synthesis of FeOOH and FeOOH/graphite high performance anode materials for lithium-ion batteries. Appl. Surf. Sci. 2022, 597, 153698. [Google Scholar] [CrossRef]
  255. Bhar, M.; Dey, A.; Ghosh, S.; van Spronsen, M.A.; Selvaraj, V.; Kaliprasad, Y.; Krishnamurthy, S.; Martha, S.K. Plasma jet printing induced high-capacity graphite anodes for sustainable recycling of lithium-ion batteries. Carbon 2022, 198, 401–410. [Google Scholar] [CrossRef]
  256. Geng, J.; Zhang, S.; Hu, X.; Ling, W.; Peng, X.; Zhong, S.; Liang, F.; Zou, Z. A review of graphene-decorated LiFePO4 cathode materials for lithium-ion batteries. Ionics 2022, 28, 4899–4922. [Google Scholar] [CrossRef]
  257. Habibi, A.; Mousavi, M.R.; Yasoubi, M.; Sanaee, Z.; Ghasemi, S. Plasma-enhanced chemical vapor deposition for fabrication of yolk-shell SnO2@Void@C nanowires, as an efficient carbon coating technique for improving lithium-ion battery performance. Mater. Sci. Semicond. Process. 2022, 149, 106901. [Google Scholar] [CrossRef]
  258. Han, M.; Kim, D.W.; Kim, Y.C. Charged polymer-coated separators by atmospheric plasma-induced grafting for lithium-ion batteries. ACS Appl. Mater. Interfaces 2016, 8, 26073–26081. [Google Scholar] [CrossRef] [PubMed]
  259. Hou, D.; Bai, F.; Dong, P.; Chen, J.; Zhang, Y.; Meng, F.; Zhang, Z.; Zhang, C.; Zhang, Y.; Hu, J. Recent development of low temperature plasma technology for lithium-ion battery materials. J. Power Sources 2023, 584, 233599. [Google Scholar] [CrossRef]
  260. Jiao, J.; Du, K.; Wang, Y.; Sun, P.; Zhao, H.; Tang, P.; Fan, Q.; Tian, H.; Li, Q.; Xu, Q. N plasma treatment on graphene oxide-MoS2 composites for improved performance in lithium ion batteries. Mater. Chem. Phys. 2020, 240, 122169. [Google Scholar] [CrossRef]
  261. Jin, S.Y.; Manuel, J.; Zhao, X.; Park, W.H.; Ahn, J.H. Surface-modified polyethylene separator via oxygen plasma treatment for lithium ion battery. J. Ind. Eng. Chem. 2017, 45, 15–21. [Google Scholar] [CrossRef]
  262. Lee, C.; Han, Y.J.; Seo, Y.D.; Nakabayashi, K.; Miyawaki, J.; Santamaría, R.; Menéndez, R.; Yoon, S.H.; Jang, J. C4F8 plasma treatment as an effective route for improving rate performance of natural/synthetic graphite anodes in lithium ion batteries. Carbon 2016, 103, 28–35. [Google Scholar] [CrossRef]
  263. Li, C.; Li, H.L.; Li, C.H.; Liu, Y.S.; Sung, Y.C.; Huang, C. Effects of low-pressure nitrogen plasma treatment on the surface properties and electrochemical performance of the polyethylene separator used lithium-ion batteries. Jpn. J. Appl. Phys. 2018, 57, 01AB03. [Google Scholar] [CrossRef]
  264. Li, T.; Nie, X. One-step fast-synthesized foamlike amorphous Co(OH)2 flexible film on Ti foil by plasma-assisted electrolytic deposition as a binder-free anode of a high-capacity lithium-ion battery. ACS Appl. Mater. Interfaces 2018, 10, 16943–16946. [Google Scholar] [CrossRef]
  265. Liu, Y.; Zhang, L.; Zhao, Y.; Shen, T.; Yan, X.; Yu, C.; Wang, H.; Zeng, H. Novel plasma-engineered MoS2 nanosheets for superior lithium-ion batteries. J. Alloys Compd. 2019, 787, 996–1003. [Google Scholar] [CrossRef]
  266. Ouyang, L.; Guo, L.; Cai, W.; Ye, J.; Hu, R.; Liu, J.; Yang, L.; Zhu, M. Facile synthesis of Ge@FLG composites by plasma assisted ball milling for lithium ion battery anodes. J. Mater. Chem. A 2014, 2, 11280. [Google Scholar] [CrossRef]
  267. Nava-Avendaño, J.; Veilleux, J. Plasma processes in the preparation of lithium-ion battery electrodes and separators. J. Phys. D Appl. Phys. 2017, 50, 163001. [Google Scholar] [CrossRef]
  268. Shen, F.; Zhang, F.; Yin, Y.; Chen, S.; Gao, J.; Li, J.; Han, X. A high performance lithium metal anode enabled by CF4 plasma treated carbon paper. Nanoscale 2021, 13, 11800–11807. [Google Scholar] [CrossRef] [PubMed]
  269. Sun, M.; Wu, J.; Lu, P.; Zhang, Z.; Zhang, Y.; Li, D. Sphere-like MoS2 and porous TiO2 composite film on Ti foil as lithium-ion battery anode synthesized by plasma electrolytic oxidation and magnetron sputtering. J. Alloys Compd. 2021, 427, 131032. [Google Scholar] [CrossRef]
  270. Sun, S.; WU, Y.; Zu, J.; Lu, C.; Sun, Y.; Wang, Z.; Chen, J. Stabilizing plasma-induced highly nitrogen-deficient g-C3N4 by heteroatom-refilling for excellent lithium-ion battery anodes. Chem. Eng. J. 2022, 892, 162075. [Google Scholar] [CrossRef]
  271. Tsai, H.S.; Hsu, C.H.; Chi, C.C.; Wang, Y.C.; Liu, F.W.; Tang, S.Y.; Tsai, C.J.; Ouyang, H.; Chueh, Y.L.; Jiang, J.H. Non-layered Ti2N synthesized by plasma process for the anodes of lithium-ion batteries. Inorg. Chem. Front. 2019, 6, 172–175. [Google Scholar] [CrossRef]
  272. Wei, Y.; Yu, H.; Li, H.; Ming, H.; Pan, K.; Huang, H.; Liu, Y.; Kang, Z. Liquid-phase plasma synthesis of silicon quantum dots embedded in carbon matrix for lithium battery anodes. Mater. Res. Bull. 2013, 48, 4072–4077. [Google Scholar] [CrossRef]
  273. Wu, J.; He, X.; Li, G.; Deng, J.; Chen, L.; Xue, W.; Li, D. Rapid construction of TiO2/SiO2 composite film on Ti foil as lithium-ion battery anode by plasma discharge in solution. Appl. Phys. Lett. 2019, 114, 043903. [Google Scholar] [CrossRef]
  274. Wu, L.Y.; Chung, F.Y.; Huang, C. Synthesis and application of nano-organosilicon coating through cyclonic plasma deposition on a polymeric separator for lithium-ion batteries. J. Coat. Technol. Res. 2022, 19, 1159–1170. [Google Scholar] [CrossRef]
  275. Zhang, X.; Jin, M.; Zhao, Y.; Bai, Z.; Wu, C.; Zhu, Z.; Wu, H.; Zhou, J.; Li, J.; Pan, X. Improved lithium-ion battery performance by introducing oxygen-containing functional groups by plasma treatment. Nanotechnology 2021, 32, 275401. [Google Scholar] [CrossRef]
  276. Zhong, X.; Liu, W.; Han, J.; Jiao, F.; Qin, W.; Liu, T.; Zhao, C. Pyrolysis and physical separation for the recovery of spent LiFePO4 batteries. Waste Manag. 2019, 89, 83–93. [Google Scholar] [CrossRef]
  277. Zhu, J.; Chen, J.; Luo, Y.; Sun, S.; Qin, L.; Xu, H.; Zhang, P.; Zhang, W.; Tian, W.; Sun, Z. Lithiophilic metallic nitrides modified nickel foam by plasma for stable lithium metal anode. Energy Storage Mater. 2019, 23, 539–546. [Google Scholar] [CrossRef]
  278. Joseph, J.; Murdock, A.T.; Seo, D.H.; Han, Z.J.; O’Mullane, A.P.; Ostrikov, K.K. Plasma enabled synthesis and processing of materials for lithium-ion batteries. Adv. Mater. Technol. 2018, 3, 1800070. [Google Scholar] [CrossRef]
  279. Hsueh, T.H.; Tsai, C.H.; Liu, S.E.; Wang, M.C.; Chang, S.M.; Shiue, A.; Chin, K.Y. LiCoO2 battery electrode fabricated by high deposition-rate atmospheric plasma spraying for lithium battery. J. Electrochem. Soc. 2022, 169, 100506. [Google Scholar] [CrossRef]
  280. Zhang, X.; Hayashida, R.; Tanaka, M.; Watanabe, T. Synthesis of carbon-coated silicon nanoparticles by induction thermal plasma for lithium ion battery. Powder Technol. 2020, 371, 26–36. [Google Scholar] [CrossRef]
  281. Liu, P.; Zhang, J.G.; Turner, J.A.; Tracy, C.E.; Benson, D.K. Lithium-manganese-oxide thin-film cathodes prepared by plasma-enhanced chemical vapor deposition. J. Electrochem. Soc. 1999, 146, 2001. [Google Scholar] [CrossRef]
  282. Cao, S.; He, X.; Nie, L.; Hu, J.; Chen, M.; Han, Y.; Wang, K.; Jiang, K.; Zhou, M. CF4 plasma-generated LiF-Li2C2 artificial layers for dendrite-free lithium-metal anodes. Adv. Sci. 2022, 9, 2201147. [Google Scholar] [CrossRef] [PubMed]
  283. Jiang, Q.; Zhang, H.; Wang, S. Plasma-enhanced low-temperature solid-state synthesis of spinel LiMn2O4 with superior performance for lithium-ion batteries. Green Chem. 2016, 18, 662–666. [Google Scholar] [CrossRef]
  284. Liang, Z.; Li, T.; Chi, H.; Ziegelbauer, J.; Sun, K.; Wang, M.; Zhang, W.; Liu, T.; Cheng, Y.T.; Chen, Z.; et al. Solvent-free manufacturing of lithium-ion battery electrodes via cold plasma. Energy Environ. Mater. 2024, 7, e12503. [Google Scholar] [CrossRef]
  285. Zhou, Q.; Zhao, Z.; Wang, Z.; Dong, Y.; Wang, X.; Gogotsi, Y.; Qiu, J. Low temperature plasma synthesis of mesoporous Fe3O4 nanorods grafted on reduced graphene oxide for high performance lithium storage. Nanoscale 2014, 6, 2286–2291. [Google Scholar] [CrossRef]
  286. Ma, X.; Azhari, L.; Wang, Y. Li-ion battery recycling challenges. Chem 2021, 7, 2843–2847. [Google Scholar] [CrossRef]
  287. Windisch-Kern, S.; Gerold, E.; Nigl, T.; Jandric, A.; Altendorfer, M.; Rutrecht, B.; Scherhaufer, S.; Raupenstrauch, H.; Pomberger, R.; Antrekowitsch, H.; et al. Recycling chains for lithium-ion batteries: A critical examination of current challenges, opportunities and process dependencies. Waste Manag. 2022, 138, 125–139. [Google Scholar] [CrossRef]
  288. Rajaeifar, M.A.; Raugei, M.; Steubing, B.; Hartwell, A.; Anderson, P.A.; Heidrich, O. Life cycle assessment of lithium-ion battery recycling using pyrometallurgical technologies. J. Ind. Ecol. 2021, 25, 1560–1571. [Google Scholar] [CrossRef]
  289. Cheret, D.; Santen, S. Battery Recycling. U.S. Patent 7,169,206 B2, 30 January 2007. [Google Scholar]
  290. Hong, R.; Su, L.; Chen, S.; Long, Z.; Wu, J. Comparison of cutinases in enzymic deinking of old newsprint. Cellulose 2017, 24, 5089–5099. [Google Scholar] [CrossRef]
  291. Wang, F.; Zhang, X.; Zhang, G.; Chen, J.; Sang, M.; Long, Z.; Wang, B. Studies on the environmentally friendly deinking process employing biological enzymes and composite surfactant. Cellulose 2018, 25, 3079–3089. [Google Scholar] [CrossRef]
  292. Mauchauffé, R.; Lee, S.J.; Han, I.; Kim, S.H.; Moon, S.Y. Improved de-inking of inkjet-printed paper using environmentally friendly atmospheric pressure low temperature plasma for paper recycling. Sci. Rep. 2019, 9, 14046. [Google Scholar] [CrossRef] [PubMed]
  293. Gaiolas, C.; Costa, A.P.; Silva, M.S.; Thielemans, W.; Amaral, M.E. Cold plasma-assisted paper recycling. Ind. Crops Prod. 2013, 43, 114–118. [Google Scholar] [CrossRef]
  294. Wang, R.; Shen, Y.; Zhang, C.; Yan, P.; Shao, T. Comparison between helium and argon plasma jets on improving the hydrophilic property of PMMA surface. Appl. Surf. Sci. 2016, 367, 401–406. [Google Scholar] [CrossRef]
  295. Fowkes, F.M. Attractive forces at interfaces. Ind. Eng. Chem. 1964, 56, 40–52. [Google Scholar] [CrossRef]
  296. Kaelble, D. Dispersion-polar surface tension properties of organic solids. J. Adhes. 1970, 2, 66–81. [Google Scholar] [CrossRef]
  297. Owens, D.; Wendt, R. Estimation of the surface free energy of polymers. J. Appl. Polym. Sci. 1969, 13, 1741–1747. [Google Scholar] [CrossRef]
  298. Kwok, D.Y.; Lam, C.N.C.; Li, A.; Leung, A.; Wu, R.; Mok, E.; Neumann, A.W. Measuring and interpreting contact angles: A complex issue. Colloids Surf. A Physicochem. Eng. Asp. 1998, 142, 219–235. [Google Scholar] [CrossRef]
  299. Drelich, J.W.; Boinovich, L.; Chibowski, E.; Volpe, C.D.; Hołysz, L.; Marmur, A.; Siboni, S. Contact angles: History of over 200 years of open questions. Surf. Innov. 2020, 8, 3–27. [Google Scholar] [CrossRef]
  300. Sinayobye, E.; Kakane, V.C.K.; Baryeh, E.A. Determination of polar and dispersive contributions of surface tensions of some macromolecular solutions. Int. J. Biol. Chem. Sci. 2012, 6, 887–896. [Google Scholar] [CrossRef]
  301. Kusano, Y.; van Acker, K.; Hutchings, I.M. Methods of data analysis for the micro-scale abrasion test on coated substrates. Surf. Coat. Technol. 2004, 183, 312–327. [Google Scholar] [CrossRef]
  302. Altay, B.N.; Ma, R.; Fleming, P.D.; Joyce, M.J.; Anand, A.; Chen, T.; Keskin, B.; Maddipatla, D.; Turkani, V.S.; Kotkar, P.R.; et al. Surface free energy estimation: A new methodology for solid surfaces. Adv. Mater. Inferf. 2020, 7, 1901570. [Google Scholar] [CrossRef]
  303. Sacher, E. An attempt to evaluate the dispersive and polar components of the surface tensions of multicomponent solutions, using kaelble’s method. J. Colloid Interf. Sci. 1981, 83, 649–651. [Google Scholar] [CrossRef]
  304. Kusano, R.; Boulton, G.; Kusano, Y. Saltwater-wettability on polymer surfaces and determination of surface tension. Int. J. Surf. Sci. Eng. 2021, 15, 281–293. [Google Scholar] [CrossRef]
  305. Kusano, R.; Kusano, Y. Symmetric expressions of surface tension components. J. Adhes. 2023, 99, 2381–2401. [Google Scholar] [CrossRef]
  306. Kusano, Y.; Kusano, R. Critical assessment of the correlation between surface tension components and Hansen solubility parameters. Colloids Surf. A Physicochem. Eng. Asp. 2023, 677, 132423. [Google Scholar] [CrossRef]
Figure 1. A diagram of the waste management hierarchy model.
Figure 1. A diagram of the waste management hierarchy model.
Materials 17 01687 g001
Figure 2. A diagram of the circular economy model.
Figure 2. A diagram of the circular economy model.
Materials 17 01687 g002
Figure 3. A configuration for a direct current (DC) thermal plasma torch.
Figure 3. A configuration for a direct current (DC) thermal plasma torch.
Materials 17 01687 g003
Figure 4. A configuration for a radiofrequency (RF) thermal plasma torch.
Figure 4. A configuration for a radiofrequency (RF) thermal plasma torch.
Materials 17 01687 g004
Figure 5. A configuration for a microwave (MW) thermal plasma torch.
Figure 5. A configuration for a microwave (MW) thermal plasma torch.
Materials 17 01687 g005
Figure 6. A configuration of corona discharge.
Figure 6. A configuration of corona discharge.
Materials 17 01687 g006
Figure 7. Configurations of DBD plasmas. (a) Volume discharge. (b) Surface discharge.
Figure 7. Configurations of DBD plasmas. (a) Volume discharge. (b) Surface discharge.
Materials 17 01687 g007
Figure 8. A configuration of a cold plasma torch.
Figure 8. A configuration of a cold plasma torch.
Materials 17 01687 g008
Figure 9. A photograph of a gliding arc.
Figure 9. A photograph of a gliding arc.
Materials 17 01687 g009
Figure 10. A diagram of the plasma gasification process.
Figure 10. A diagram of the plasma gasification process.
Materials 17 01687 g010
Table 1. Typical plasma parameters of plasma processes in air (values are rounded).
Table 1. Typical plasma parameters of plasma processes in air (values are rounded).
Type of PlasmaMax. Temp. (K)Electron
Density (cm−3)
ApplicationsRefs.
ElectronGas
ThermalPlasma torch (DC, RF, MW)1041041015Gasification
Metal recovery
[19]
Spark plasma1041041015–1018Sintering[20,21]
NonthermalCorona discharge105102–103108Surface modification
Plasma reforming
[20,22,23]
Dielectric barrier discharge104–1051014–1015[24]
Cold plasma torch1041014[25]
Gliding arc1042 × 1031015[26,27]
Table 2. Comparison of plasma torches used for recycling.
Table 2. Comparison of plasma torches used for recycling.
ElectrodeTypical Max. Power
Reported [kW]
AdvantageDisadvantage
DC plasma torchNeeded118.8 [83]High power applicableContamination by
electrode erosion
RF plasma torchNo need75 [84]ElectrodelessUse of matching network
MW plasma torchNo need10 [85,86]Electrodeless,
arguably energy efficient
Use of ignition system
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

Kusano, R.; Kusano, Y. Applications of Plasma Technologies in Recycling Processes. Materials 2024, 17, 1687. https://doi.org/10.3390/ma17071687

AMA Style

Kusano R, Kusano Y. Applications of Plasma Technologies in Recycling Processes. Materials. 2024; 17(7):1687. https://doi.org/10.3390/ma17071687

Chicago/Turabian Style

Kusano, Reinosuke, and Yukihiro Kusano. 2024. "Applications of Plasma Technologies in Recycling Processes" Materials 17, no. 7: 1687. https://doi.org/10.3390/ma17071687

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