Next Article in Journal
Use of Thermodynamically Consistent Phase Equilibrium Data to Obtain a Generalized Padé-Type Model for the Henry’s Constants of Gases in Ionic Liquids
Next Article in Special Issue
Correlations between a Friabilin Content Indicator and Selected Physicochemical and Mechanical Properties of Wheat Grain for Processing Suitability Assessment
Previous Article in Journal
Ionic Liquids as Potential Cleaning and Restoration Agents for Cellulosic Artefacts
Previous Article in Special Issue
Prospects of Processing Tungsten Ores from the Akchatau Deposit
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pulsed Electric Field Technology for Recovery of Proteins from Waste Plant Resources and Deformed Mushrooms: A Review

by
Ramya Ramaswamy
1,*,
Sivaneasan Bala Krishnan
2 and
Susanna Su Jan Leong
1
1
Food, Chemical and Biotechnology Cluster, Singapore Institute of Technology, Dover Drive, Singapore 138683, Singapore
2
Engineering Cluster, Singapore Institute of Technology, Dover Drive, Singapore 138683, Singapore
*
Author to whom correspondence should be addressed.
Processes 2024, 12(2), 342; https://doi.org/10.3390/pr12020342
Submission received: 20 January 2024 / Revised: 2 February 2024 / Accepted: 2 February 2024 / Published: 6 February 2024

Abstract

:
Proteins are complex molecules, which play a vital role in our body’s function, the building of tissues, and the regulation of metabolic activity. They are crucial to children’s growth and serve as a key component in the body’s process of distributing oxygen. Proteins fuel the body by supplying the required nutrition and energy. Currently, there is an increasing demand for proteins on large scales with no detrimental effects. The adverse health effects of animal proteins have resulted in a growing preference for plant-based proteins, which offer a healthier daily dosage. Valuable proteins can be extracted from various parts of the plant, including stems, leaves, seeds, fruits, vegetables, and roots. Notably, protein extraction from waste plant and mushroom parts minimizes the product wastage and improves the overall production to support economic sustainability. There are several protein extraction techniques available, where the replacement of non-thermal methods with thermal ones is promising nowadays due to the appreciable retainment of protein quality. Pulsed Electric Field (PEF) technology is one of the most efficient non-thermal tools used to assist with extracting these proteins at the minimum processing time and energy consumption when compared with thermal techniques. It relies on the application of a high-voltage pulse between two electrodes to treat samples inside the treatment chamber. While electrode shapes and treatment chamber designs primarily govern the electric field’s application, optimizing process parameters such as electric field strength, pulse width, number of pulses, and pulse waveshape assists in obtaining a desirable enhancement in the protein yield. The primary objective of this review is to explain the PEF-assisted protein extraction process applicable to waste plant parts and deformed mushrooms. While PEF is not a novel concept, utilizing it as a pre-extraction treatment to the aforementioned waste resources would aid in improving the production of value-added protein products economically. So far, PEF has shown immense promise in assisting with protein extraction studies, but requires further research in order to establish this area for large-scale industrial applications.

1. Introduction

Based on the Food and Agriculture Organization, around one third of the food produced for human consumption is wasted globally, i.e., 1.3 billion tons per year [1]. According to the United Nations Environment Programme’s Food Waste Index Report, 17 percent of our food ends up being wasted in retail and by consumers, particularly in households [2]. For instance, India is among the countries that remains high in the production of fruits and vegetables. However, it suffers from 40% wastage of total food production. In that, fruits and vegetables contribute to 18% wastage of total food production [3]. These wastes comprise leaves, peels, stems, roots, and seeds, which have been found to have high nutritional value and contain a substantial amount of valuable bio-compounds, including proteins [4]. Upon extraction, the byproducts of these wastes can be efficiently utilized in food processing and pharma industries to produce nutritional and pharmaceutical supplements.
Several research studies have observed the significance of the protein requirement to the body [5,6]. This requirement is increasing globally due to the high population growth and the awareness of protein’s positive impact on the body. Research studies indicate the Recommended Dietary Allowance of protein to be 0.8 g per kg body weight (BW) per day since insufficient protein intake can lead to anemia, physical weakness, stunting, edema, vascular dysfunction, and impaired immunity [7]. Dietary recommendations from other countries, including the joint FAO/WHO committee, do not significantly deviate from the US recommendations for macronutrient intake [8,9]. In the case of insufficient protein intake, protein supplements can provide dietary compensation. Hence, it is necessary to consume a protein-rich diet every day since protein’s nutritional benefits are decided by both the quality and quantity. It has been reported that sufficient consumption of plant proteins may reduce the risk of all-cause and cardiovascular disease mortality and may increase longevity [10]. Thus, proteins play a significant role in providing nutritional value to foods and act as thickening, stabilizing, emulsifying, foaming, gelling, and binding agents.
Over the past few decades, a consistent increase in the demand for proteins has been observed due to multiple health benefits. With a global population of 7.3 billion, the demand for protein consumption has increased to approximately 202 million tons [11]. This demand may exceed the protein supply in future decades if the available resources are not efficiently utilized. This has led to the search for plant-based protein resources since the perception of the negative health effects of animal proteins has escalated. This extensive search is carried out with the objective of producing value-added products from the plant protein isolates in order to meet the ever-increasing protein demand. In order to improve the socioeconomic viability and to minimize product wastage, research is currently in the phase of extracting proteins from waste plant resources such as vegetable peels, fruit peels, seeds, stems, and roots from the processing industries. When compared with animal-derived waste, plant-derived waste makes up 63% of the overall food supply chain [12]. Globally, huge quantities of mushroom and plant pieces are discarded from packaging industries since they do not meet the required standards for size and shape regulations. These rejected or wasted plant parts, which are excellent sources of proteins and other necessary bioactive compounds, will typically be disposed of or utilized using conventional practices [13]. Here, valorization can play a major role in the efficient utilization of these resources to produce value-added products and minimize the overall wastage. Figure 1 illustrates the valorization of waste plant resources [14].
Plant waste valorization is the process of transforming plant wastes into higher-value products by converting them into chemicals or fuels or extracting them to produce nutrient-rich byproducts or supplements. The significant advantages of plant wastes include (i) ease of access, (ii) sustainability, and (iii) biodegradability [15]. Pertaining to the global protein requirement, efficient protein extraction techniques are prioritized to meet the rising demand for high-quality proteins. The food industry has constantly explored pathways to the extraction of plant-based proteins using several conventional processing methods, such as Soxhlet extraction, thermosonication, heat reflux, maceration, bead milling, mechanical grinding, ultrasonic treatment, osmotic treatments, and acid and alkaline pretreatments [16,17,18,19], which involve solvent usage, chemicals, heat, and the heavy mechanical disruption of plant cells, resulting in a requirement for additional processing due to purification and the low extraction efficiency. These poor results depend on the size of the testing particle, the usage of solvents, the pH value, the extraction time, and the temperature, which eventually degrade the quality and affect the functional properties of extracted proteins. Emerging non-thermal methods are gaining recognition due to their eco-friendly nature, consumption of less energy, and generation of minimal to no thermal effects on the extracted bio-compounds. These methods do not have an environmental impact and preserve the overall quality and nutrition of the resulting proteins. The application of a Pulsed Electric Field (PEF) is supposed to be a good alternative to thermal treatments in order to overcome the undesirable degradation in the protein quality. This non-thermal technique has been used for several years to inactivate food-borne pathogens in order to extend the shelf life and preserve the quality of foods [20,21,22] throughout their storage time [23,24]. When compared with other thermal food processing methods, PEF preserves the nutritive value of foods with the minimum degree of damage [25,26,27]. Initial PEF studies were reviewed by multiple authors [28,29,30,31,32,33,34,35,36], which aided in the development of PEF into having a global reach. The gradual scaling-up of lab-scale to industrial systems has led to developments in the design of high-voltage pulse generators that provide higher-end performance to suit PEF industrial processing [37,38]. Also, the optimization of electrode shapes and chamber designs have overcome the existing drawbacks of electric field enhancement and dielectric breakdown in foods [39]. Due to the myriad advantages of PEF, recent research has focused on using the technique in protein extraction procedures, where the finally derived protein isolates would be used for producing value-added products including health supplements and nutraceutical and pharmaceutical products. Though PEF enables physiological, functional, and enzymatic changes in proteins, the area still requires further insights into the factors causing these changes in order to promote them for industrial processing.
This review discusses protein extraction from waste plant resources and deformed mushrooms using PEF treatments. This paper covers waste resources, protein extraction methods, the Pulsed Electric Field technique, PEF-assisted extraction of proteins, optimization of the PEF-assisted extraction process, industrial processing, existing challenges, and future directions.

2. Waste Resources

2.1. Plant Wastes

Protein is the most significant macronutrient and is essential to our daily survival needs. Though animal proteins, including meat, fish, milk, and eggs, are rich in essential amino acids, minerals, and vitamins, they use extensive land and feed resources. Also, animal proteins cause adverse health effects that nowadays necessitate low-cost alternative protein sources [40]. Plant protein sources have several advantages, such as a lower cost of production, ease of access, minimal maintenance, and better environmental sustainability. Developing countries rely on plant proteins as they are more economical than animal proteins [9]. Specifically, when the proteins are extracted from waste plant resources, such as seeds, stems, peels, and roots, the extraction of proteins can be made more economical. Vegetable and fruit processing industries generate a huge amount of plant waste (about 90 to 92%) [41]. This leads to a high percentage of wastage at a global level. Plant wastes can be recycled for value-added purposes, thereby reducing environmental pollution and enabling sustainable green development. Some examples of nutrient-rich industrial plant wastes include papaya seeds, sesame press cake, olive pomace, flaxseed hulls, tomato residues, and brewer spent grains. These wastes will be dumped in the garbage, landfilled, or used as cattle feed. Selva Ganesh et al. reviewed the downsides of excessive vegetable and fruit waste and discussed the emerging opportunities to generate value-added products from these wastes [42]. They have also indicated the ever-increasing level of global food wastage and have recommended that we evaluate efficient waste management techniques to valorize and minimize wastage in the long term [42]. In this regard, the accurate monitoring of plant waste is necessary in order to minimize overall wastage. Although plant wastes are produced in all stages of the fruit and vegetable manufacturing chain, recovery and recycling are essential factors to be considered to improve sustainability. The recycling of plant wastes for the recovery of bioactive compounds to produce value-added products minimizes the environmental stress in the global scenario. Table 1 refers to examples of protein-rich plant wastes that can be used for ongoing protein extraction studies.
For a protein source to be considered valuable, it must have substantial protein content with an essential amino acid composition [64]. From Table 1, plant wastes such as sesame cake, papaya seeds, brewer spent grains, rapeseed meal, and inodorus melon peel contain a higher percentage of proteins than other resources that, if extracted, would be beneficial in producing value-added products in an economical manner. Effective valorization of these plant wastes can improve food safety, security, and sustainability in food production.

2.2. Deformed Mushroom Wastes

Mushrooms have the potential to expel large quantities of proteins when extracted. Mushrooms belong to the fungi family, which has a ‘plant-like’ structure consisting of stems and heads lacking chlorophyll. Edible mushrooms are low-calorie, no-fat, and no-cholesterol foods that are rich in protein, fibers, and antioxidants. The protein quality of mushrooms has been found to be superior to that of plant sources. The nutritional value of edible mushrooms is primarily due to their protein content, consisting of all essential amino acids required to meet each day’s dietary requirements, and they are more economical than plant and animal sources [65]. Generally, edible mushrooms have a protein level ranging from 6.60 to 36.87 g/100 g dry weight, with an average of 23.80 g/100 g [66]. The amino acid composition of the proteins in mushrooms is equivalent to that of animal proteins and, hence, they are considered to be more valuable than plant proteins [67,68]. However, different mushroom species will differ in terms of the overall amino acid composition [69]. Based on earlier studies, the protein digestibility of mushrooms ranges from 60% to 70% [70], which can be further improved by eliminating the components that prevent proteins from being hydrolyzed in the gastrointestinal system [70]. Additionally, mushrooms have a significant role in preventing and suppressing disease-causing factors [71].
Mushrooms are consumed widely due to their excellent organoleptic properties and are considered to be novel protein alternatives [65]. Due to their flavor-enhancing and nutritional characteristics, the demand for mushroom consumption is increasing every year. Mushrooms possess antiviral, anti-inflammatory, antifungal, antithrombotic, and antimutagenic functional properties, which are used extensively in foods and medicinal formulations [72,73,74,75]. Around 2000 species of mushrooms exist in our environment but only a few species can be used for edible purposes [76]. The most widely grown edible mushrooms are Agaricus bisporus, Lentinula edodes, and Pleurotus species, which are produced primarily in China [77]. The consumption of edible mushrooms on a per-person basis increased by approximately fourfold between 1997 and 2013 [78]. In addition, it is anticipated that the global production of mushrooms will increase to 20.84 million tons by 2026 [79]. On the other hand, the edible mushroom industry’s accelerating growth and increasing production has led in parallel to a regrettable level of wastage (60 million tons of mushrooms per year) [80].
In the overall wastage, large quantities of misshapen parts are produced, whose size and shape do not comply with commercial standards. Since these misshapen mushroom parts are still edible, they are used as cattle feed to overcome wastage. It is noted that these misshapen parts are rich sources of carbohydrates, proteins, vitamins, minerals, and healthy fats. Unfortunately, there are few studies on mushroom-extracted bio-valuable compounds [80].
It is a known fact that proteins play a significant role in providing adequate nutrition to the body, and mushrooms are considered to be highly valuable sources of proteins in the human diet. Hence, in order to not neglect these valuable nutritional sources and to overcome food and nutrient wastage, extraction is a wise approach to addressing food sustainability and nutritional quality issues. Due to the extensive demand for alternate protein sources these days, rejected mushroom parts have been found to be an economic resource to boost protein production. Thus, the functional properties of rejected mushroom parts can be improved by extracting valuable proteins from them and processing them to produce food, medicinal, and health supplements. Numerous studies have projected the health benefits of mushrooms due to their medicinal properties. Table 2 lists the commercially cultivated mushroom varieties that have been recommended for protein extraction studies. When these mushrooms are transported to packaging and processing industries, large quantities of rejected waste would be generated, which are not only edible but also have numerous medicinal uses.
Mushrooms are a great source of nutritionally valuable compounds, including proteins, lipids, polysaccharides, polyphenols, micronutrients, and vitamins. Rosello et al. describe the potential use of some novel non-conventional methods, including enzyme-assisted extraction, pulsed electric field, ultrasound, microwaves, and subcritical and supercritical fluid extraction, for the recovery of valuable compounds from mushrooms [107]. PEF technology is one such essential tool that can improve the functionality, extractability, and recovery of nutritionally valuable compounds as well as the bioavailability of micronutrients and components in a diverse variety of foods. To exploit mushroom extracts for various applications in food, pharmaceutical, and nutraceutical industries, they need to be characterized for their bioactive compounds through several extraction processes. The residues from the protein extraction can be further valorized into valuable products based on their potential, such as prebiotics, dietary fiber, and compost. Several research studies on mushrooms have provided a sustainable solution in the valorization of mushroom byproducts to produce valuable functional ingredients with health benefits, especially for the gastrointestinal system [108,109].

3. Protein Extraction Methods

An extraction process is described as the recovery of valuable bioactive compounds from plant cells by means of a mass transport phenomenon, where the intracellular components of the cell are transmitted to the solvent [110,111,112]. The stages of an extraction process typically include penetration of the solvent into the plant matrix, dissolution of solutes in the solvent, diffusion of the solute out of the matrix, and collection of extracted solutes [113]. Since these extracted solutes are bioactive compounds for which a high degree of purity is preferred, it is necessary to find efficient extraction techniques that are quick, easy, safe, and reliable, provide better yields, and, most importantly, require a minimal number of purification steps. Existing conventional protein extraction methods, including maceration, percolation, heat reflux extraction, and Soxhlet extraction, are widely adopted due to their simple operational techniques. However, these methods use large volumes of solvents and heat, have a long processing time, and are inefficient [114]. Under these methods, increasing the temperature increases the solubility and diffusion, leading to increased extraction. However, the huge rise in the temperature leads to the extraction of other undesirable impurities, which require additional processing for purification. While we discuss a few examples, alkaline extraction is an inexpensive chemical extraction method in which protein dissolution occurs in an alkaline medium, with the disadvantage of protein denaturation under highly alkaline conditions during the extraction procedure [115]. Also, the increase in temperature and extraction time induces thermal precipitation and denaturation effects that may degrade the quality of proteins [115]. Enzyme-assisted extraction is a biochemical method that relies on the use of enzymatic breaking of the rigid cell wall to boost the protein yield. Though the enzyme usage maintains the optimal pH in order to minimize protein denaturation, the process is an expensive one for industrial applications and the enzymes cannot achieve complete cell breakdown [116].
Since the primary goal is to achieve an enhanced yield in a shorter extraction time without affecting the protein quality to a high degree, the selection of the extraction technique is essential. As discussed earlier, conventional extraction methods have a long extraction time, have a high solvent requirement, cause increased damage to cells, and have low extraction efficiency. Modern extraction methods, including ultrasound-assisted extraction, high-voltage electrical-discharge-assisted extraction, high hydrostatic pressure extraction, and pulsed electric field-assisted extraction, are crucial in preserving a sustainable food chain by the recovery of essential bioactive compounds from plant wastes. These ecofriendly methods have several advantages, such as a minimal temperature increase, less solvent usage, a short extraction time, a quicker process, less damage to the plant matrix, and a higher quality of extracted products [117,118,119,120]. This can be made possible by optimizing the extraction parameters throughout the experimental approach. These physical methods of cell degradation provide enhanced protein yields with minimal or no usage of solvents and heat with no environmental impacts. Although research on these techniques is promising, large-scale applications would require a long period of time [121]. In the following section, we discuss the Pulsed Electric Field technique, which works under the phenomenon of electroporation. This method has shown favorable extraction qualities with minimal damage to the plant cell matrix.

4. PEF Technique

4.1. General Features

PEF is a non-thermal technique that uses ultrashort electric field pulses in various applications such as food processing, tissue ablation, electrochemotherapy, and electro-extraction processes [20,122,123,124]. In food applications, PEF has been used in microbial inactivation studies, the extraction of bio-compounds, assisting with the drying process, and oil extraction [125,126,127,128]. While applications in microbial inactivation were initially invented and earned a greater amount of potential in research, recent research has focused on using the method for the extraction of bioactive compounds from the plant cell matrix [129,130]. When compared with thermal methods, the qualities of plant-protein-based foods were minimally affected after the PEF treatment and were maintained throughout the storage period [24]. The PEF method involves subjecting the plant cells to intense high-voltage pulses at an electric field strength range between 0.5 and 50 kV/cm with a pulse duration of milliseconds, microseconds, or nanoseconds in samples positioned between two electrodes to inactivate microorganisms in foods and to improve the rate of mass transfer for extracting intracellular bioactive compounds [131,132]. Basically, the cell membrane is made up of a lipid bilayer with embedded protein channels that play a critical role in separating the intracellular organelles of the cell from the surrounding environment and modulate the transfer of ions in and out of the cell to maintain the cell balance. The application of high-voltage electric field pulses destabilizes the lipid bilayer by increasing the cell wall permeability, providing adequate contact between the solvent (plain water in this case) and the intracellular compounds. Hence, while the plant cell matrix receives the electric field pulses, the permeability of the cell wall would increase, resulting in the induction of a transmembrane voltage across the cell membrane, the intensity of which depends on the intensity of the electric field, the pulse profile, the nature of cells, and the cell orientation with reference to the direction of the electric field vector [133]. In general, increasing the transmembrane potential increases the electromechanical compression of the cell membrane, leading to electropermeabilization or electroporation [134,135,136]. As a consequence of electroporation, plant tissues will be softened, enabling them to be well equipped for the subsequent extraction of bio-compounds in a shorter processing time. Thus, electroporation has a wide range of potential applications, from changing the structure and function of plant cells to causing permanent cell death. Figure 2 illustrates the potential uses of PEF in food applications.
Beneficial features of the PEF method rely on not using harsh chemicals, heavy solvents, acids, alkalis, or high-frequency mechanical effects on cells. It merely focuses on optimizing ultrashort electric field pulses to enhance the electroporation effect on plant cells through mass transport phenomena for efficient protein extraction [112,126]. When compared with thermal methods, it is a promising method for preserving the quality of proteins throughout the storage period [24]. Additionally, using plain water as a solvent minimizes environmental impacts. The interest in using PEF-assisted extraction has increased significantly since it has developed into an environmentally friendly and green extraction technique [137,138] that can contribute to sustainable development in future food systems. During the PEF bio-extraction procedure, electroporation selectively obtains the valuable bio-compounds from the plant cells by optimizing the intensity of the electric field strength and pulse parameters [139]. Thus, PEF-assisted extraction is advantageous in providing high protein yields when compared with other conventional techniques [130,140]. Moreover, combining other extraction processes with PEF has been reported to increase the protein yield [141,142] when compared with PEF standalone treatments. In addition to the enhanced yield, this method also preserves the quality attributes and nutrition of extracted bio-compounds [143,144]. For instance, PEF enhanced the functional and emulsifying properties of rice bran protein by about 20.29–22.64% and 3.3–12.0%, respectively. Additionally, the foaming capacity and foam stability were increased by 1.8 to 2.9-fold, respectively. Moreover, the in vitro digestibility of proteins was also enhanced [145].
Optimizing process parameters plays a significant role in changing the structure and function of proteins. Some of the structural changes, such as protein unfolding, dissociation, protein denaturation, and reaggregation, highly depend on optimizing the intensity of electroporation, which in turn depends on the electric field strength. When PEF effects on proteins are analyzed, the PEF-induced electric field stress causes the electrostatic interactions to be disrupted inside the polypeptide chain of the protein, leading to the deformation of secondary structures [146,147]. Multi-protein food systems have more complex interactions in an electric field, which can impact the stability of the systems [148]. For instance, with an increase in the electric field strength and number of pulses, the rheological properties of soy milk, such as the viscosity, increase with the change in consistency [149]. Enhancing the surface hydrophobicity of the protein structure enhances the flavor of foods. In this regard, increasing the electric field strength and number of pulses resulted in protein unfolding and the destruction of hydrophobic interactions, leading to structural modifications that increased the surface hydrophobicity in the case of whey protein isolate [150].

4.2. PEF Setup

The working principle of PEF relies on the application of high-voltage pulses to samples placed between two electrodes, where one electrode will be charged and the other will be at the ground potential. A PEF prototype comprises a high-voltage pulse generator unit, a treatment chamber with two electrodes, a peristaltic pump for circulating samples, an oscilloscope for displaying pulses, and a control unit [22]. The untreated sample will be transferred to the PEF treatment chamber before it is subjected to the required high-voltage pulse. Samples will be tested in batches in the case of static treatment chambers or will be pumped into the chamber for continuous testing purposes. The pulse-forming network (not shown) plays a significant role in delivering the required pulse waveshape to the testing medium placed inside the PEF chamber. The pulses can be delivered in the form of square, exponential, or oscillatory waveshapes. The treatment chamber and the electrodes must be carefully designed since they govern the electric field distribution inside the chamber. Figure 3 represents a simplified sketch of a Pulsed Electric Field setup.
There are three basic electrode configurations, including parallel plate, coaxial, and collinear electrodes [151], the shapes of which can be optimized further to suit either batch or continuous modes of operation [152,153,154,155]. Among the designs, the parallel plate configuration was the earliest-invented design of high importance, with a simple construction that delivers a uniform electric field inside the treatment gap [156]. While applying the electric field pulses to the testing medium, the electrodes act as a mediating factor between the high-voltage pulse generator and the treatment gap. Though the treatment chamber and electrode designs play the primary role in providing the optimum electric field distribution, process parameters such as electric field intensity, pulse waveshape, pulse duration, and pulse frequency should be adjusted to produce beneficial impacts on the treated samples [157,158]. The impact of the treatment also depends on the properties of the testing medium, such as the electrical conductivity, pH, concentration, and permittivity [111,159]. Hence, to obtain the best treatment outcomes, the researcher has to optimize the combination of all the aforementioned parameters.
High-voltage pulse generators deliver the desired electric pulses repetitively to treatment chambers through a pulse-forming network (PFN) and high-voltage switches. These generators provide the required waveshape, polarity, and amplitude of the pulses based on the construction of the PFN. While square and exponential pulses were frequently used in earlier studies [160,161], the square waveshape has been shown to be more efficient in treating the samples due to the longer retainment of the peak voltage during the treatment [161]. Exposing the cells to oscillatory decay pulses were found to be less effective due to the shorter cell exposure time. To minimize undesirable electrochemical reactions and associated thermal effects, bipolar pulses are generally preferred over monopolar pulses [160,162,163] due to the alternate stress effects on the cell membrane.
During the PEF method’s application, the sample will be treated either in batch or in continuous mode inside the treatment chamber [164]. Insulating spacers will be positioned between two electrode edges to facilitate the application of high-voltage pulses only in the treatment gap [112]. In earlier studies, batch testing modes were preferred for small-scale research and continuous modes were used for industrial testing purposes. Nowadays, lab-scale studies use both modes of operation. Batch operation involves testing the samples in batches using a standard treatment chamber. In this mode, there will be no continuous flow of the testing medium and the samples will be subjected to the static application of an electric field in the treatment gap for a specific treatment time. This mode of operation is slow to process due to the manpower involved in placing the testing medium inside the treatment chamber, transferring the medium, and collecting it after the treatment. Using a batch treatment chamber, the extraction of essential bio-compounds, including proteins, has been observed in several plant cells and waste plant tissues [4,165]. The sample to be treated will be placed either in small cuvettes or in a customized chamber, depending on the testing volume requirements. The product will be mixed with mild solvents or tap water in a prescribed ratio before it is subjected to the required high-voltage pulses [143].
During continuous operation, the peristaltic pump circulates the untreated samples and delivers them to the treatment chamber for the application of high-voltage pulses. When compared with batch systems, continuous systems are attractive due to their increased capacity to treat foods continuously in a shorter treatment time. Due to their high efficiency [166], continuous systems are preferred for large-scale industrial testing purposes not only in PEF-assisted extraction but also in other areas such as freezing, drying, cutting, and cold pasteurization [111,167]. In continuous mode, the electrodes and the chamber design will be constructed to facilitate the continuous flow of the testing medium inside the chamber at a constant flow rate per second/minute/hour depending on the volume of the testing sample. Tamborrino et al. used a collinear tubular electrode configuration for a continuous PEF-assisted olive oil extraction process, which resulted in an increase in the process efficiency and quantity [168]. One minor drawback of continuous chambers lies in the difficulty of having the product flow uniformly experience the electric field and subsequent electroporation effects. This has been observed in both microbial inactivation and bio-compound extraction studies [168,169]. The parallel plate and coaxial electrodes were generally used for both batch and continuous treatment processes [154,155,164,170,171,172] whereas collinear designs were allocated for the continuous treatment process in the majority of studies [168,173]. Despite fringing effects [174], the parallel plate design provides a more homogeneous electric field in the treatment area, whereas the coaxial shape ensures a uniform flow of the sample that is suitable for industrial-scale applications. Though all three electrode shapes are used for continuous PEF operations [175], the collinear design facilitates the easy flow of product through the chamber. The efficiency of both the batch and continuous testing modes is dependent on several factors, such as the chamber design, electrode design, electric field strength, number of pulses, pulse width, pulse waveshape, and properties of the testing medium such as the electrical conductivity, pH, permittivity, and concentration of the sample.

4.3. Mechanism of Extraction

PEF assists in the efficient extraction of intracellular proteins by means of a mass transport phenomenon through electroporation. Basically, the cell plasma membrane is considered to be a barrier, which protects the inside of the cell from dynamic environmental conditions. It is a semipermeable membrane that regulates the passage of nutrients and the ionic flow in and out of the cell, thereby creating an osmotic balance between the cell and its surroundings. This membrane acts as an insulator made up of lipid molecules consisting of polar (i.e., hydrophilic) and non-polar (i.e., hydrophobic) sections that together form a bilayer structure. When this membrane is subjected to high electric stress, the permeability and conductance of these membranes increase, leading to cell disruption that results in electropermeabilization or electroporation [176]. Electroporation is currently gaining recognition due to its cost-effective, time-saving, and chemical-free application in several areas, such as food processing, electrochemotherapy, tissue ablation, and bio-compound extraction protocols. Fundamentally, it is the process of inducing pore formation in the biological cell membrane, the radius of which depends on the intensity of the electric field strength [177]. The capacity of electroporation to cause the mass transport phenomenon in the plant matrix depends on the electric field strength, pulse number, pulse frequency, and total treatment duration [178]. The rationale behind producing electrical pores is to extract valuable bio-compounds from the cell wall matrix. Due to the increase in the electric field, transmembrane potential ( ¥ tmp) will be induced across the cell membrane that enables the membrane to experience electromechanical stress leading to the creation of pores. The pores will continue to increase depending on the supercritical electric field Ec [177]. Hence, the electroporation mechanism involves a significant number of stages, including application of ultrashort electric field pulses, lipid bilayer charging, local structural rearrangements within the cell, perforation of the cell membrane, and increased ionic and molecular transport through the pores [134]. The structural rearrangements indicate the permeation of the lipid bilayer leading to the formation of hydrophilic pores for an increase in the transmembrane potential value [134]. After being exposed to an electric field, cells cannot regain their original semipermeable properties completely [178].
When the electric field E is increased such that E > Ec, structural changes, cell rupture, and an increase in the cell membrane permeability will take place, leading to pore formation. Based on the increase in the electric field E, a larger number of pores will be formed, resulting in the leakage of intracellular biomolecules, including proteins, lipids, and carbohydrates. After the pulse’s application and the subsequent release of intracellular components, the resealing of pores will occur, and the cell will return back to its original state in the case of reversible electroporation (RE). The life span of pores ranges from milliseconds to minutes after the pulse’s application [176]. When the increase in the electric field E is very high such that E >> Ec, which progresses beyond the radius of the pores, permanent cell damage will occur, leading to irreversible electroporation (IRE) [179,180]. Therefore, the electroporation parameters should be optimized based on the treatment requirements and cell types. Figure 4 illustrates the electroporation stages in a plant cell.

5. PEF-Assisted Extraction of Proteins

In PEF-assisted extraction, the intensity of the electric field can significantly influence electroporation to soften the plant tissues and facilitate a better protein yield from the cell matrix. Hence, PEF-induced permeability changes in the cell membrane depend on the electric field intensity, extraction time, pH, specific energy, temperature, and properties of the cell matrix, especially the conductivity [181]. These electroporation effects can be observed on the cell surface using scanning electron microscopy (SEM), which can help the researcher to make analogous studies with other treatment effects on the cells [182]. Figure 5 illustrates the PEF-assisted protein extraction process.
Over the past few years, studies on PEF-assisted protein extraction have escalated exponentially due to the potential of this treatment to provide better extraction yields with quality retainment. This technique to extract proteins through the mass transport mechanism works by the diffusion of the solvent into the intracellular medium and the dissolution of the solute out of the cell matrix. Due to its beneficial advantages, several studies have incorporated PEF into protein extraction research [4,114,119,129,159,183,184,185]. Moreover, thermolabile compounds are well preserved using PEF when compared with existing thermal methods [18,112]. Table 3 presents several remarkable studies on PEF-assisted protein extraction from waste plant parts and mushrooms. These studies represent the protein yield through process parameter optimization.
The integration of PEF into the protein extraction process in mushrooms and waste plant parts is evident from earlier studies [4,119,143,165]. In addition to protein extraction, PEF–US combination techniques increased the drying rates, decreased the drying time, provided better mass transfer and rehydration capacities, changed the structure of the cell tissue, and showed high antioxidant activity in shiitake mushrooms [189]. In 2014, Boussetta et al. studied the impact of pulsed electric field-assisted extraction of polyphenols from waste flaxseed hulls. Their findings highlighted a research gap in protein extraction from flaxseed hulls [190].
PEF has also been proven to be effective in bio-compound extraction from seeds. Sesame seeds are extensively used in cooking to prepare tasty dishes and act as a flavoring agent in many recipes. These seeds are highly nutritious, are a good source of vitamins, healthy fats, protein, minerals, fiber, and antioxidants, and contain other beneficial characteristics for health. These seeds contain about 20% protein [191]. After oil extraction, the remaining byproducts of sesame seeds are still high in nutritional value. Sarkis et al. carried out extraction studies on sesame press cake for the protein and polyphenol content. A significant increase in protein extraction was observed at an optimum temperature of 40 °C, an electric field strength of 13.3 kV/cm, and a pulse duration of 10 µS [183]. The PEF method’s application increased the electrical conductivity and cell permeability of the residual seed cake, causing minimal cell damage. This enabled an easier extraction in a shorter time [183].
Papaya seeds are rich in isothiocyanates, phenolic compounds, and proteins. Parniakov et al. observed an increased degree of disintegration in papaya seeds at 13.3 kV/cm and 2000 pulses when coupled with supplementary aqueous extraction (SAE). The authors recovered a significant amount of protein after three hours of PEF-assisted extraction. Upon the PEF method’s application, the degree of extraction of ionic components facilitated the release of embedded proteins from the cell wall matrix during SAE [159]. Therefore, PEF and supplementary aqueous extraction together produced a higher protein yield, opening the door for industrial processing [159]. Yu et al. studied the PEF-assisted extraction of proteins from rapeseed stems and leaves, which primarily depended on the electric field strength and growth phase of the plant. The specific parameters that caused tissue damage were observed at 800 V/cm and a protein extraction yield of up to 80% was achieved at 20 kV/cm [4]. Kamboj et al. observed an increase in the protein yield of up to 10.58% on perilla seed meal at processing parameters of a 4 kV/cm electric field and a 120 µS pulse width [192]. Though PEF-assisted protein extraction studies have been performed on mushrooms and plant wastes, these edible products are currently in the amateur stage of study and require further exploration to improve the productivity economically.

6. Optimization of the PEF-Assisted Extraction Process

The optimization of PEF-assisted extraction requires the execution of systematic procedures to enhance the recovery of proteins from the cell matrices. The primary factors that influence the PEF performance in the extraction process include the treatment chamber design and the electrode shape, which govern the electric field distribution inside the chamber. These factors must be carefully designed in order to provide the optimal treatment to the testing samples. Other process parameters that need to be adjusted include the electric field strength, number of pulses, pulse width, pulse waveshape, specific energy input, and temperature of the untreated sample, as these factors optimize the electroporation effects on the cells to be treated [193], facilitating the mass transport phenomenon to assist with extraction. If these process parameters are increased, the ability of electroporation to puncture the plant cells would increase, enhancing the overall extraction process. Also, a decrease in impedance would be noted on the cell tissues upon increasing the electric field strength [194]. However, increasing the process parameters by many folds beyond the extractable threshold of the cell matrix and a prolonged extraction time would cause a significant reduction in protein extraction levels. On the other hand, the electroporation effects also depend on the characteristics of the cell, such as the cell wall rigidity, cell arrangement, structure, water content, and usage of solvents. Hence, the process parameter optimization may be variable depending on each cell matrix.
Electric field strength is one of the most significant parameters to be optimized in order to obtain an enhanced protein extraction rate. While PEF technology relies on the application of an electric field in a lower range from 0.1 to 0.3 kV/cm and in a higher range from 20 kV/cm to 80 kV/cm [131,195], reducing the pulse duration with a higher electric field would aid in providing a higher extraction yield in a short processing time [126]. In dairy and plant proteins, the intensity of the electric field can modify the structure and techno-functional characteristics of proteins. In addition, a higher electric field strength can yield better emulsifying properties of proteins and protein–polysaccharide complexes [196]. Xue and Farid analyzed the impact of PEF-assisted protein extraction on white button mushrooms (Agaricus Bisporus) and obtained 49% ± 0.09 protein at a 38.4 kV/cm field intensity for 272 μs of treatment time and a treatment temperature of 85 °C. In addition, reducing the pulse width to 136 μs while keeping constant the electric field value retained the pink-to-reddish color similar to fresh extracts [119]. Although a high electric field intensifies the electroporation effect, care must be taken not to increase the electric field exponentially as it may synergistically increase the product temperature, resulting in a loss of nutrients in the extracted outcome. So far, more research findings have been reported on the fact that the electric field strength induces higher protein yields in microalgal cells [124,188,197] than in plant resources and mushrooms. The aim of this article is to open up new avenues for more PEF-assisted protein extraction research on waste plant tissues and mushrooms.
The PEF treatment time can be calculated by multiplying the pulse duration by the number of pulses. Modulating either the pulse duration or the pulse number would modify the treatment time. The extraction time and temperature are the essential parameters to be optimized during the extraction process. An exponential increase in either of these parameters would degrade the quality of the testing medium. For instance, Boussetta and Segovia reported an increase in the yield of valuable compounds in grape seeds and borage leaves at a lower temperature and shorter extraction time than the control samples [137,198]. Hence, these two factors must be regulated appropriately to preserve nutritional and sensory aspects in the samples.
The efficiency of PEF technology also depends on the conductivity of the sample [20]. In microbial inactivation studies, high-conductivity foods created a peak electric field across the treatment chamber, which intervened into the electric field’s application during the treatment, resulting in a low-efficiency outcome. Hence, it is always preferred to choose low-conductivity foods to obtain optimum results in electroporating microorganisms. In plant-based bio-extraction studies, samples with high conductivity increased the ionic strength of the medium, which hastened the cell membrane charging, resulting in enhanced protein extraction levels. Hence, according to Buchmann, a highly conductive extracellular medium would enable the diffusion of proteins through the permeable membrane, resulting in an enhanced protein yield [124].
The application of ultrashort pulses (microseconds or nanoseconds) is considered to be one of the major advantages of this treatment that renders it non-thermal and consumes less energy when compared with thermal methods. This feature would assist in lowering the processing cost, though the initial cost of the equipment is high [199]. Hence, a suitable combination of optimized pulse parameters would provide effective extraction results. However, the process parameters must be selected based on the plant tissues in order to achieve high protein yields. Generally, the application of high-voltage pulses may not have an impact on intracellular organelles due to the presence of the rigid cell wall. However, if the pulse duration is made to be ultrashort, i.e., nanoseconds, sub-nanoseconds, or picoseconds, the electric shots will directly charge intracellular organelles in the cytoplasm without charging the outer cell membrane [200]. Hence, the cell wall characteristics play a significant role in the extractability of proteins [201]. However, when a long pulse duration is applied, it may lead to an undesirable increase in the temperature, resulting in adverse electrode reactions and thereby degrading the quality of the testing medium [202]. Hence, the application of a large number of ultrashort pulses would help in minimizing the temperature increase and associated detrimental effects [203].
The use of solvent at a suitable concentration has been highly beneficial in achieving the mass transport phenomenon in extraction studies [204]. By using PEF, the amount of energy and organic solvent needed to extract the desired compounds would be minimized [137,201]. Water is recognized as a green solvent liquid. However, the low solubility of valuable compounds in water makes it less preferable [205,206]. While the majority of extraction methods use ethanol, methanol, aqueous NaOH, KOH, NaHCO3, or NaCl as the solvent in a high concentration, PEF-assisted extraction has used different solvent types, such as water alone and a mixture of low-concentration organic solvents with water, for research. For instance, Min Wang [207] reported a higher protein yield from Chlorella using water alone as a solvent. This study gave the best extraction efficiency at an extraction time of 180 min at 3 kV/cm, 44 pulses, and 99 kJ/kg energy. Barba et al. studied PEF-assisted extraction of valuable bio-compounds from stevia leaves using water alone as a solvent, which facilitated an increase in the extraction of soluble matter (~33%), an increase in the activity of phenolic compounds (80%), chlorogenic acid (93%), caffeic acids (55%), ferulic acid (90%), and protocatechuic acids (45%), and increased the conductivity by 25% [111,208]. Parniakov et al. [209] compared one-stage and two-stage extraction procedures using the microalgae Nannochloropsis spp. In two-stage procedures, the combination of less-concentrated organic solvents and water obtained high non-degraded protein levels at 20 kV/cm for 400 pulses with a one-minute pause after each application of 200 pulses for better temperature control to 20 °C. The one-stage procedure resulted in the undesirable degradation of proteins due to the high concentration of solvents used. Serena Carpentieri et al. studied the PEF-assisted extractability of aroma and bioactive compounds from rehydrated plant materials. Here, green solvents such as a water–ethanol mixture and propylene glycol were used in solid–liquid extraction after the PEF application. PEF improved the solvent’s penetration into the plant cell’s cytoplasm and the diffusion rate of solubilized compounds during solid–liquid extraction, thereby enhancing the overall extractability rate [210].

7. Industrial Processing

In the 1960s, several possible applications were identified using electroporation in food and bioprocessing areas [211]. A PEF commercial installation for fruit juice processing was first achieved in 2006 in the USA followed by a line of installations in Europe in 2009, which was extended to applications for the pretreatment of vegetables in 2010 [211]. A PEF system has recently been installed in several countries, including Italy (Amica Chips), Canada (Potatopro), and New Zealand (McCain Foods), from machine manufacturers including Pulsemaster from The Netherlands and Elea Technology GmbH from Germany, allowing these countries to manufacture their own food products. For industrial applications, sustainable production is essential to feasible applicability in PEF-assisted protein extraction research, which requires further insights to become a commercially mature technique. Substantial factors to be considered for this analysis include reliability, economic evaluation, environmental impact, social impact, and relevant contribution to society. Therefore, it is essential to analyze all these factors for the successful integration of PEF into the protein extraction process in order to evaluate its efficiency. Based on earlier studies, PEF has been proven to be highly energy efficient in terms of applying ultrashort pulses when compared with other thermal processing methods. Due to the low energy consumption, the temperature of the PEF-treated sample would be lower when compared with conventional thermal techniques [212]. Hence, the overall processing cost would be lower, even though the initial cost of the machine is high [199].
Currently, there is high demand for green extraction processes that replace the non-ecofriendly conventional chemical and heavy solvent usage techniques. The objective of green techniques is to minimize the interference with solvent usage and enhance the qualitative and quantitative characteristics of the target extractable compounds. The PEF technology started with testing small quantities in batches using simple treatment chamber designs and electrodes at laboratory scales and has grown into a sustainable and commercially established technique in microbial inactivation studies. Later, many other divisions of application emerged, including solid and liquid food processing, spore inactivation, enzyme inactivation, dehydration, freezing, and medical science areas [213]. Moreover, the increase in demand led to an increase in the flexibility of PEF systems appropriate for each testing sample [112,214,215]. Over the past few years, PEF has been found to be effective in bio-compound extraction studies, but they are still in a preliminary stage [126]. Hence, further research is required in extraction studies to streamline the energy efficiency and safety processes suitable for industrial applications.
Due to the growing consumer demand for nutritional, safe, and high-quality food products, PEF technology has upscaled systems for industrial production. Section 4 discusses the PEF layout in detail. There are many significant parameters involved in the successful operation of a PEF system at the commercial scale. An effective treatment depends on optimizing these process parameters based on the volume of the sample treated at a larger scale. Arshad et al. have discussed the interdependency of these process parameters and the difficulty of modulating these parameters in larger-scale operations [216].
PEF systems deliver high-voltage pulses with high functional capabilities that meet large-scale industrial standards [217]. Stefan Toepfl installed high-power semiconductors that facilitated the evolution of industrial-scale PEF systems with a high degree of competence [218]. Research on the scaling-up of the PEF extraction process can be understood from several studies [219,220]. Toepfl’s analysis on upscaling from the lab scale to the industrial scale was achieved through an appropriate transfer of processing conditions suitable for larger scales. In this regard, the major processing parameters, such as the electric field strength, specific energy input, and temperature, were optimized, where a specific energy input of 5–10 kJ/kg was required for the permeabilization of plant tissue [38].

8. Existing Challenges for PEF Systems

At present, the occurrence of electrochemical reactions is a significant concern for pulsed electric field treatment in the food industry [221,222]. The main consequences of electrochemical reactions include corrosive effects, electrode fouling, migration of metallic particles from electrodes, chemical changes, and water electrolysis in the food matrix [222,223]. So far, research has been performed to minimize the electrolysis by (i) using bipolar square wave pulses [162,224], (ii) reducing the pulse duration [220,225], (iii) maintaining the minimum peak voltage [225], and (iv) coating the electrode [155]. Moreover, the availability of commercially established PEF units is limited worldwide, which requires further research to overcome the existing limitations [199]. Johan Morren et al. studied the electrolysis that occurs during PEF treatment, the effect of which could be minimized in the treated food by using small pulses [221]. If the electrodes are corroded, the PEF-treated samples would be contaminated with metal particles [221].
Another drawback is the occurrence of gas bubbles during the PEF treatment that affect the uniformity of the electric field’s application in the treatment area, resulting in operational issues [196]. Moreover, the majority of food samples are homogeneous in nature. The PEF’s application in heterogeneous foods leads to dielectric ruptures due to air and gas bubbles in the food product that occur due to the electrical discharge between the two electrodes inside the treatment chamber. This would create sparks, thereby increasing the probability of explosions [226].
The major disadvantage of a pulsed electric field is the extensive initial cost of the machine, which requires significant financial investment [30,227,228]. However, since the operational cost is very low [216,228], many industries and research institutes propose to buy the equipment, thereby evaluating PEF as a long-term profitable food processing business that has the potential to provide high research integrity.
Moreover, during the PEF’s application, varying the process parameters would cause electric-field-induced temperature changes in the food samples based on the conductivity value of the food. High-conductivity foods will have a high ionic concentration that would produce a peak electric field across the treatment chamber, thereby inducing heat energy. Hence, the temperature should be monitored and controlled to minimize the degradation of the thermolabile compounds in the plant cells [113]. For this purpose, mild treatments or circulating cooling water would aid in minimizing the associated temperature increase in the system, which would also help to preserve the quality attributes of the tested food [167,229,230].
The extraction of proteins from the plant matrix involves process parameter optimization, which poses a greater challenge to protein extraction research teams. Varying degrees of optimization would be a requirement to suit each type of plant matrix to obtain enhanced protein levels. In addition, promoting this research area on a larger scale requires established research reports pertaining to specific doses of the PEF to each protein source. Hence, further research is still required on molecular perception to find out the changes that occur in the protein levels from the plant matrix for each processing condition. In addition to protein yield, other factors such as the initial cost of equipment, maintenance and operating costs, sustainability, and the feasibility of PEF application under varying environmental conditions should be analyzed to upscale this technology to industrial processing. However, studies involving all of the above factors are still limited and we need deeper insights for further progress.

9. Conclusions and Future Directions

Protein is a mandatory requirement for the body’s everyday functioning. The protein requirement varies based on age and conditions in the human body. When it comes to plant-based dietary proteins, the extraction of proteins from plant wastes using green techniques favors both the economy and the environment, promoting sustainable production. From a global perspective, valorizing raw plant wastes and rejected mushroom parts would help to reduce the amount of fruit, vegetable and mushroom wastage and provide value-added products. In this regard, the PEF-assisted protein extraction method has the potential to exponentially grow in the future to meet the ever-escalating protein demand since this technique facilitates a high protein yield in a non-thermal manner that preserves the bioactive and associated thermolabile compounds. However, further studies are required in order to optimize process parameters and scale up the technology for industrial applications. PEF is considered to be a green technique, due to which its utilization will increase in the following years. As a limited number of studies are available on PEF-assisted protein extraction, emerging unknown aspects need to be examined in the future. In particular, the impact of different electric field strength values must be studied to find out the optimum value for enhanced protein yield at a controlled temperature, since temperature is one of the most important parameters to monitor. Other process parameters should also be optimized with the electric field until we obtain clear conclusions, which should be evaluated multiple times to obtain valid and established reports. These studies would facilitate the technique for industrial applications.
Recently, combinations of techniques with PEF have been found to be effective in obtaining enhanced protein levels when compared with PEF standalone treatments [231,232]. However, the additional method should not use severe processing conditions to degrade the proteins, which may in turn require further processing steps including purification. Zhou et al. studied protein extraction with a combination of PEF and pressurized liquid extraction (PLE) treatments compared with a PEF pretreatment with conventional extraction methods. The enhanced extraction was due to the initial cell breakdown effect caused by PEF followed by the easier diffusion of the solvent into the cell during PLE to achieve a high level of protein recovery from Spirulina [233]. Carullo et al. showed improved recovery of bioactive compounds including proteins from a combination of high-shear homogenization (HSH) and PEF treatments on A. platensis suspensions due to synergistic effects. The disaggregating effect of A. platensis trichome clusters due to the HSH pretreatment enhanced the cell membrane permeabilization, followed by the PEF treatment, thus intensifying the extractability of valuable compounds [142]. The experiment was maintained at a feasible temperature in order to achieve non-thermal effects in the samples. Therefore, the integration of two techniques would aid in enhancing the overall protein levels in the samples. However, further studies should be conducted in order to select an appropriate green technology to combine with PEF to obtain high protein yields, also facilitating the combination for larger scales. Moreover, the cost effectiveness and sustainability should also be considered in order to promote combined techniques to industrial levels.
The challenges and future developments that should be addressed are as follows:
The protein extraction research on PEF process parameter optimization should be broadened to promote the applications in a large-scale manner.
Deeper insights should be gained into PEF-assisted protein extraction in non-conforming botanical parts such as rejected stems, leaves, roots, flowers, fruits, and other green waste resources from industrial processing.
Research on combining other green techniques with PEF to obtain the optimum protein yield should be extended.
The research on electrolysis is underdeveloped in extraction studies and can be further improved to minimize the effects in the final byproducts.
Advanced simulation studies should be conducted to find out the protein’s structural and functional changes due to electric stress, which would improve our understanding of the interaction between the protein molecule and electric field and consequently be beneficial for future experimental research.

Funding

This work is supported by Singapore Institute of Technology (SIT) under the Food Extraction Programme.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Global Food Losses and Food Waste. Available online: https://www.fao.org (accessed on 22 November 2023).
  2. UNEP Food Waste Index Report 2021|UNEP—UN Environment Programme. Available online: https://www.unep.org (accessed on 22 November 2023).
  3. NPCS Team. India Emerging Business Opportunities: Cold Chain Sector (Why to Invest, Project Potential, Core Financials, Market Size & Industry Analysis); NIIR Project Consultancy Services: New Delhi, India, 2014; pp. 11–12. [Google Scholar]
  4. Yu, X.; Bals, O.; Grimi, N.; Vorobiev, E. A new way for the oil plant biomass valorization: Polyphenols and proteins extraction from rapeseed stems and leaves assisted by pulsed electric fields. Ind. Crops Prod. 2015, 74, 309–318. [Google Scholar] [CrossRef]
  5. Lonnie, M.; Hooker, E.; Brunstrom, J.M.; Corfe, B.M.; Green, M.A.; Watson, A.W.; Williams, E.A.; Stevenson, E.J.; Penson, S.; Johnstone, A.M. Protein for Life: Review of Optimal Protein Intake, Sustainable Dietary Sources and the Effect on Appetite in Ageing Adults. Nutrients 2018, 10, 360. [Google Scholar] [CrossRef] [PubMed]
  6. Pikosky, M.A.; Ragalie-Carr, J.; Miller, G.D. Recognizing the importance of protein quality in an era of food systems transformation. Front. Sustain. Food Syst. 2022, 6, 1–7. [Google Scholar] [CrossRef]
  7. Wolfe, R.R.; Cifelli, A.M.; Kostas, G.; Kim, I.-Y. Optimizing Protein Intake in Adults: Interpretation and Application of the Recommended Dietary Allowance Compared with the Acceptable Macronutrient Distribution Range. Adv. Nutr. 2017, 8, 266–275. [Google Scholar] [CrossRef] [PubMed]
  8. FAO; WHO; UNU. Technical Report Series 935. Protein and Amino Acid Requirements in Human Nutrition; WHO Press: Geneva, Switzerland, 2007; pp. 1–265. [Google Scholar]
  9. Institute of Medicine (IOM). Dietary Reference Intakes for Energy, Carbohydrates, Fiber, Fat, Fatty Acids, Cholesterol, Proteins, and Amino Acids; The National Academies Press: Washington, DC, USA, 2005. [Google Scholar]
  10. Naghshi, S.; Sadeghi, O.; Willett, W.C.; Esmaillzadeh, A. Dietary intake of total, animal, and plant proteins and risk of all cause, cardiovascular, and cancer mortality: Systematic review and dose-response meta-analysis of prospective cohort studies. BMJ 2020, 370, m2412. [Google Scholar] [CrossRef] [PubMed]
  11. Henchion, M.; Hayes, M.; Mullen, A.M.; Fenelon, M.; Tiwari, B. Future Protein Supply and Demand: Strategies and Factors Influencing a Sustainable Equilibrium. Foods 2017, 6, 53. [Google Scholar] [CrossRef] [PubMed]
  12. Jin, Q.; Yang, L.; Poe, N.; Huang, H. Integrated processing of plant-derived waste to produce value-added products based on the biorefinery concept. Trends Food Sci. Technol. 2018, 74, 119–131. [Google Scholar] [CrossRef]
  13. Basri, M.S.M.; Shah, N.N.A.K.; Sulaiman, A.; Tawakkal, I.S.M.A.; Nor, M.Z.M.; Ariffin, S.H.; Ghani, N.H.A.; Salleh, F.S.M. Progress in the Valorization of Fruit and Vegetable Wastes: Active Packaging, Biocomposites, By-Products, and Innovative Technologies Used for Bioactive Compound Extraction. Polymers 2021, 13, 3503. [Google Scholar] [CrossRef]
  14. Esparza, I.; Jiménez-Moreno, N.; Bimbela, F.; Ancín-Azpilicueta, C.; Gandía, L.M. Fruit and vegetable waste management: Conventional and emerging approaches. J. Environ. Manag. 2020, 265, 110510. [Google Scholar] [CrossRef]
  15. Rana, A.K. Green approaches in the valorization of plant wastes: Recent insights and future directions. Curr. Opin. Green Sustain. Chem. 2022, 38, 1–9. [Google Scholar] [CrossRef]
  16. Barbarino, E.; Lourenço, S.O. An evaluation of methods for extraction and quantification of protein from marine macro- and microalgae. J. Appl. Phycol. 2005, 17, 447–460. [Google Scholar] [CrossRef]
  17. Fleurence, J.; Le Coeur, C.; Mabeau, S.; Maurice, M.; Landrein, A. Comparison of different extractive procedures for proteins from the edible seaweeds Ulva rigida and Ulva rotundata. J. Appl. Phycol. 1995, 7, 577–582. [Google Scholar] [CrossRef]
  18. Urango, A.C.M.; Strieder, M.M.; Silva, E.K.; Meireles, M.A.A. Thermosonication Process Design for Recovering Bioactive Compounds from Fennel: A Comparative Study with Conventional Extraction Techniques. Appl. Sci. 2021, 11, 12104. [Google Scholar] [CrossRef]
  19. Harnedy, P.A.; FitzGerald, R.J. Extraction of protein from the macroalga Palmaria palmata. LWT—Food Sci. Technol. 2013, 51, 375–382. [Google Scholar] [CrossRef]
  20. Nowosad, K.; Sujka, M.; Pankiewicz, U.; Kowalski, R. The application of PEF technology in food processing and human nutrition. J. Food Sci. Technol. 2021, 58, 397–411. [Google Scholar] [CrossRef] [PubMed]
  21. Lammerskitten, A.; Mykhailyk, V.; Wiktor, A.; Toepfl, S.; Nowacka, M.; Bialik, M.; Czyżewski, J.; Witrowa-Rajchert, D.; Parniakov, O. Impact of pulsed electric fields on physical properties of freeze-dried apple tissue. Innov. Food Sci. Emerg. Technol. 2019, 57, 1–7. [Google Scholar] [CrossRef]
  22. Min, S.; Evrendilek, G.A.; Zhang, H.Q. Pulsed Electric Fields: Processing System, Microbial and Enzyme Inhibition, and shelf life extension of Foods. IEEE Trans. Plasma Sci. 2007, 35, 59–73. [Google Scholar] [CrossRef]
  23. Dziadek, K.; Kopeć, A.; Dróżdż, T.; Kiełbasa, P.; Ostafin, M.; Bulski, K.; Oziembłowski, M. Effect of pulsed electric field treatment on shelf life and nutritional value of apple juice. J. Food Sci. Technol. 2019, 56, 1184–1191. [Google Scholar] [CrossRef] [PubMed]
  24. Odriozola-Serrano, I.; Aguiló-Aguayo, I.; Soliva-Fortuny, R.; Martín-Belloso, O. Pulsed electric fields processing effects on quality and health-related constituents of plant-based foods. Trends Food Sci. Technol. 2013, 29, 98–107. [Google Scholar] [CrossRef]
  25. Kantala, C.; Supasin, S.; Intra, P.; Rattanadecho, P. Evaluation of Pulsed Electric Field and Conventional Thermal Processing for Microbial Inactivation in Thai Orange Juice. Foods 2022, 11, 1102. [Google Scholar] [CrossRef]
  26. Barsotti, L.; Dumay, E.; Mu, T.H.; Diaz, M.D.F.; Cheftel, J.C. Effects of high voltage electric pulses on protein-based food constituents and structures. Trends Food Sci. Technol. 2001, 12, 136–144. [Google Scholar] [CrossRef]
  27. Zulueta, A.; Esteve, M.J.; Frasquet, I.; Frígola, A. Fatty acid profile changes during orange juice-milk beverage processing by high-pulsed electric field. Eur. J. Lipid Sci. Technol. 2007, 109, 25–31. [Google Scholar] [CrossRef]
  28. Wouters, P.C.; Smelt, J.P. Inactivation of microorganisms with pulsed electric fields: Potential for food preservation. Food Biotechnol. 1997, 11, 193–229. [Google Scholar] [CrossRef]
  29. Barsotti, L.; Cheftel, J.C. Food processing by pulsed electric fields. II. Biological aspects. Food Rev. Int. 1999, 15, 181–213. [Google Scholar] [CrossRef]
  30. Jeyamkondan, S.; Jayas, D.S.; Holley, R.A. Pulsed electric field processing of foods: A review. J. Food Prot. 1999, 62, 1088–1096. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Barbosa-Cánovas, G.V.; Swanson, B.G. Engineering aspects of pulsed electric field pasteurization. J. Food Eng. 1995, 25, 261–281. [Google Scholar] [CrossRef]
  32. Ho, S.; Mittal, G.S. High voltage pulsed electrical field for liquid food pasteurization. Food Rev. Int. 2000, 16, 395–434. [Google Scholar] [CrossRef]
  33. Dunn, J.E.; Pearlman, J.S.; La Costa, R. Methods and Apparatus for Extending the Shelf Life of Fluid Food Products. U.S. Patent 4,838,154, 22 September 1987. [Google Scholar]
  34. Dunn, J. Pulsed electric field processing: An overview. In Pulsed Electric Fields in Food Processing, Fundamental Aspects and Applications; Barbosa-Cánovas, G., Zhang, Q.H., Eds.; Technomic Press: Lancaster, PA, USA, 2001; p. 130. [Google Scholar]
  35. Sampedro, F.; Rodrigo, D.; Martínez, A.; Barbosa-Cánovas, G.V.; Rodrigo, M. Review: Application of Pulsed Electric Fields in Egg and Egg Derivatives. Food Sci. Technol. Int. 2006, 12, 397–405. [Google Scholar] [CrossRef]
  36. Góngora-Nieto, M.M.; Sepúlveda, D.R.; Pedrow, P.; Barbosa-Cánovas, G.V.; Swanson, B.G. Food Processing by Pulsed Electric Fields: Treatment Delivery, Inactivation Level, and Regulatory Aspects. LWT—Food Sci. Technol. 2002, 35, 375–388. [Google Scholar] [CrossRef]
  37. Puértolas, E.; Luengo, E.; Álvarez, I.; Raso, J. Improving Mass Transfer to Soften Tissues by Pulsed Electric Fields: Fundamentals and Applications. Annu. Rev. Food Sci. Technol. 2012, 3, 263–282. [Google Scholar] [CrossRef]
  38. Toepfl, S. Pulsed Electric Field food treatment—Scale up from lab to industrial scale. Procedia Food Sci. 2011, 1, 776–779. [Google Scholar] [CrossRef]
  39. Góngora-Nieto, M.M.; Pedrow, P.D.; Swanson, B.G.; Barbosa-Cánovas, G.V. Impact of air bubbles in a dielectric liquid when subjected to high field strengths. Innov. Food Sci. Emerg. Technol. 2003, 4, 57–67. [Google Scholar] [CrossRef]
  40. Mariotti, F. Animal and Plant Protein Sources and Cardiometabolic Health. Adv. Nutr. 2019, 10, S351–S366. [Google Scholar] [CrossRef] [PubMed]
  41. Rifna, E.J.; Misra, N.N.; Dwivedi, M. Recent advances in extraction technologies for recovery of bioactive compounds derived from fruit and vegetable waste peels: A review. Crit. Rev. Food Sci. Nutr. 2023, 63, 719–752. [Google Scholar] [CrossRef]
  42. Ganesh, K.S.; Sridhar, A.; Vishali, S. Utilization of fruit and vegetable waste to produce value-added products: Conventional utilization and emerging opportunities-A review. Chemosphere 2022, 287, 132221. [Google Scholar] [CrossRef]
  43. Mathews, A.D.; Tangirala, A.S.; Kumar, S.; Anandharaj, A.; Rawson, A. Extraction and modification of protein from sesame oil cake by the application of emerging technologies. Food Chem. Adv. 2023, 2, 100326. [Google Scholar] [CrossRef]
  44. Kumoro, A.C.; Alhanif, M.; Wardhani, D.H. A Critical Review on Tropical Fruits Seeds as Prospective Sources of Nutritional and Bioactive Compounds for Functional Foods Development: A Case of Indonesian Exotic Fruits. Int. J. Food Sci. 2020, 2020, 4051475. [Google Scholar] [CrossRef] [PubMed]
  45. Ruschioni, S.; Loreto, N.; Foligni, R.; Mannozzi, C.; Raffaelli, N.; Zamporlini, F.; Pasquini, M.; Roncolini, A.; Cardinali, F.; Osimani, A.; et al. Addition of Olive Pomace to Feeding Substrate Affects Growth Performance and Nutritional Value of Mealworm (Tenebrio Molitor L.) Larvae. Foods 2020, 9, 317. [Google Scholar] [CrossRef]
  46. Herchi, W.; Bahashwan, S.; Sakouhi, F.; Boukhchina, S. Influence of harvest year in the physicochemical properties and antioxidant activity of flaxseed hull oils from Tunisia. Food Sci. Technol. 2015, 35, 175–182. [Google Scholar] [CrossRef]
  47. Liadakis, G.; Kekes, T.; Frakolaki, G.; Giannou, V.; Tzia, C. Ingredients for Food Products in Mejdi Jeguirim. Tomato Processing By-Products; Zorpas, A., Ed.; Academic Press: Cambridge, MA, USA, 2022; pp. 117–148. [Google Scholar]
  48. Wen, C.; Zhang, J.; Duan, Y.; Zhang, H.; Ma, H. A Mini-Review on Brewer’s Spent Grain Protein: Isolation, Physicochemical Properties, Application of Protein, and Functional Properties of Hydrolysates. J. Food Sci. 2019, 84, 3330–3340. [Google Scholar] [CrossRef]
  49. Mustafa, A.F.; Baurhoo, B. Evaluation of dried vegetables residues for poultry: II. Effects of feeding cabbage leaf residues on broiler performance, ileal digestibility and total tract nutrient digestibility. Poult. Sci. 2017, 96, 681–686. [Google Scholar] [CrossRef]
  50. Kaur, M.; Gautam, A.; Kaur, H. Nutritional, techno-functional, structural, and rheological properties of potato peel powder: A valuable biowaste being potential source of dietary fibre and antioxidants in cookie formulation. J. Food Process. Preserv. 2022, 46, 1–13. [Google Scholar] [CrossRef]
  51. Lobo, M.G.; Dorta, E. Utilization and management of horticultural waste in Elhadi. In Postharvest Technology of Perishable Horticultural Commodities; Yahia, M., Ed.; Woodhead Publishing: Cambridge, UK, 2019; pp. 639–666. [Google Scholar]
  52. Wikandari, R.; Nguyen, H.; Millati, R.; Niklasson, C.; Taherzadeh, M.J. Improvement of biogas production from orange peel waste by leaching of limonene. Biomed. Res. Int. 2015, 494182. [Google Scholar] [CrossRef] [PubMed]
  53. Cunha, T.J. Value of Feeds for Horses in Horse Feeding and Nutrition, 2nd ed.; Academic Press: San Diego, CA, USA, 2012; pp. 233–273. [Google Scholar]
  54. Hussain, A.; Kausar, T.; Sehar, S.; Sarwar, A.; Ashraf, A.H.; Jamil, M.A.; Noreen, S.; Rafique, A.; Iftikhar, K.; Quddoos, M.Y.; et al. A Comprehensive review of functional ingredients, especially bioactive compounds present in pumpkin peel, flesh and seeds, and their health benefits. Food Chem. Adv. 2022, 1, 100067. [Google Scholar] [CrossRef]
  55. Zaini, H.M.; Roslan, J.; Saallah, S.; Munsu, E.; Sulaiman, N.S.; Pindi, W. Banana peels as a bioactive ingredient and its potential application in the food industry. J. Funct. Foods 2022, 92, 105054. [Google Scholar] [CrossRef]
  56. Ain, H.B.U.; Tufail, T.; Bashir, S.; Ijaz, N.; Hussain, M.; Ikram, A.; Farooq, M.A.; Saewan, S.A. Nutritional importance and industrial uses of pomegranate peel: A critical review. Food Sci. Nutr. 2023, 11, 2589–2598. [Google Scholar] [CrossRef] [PubMed]
  57. Gómez-García, R.; Campos, D.A.; Oliveira, A.; Aguilar, C.N.; Madureira, A.R.; Pintado, M. A chemical valorisation of melon peels towards functional food ingredients: Bioactives profile and antioxidant properties. Food Chem. 2021, 335, 127579. [Google Scholar] [CrossRef] [PubMed]
  58. Maloney, K.P.; Truong, V.D.; Allen, J.C. Susceptibility of sweet potato (Ipomoea batatas) peel proteins to digestive enzymes. Food Sci. Nutr. 2014, 2, 351–360. [Google Scholar] [CrossRef] [PubMed]
  59. Dotto, J.M.; Chacha, J.S. The potential of pumpkin seeds as a functional food ingredient: A review. Sci. Afr. 2020, 10, e00575. [Google Scholar] [CrossRef]
  60. Apprich, S.; Tirpanalan, O.; Hell, J.; Reisinger, M.; Böhmdorfer, S.; Siebenhandl-Ehn, S.; Novalin, S.; Kneifel, W. Wheat bran-based biorefinery 2: Valorization of products. LWT—Food Sci. Technol. 2014, 56, 222–231. [Google Scholar] [CrossRef]
  61. Kalpanadevi, C.; Muthukumar, S.P.; Govindaraju, K.; Subramanian, R. Rice bran protein: An alternative plant-based protein to ameliorate protein malnourishment. J. Cereal Sci. 2021, 97, 103154. [Google Scholar] [CrossRef]
  62. Kumar, M.; Barbhai, M.D.; Hasan, M.; Punia, S.; Dhumal, S.; Radha; Rais, N.; Chandran, D.; Pandiselvam, R.; Kothakota, A.; et al. Onion (Allium cepa L.) peels: A review on bioactive compounds and biomedical activities. Biomed. Pharmacother. 2022, 146, 112498. [Google Scholar] [CrossRef]
  63. Kamal, H.; Le, C.F.; Salter, A.M.; Ali, A. Extraction of protein from food waste: An overview of current status and opportunities. Compr. Rev. Food Sci. Food Saf. 2021, 20, 2455–2475. [Google Scholar] [CrossRef]
  64. Kumar, M.; Tomar, M.; Punia, S.; Dhakane-Lad, J.; Dhumal, S.; Changan, S.; Senapathy, M.; Berwal, M.K.; Sampathrajan, V.; Sayed, A.A.; et al. Plant-based proteins and their multifaceted industrial applications. LWT 2022, 154, 112620. [Google Scholar] [CrossRef]
  65. Ayimbila, F.; Keawsompong, S. Nutritional Quality and Biological Application of Mushroom Protein as a Novel Protein Alternative. Curr. Nutr. Rep. 2023, 12, 290–307. [Google Scholar] [CrossRef]
  66. Zhou, J.; Chen, M.; Wu, S.; Liao, X.; Wang, J.; Wu, Q.; Zhuang, M.; Ding, Y. A review on mushroom-derived bioactive peptides: Preparation and biological activities. Food Res. Int. 2020, 134, 109230. [Google Scholar] [CrossRef]
  67. Meenu, M.; Xu, B. Application of vibrational spectroscopy for classification, authentication and quality analysis of mushroom: A concise review. Food Chem. 2019, 289, 545–557. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, X.M.; Zhang, J.; Wu, L.H.; Zhao, Y.L.; Li, T.; Li, J.Q.; Wang, Y.Z.; Liu, H.G. A mini-review of chemical composition and nutritional value of edible wild-grown mushroom from China. Food Chem. 2014, 151, 279–285. [Google Scholar] [CrossRef] [PubMed]
  69. Elhusseiny, S.M.; El-Mahdy, T.S.; Awad, M.F.; Elleboudy, N.S.; Farag, M.M.S.; Aboshanab, K.M.; Yassien, M.A. Antiviral, Cytotoxic, and Antioxidant Activities of Three Edible Agaricomycetes Mushrooms: Pleurotus columbinus, Pleurotus sajor-caju, and Agaricus bisporus. J. Fungi—Open Access Mycol. J. 2021, 7, 645. [Google Scholar] [CrossRef]
  70. González, A.; Nobre, C.; Simões, L.S.; Cruz, M.; Loredo, A.; Rodríguez-Jasso, R.M.; Contreras, J.; Texeira, J.; Belmares, R. Evaluation of functional and nutritional potential of a protein concentrate from Pleurotus ostreatus mushroom. Food Chem. 2021, 346, 128884. [Google Scholar] [CrossRef]
  71. Miles, P.G.; Chang, S.-T. Mushrooms: Cultivation, Nutritional Value, Medicinal Effect, and Environmental Impact, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar] [CrossRef]
  72. Rangsinth, P.; Sillapachaiyaporn, C.; Nilkhet, S.; Tencomnao, T.; Ung, A.T.; Chuchawankul, S. Mushroom-derived bioactive compounds potentially serve as the inhibitors of SARS-CoV-2 main protease: An in silico approach. J. Tradit. Complement. Med. 2021, 11, 158–172. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, M.; Belwal, T.; Devkota, H.P.; Li, L.; Luo, Z. Trends of utilizing mushroom polysaccharides (MPs) as potent nutraceutical components in food and medicine: A comprehensive review. Trends Food Sci. Technol. 2019, 92, 94–110. [Google Scholar] [CrossRef]
  74. Lu, X.; Brennan, M.A.; Serventi, L.; Liu, J.; Guan, W.; Brennan, C.S. Addition of mushroom powder to pasta enhances the antioxidant content and modulates the predictive glycaemic response of pasta. Food Chem. 2018, 264, 199–209. [Google Scholar] [CrossRef]
  75. Venturella, G.; Ferraro, V.; Cirlincione, F.; Gargano, M.L. Medicinal Mushrooms: Bioactive Compounds, Use, and Clinical Trials. Int. J. Mol. Sci. 2021, 22, 634. [Google Scholar] [CrossRef]
  76. Das, A.K.; Nanda, P.K.; Dandapat, P.; Bandyopadhyay, S.; Gullón, P.; Sivaraman, G.K.; McClements, D.J.; Gullón, B.; Lorenzo, J.M. Edible Mushrooms as Functional Ingredients for Development of Healthier and More Sustainable Muscle Foods: A Flexitarian Approach. Molecules 2021, 26, 2463. [Google Scholar] [CrossRef]
  77. Jacinto-Azevedo, B.; Valderrama, N.; Henríquez, K.; Aranda, M.; Aqueveque, P. Nutritional value and biological properties of Chilean wild and commercial edible mushrooms. Food Chem. 2021, 356, 129651. [Google Scholar] [CrossRef] [PubMed]
  78. Ghose, A.; Mitra, S. Spent waste from edible mushrooms offers innovative strategies for the remediation of persistent organic micropollutants: A review. Environ. Pollut. 2022, 305, 119285. [Google Scholar] [CrossRef] [PubMed]
  79. Atallah, E.; Zeaiter, J.; Ahmad, M.N.; Leahy, J.J.; Kwapinski, W. Hydrothermal carbonization of spent mushroom compost waste compared against torrefaction and pyrolysis. Fuel Process. Technol. 2021, 216, 106795. [Google Scholar] [CrossRef]
  80. Leong, Y.K.; Ma, T.-W.; Chang, J.-S.; Yang, F.-C. Recent advances and future directions on the valorization of spent mushroom substrate (SMS): A review. Bioresour. Technol. 2021, 344, 126157. [Google Scholar] [CrossRef] [PubMed]
  81. Usman, M.; Murtaza, G.; Ditta, A. Nutritional, Medicinal, and Cosmetic Value of Bioactive Compounds in Button Mushroom (Agaricus bisporus): A Review. Appl. Sci. 2021, 11, 5943. [Google Scholar] [CrossRef]
  82. Carrasco-González, J.A.; Serna-Saldívar, S.O.; Gutiérrez-Uribe, J.A. Nutritional composition and nutraceutical properties of the Pleurotus fruiting bodies: Potential use as food ingredient. J. Food Compos. Anal. 2017, 58, 69–81. [Google Scholar] [CrossRef]
  83. Abidin, M.H.Z.; Abdullah, N.; Abidin, N.Z. Therapeutic properties of Pleurotus species (oyster mushrooms) for atherosclerosis: A review. Int. J. Food Prop. 2017, 20, 1251–1261. [Google Scholar] [CrossRef]
  84. Valverde, M.E.; Hernández-Pérez, T.; Paredes-López, O.; Hernández-Pérez, T.; Paredes-López, O. Edible Mushrooms: Improving Human Health and Promoting Quality Life. Int. J. Microbiol. 2015, 2015, 376387. [Google Scholar] [CrossRef]
  85. Li, Y.; Chen, H.; Zhang, X. Cultivation, nutritional value, bioactive compounds of morels, and their health benefits: A systematic review. Front. Nutr. 2023, 10, 1159029. [Google Scholar] [CrossRef]
  86. Zhang, Q.; Wu, C.; Fan, G.; Li, T.; Sun, Y. Improvement of antioxidant activity of Morchella esculenta protein hydrolysate by optimized glycosylation reaction. CyTA—J. Food 2018, 16, 238–246. [Google Scholar] [CrossRef]
  87. Das, S.; Prakash, B. Edible mushrooms: Nutritional composition and medicinal benefits for improvement in quality life. In Research and Technological Advances in Food Science; Prakash, B., Ed.; Academic Press: Cambridge, MA, USA, 2022; pp. 269–300. [Google Scholar]
  88. Spim, S.R.V.; Castanho, N.R.C.M.; Pistila, A.M.H.; Jozala, A.F.; Oliveira Júnior, J.M.; Grotto, D. Lentinula edodes mushroom as an ingredient to enhance the nutritional and functional properties of cereal bars. J. Food Sci. Technol. 2021, 58, 1349–1357. [Google Scholar] [CrossRef]
  89. Bisen, P.S.; Baghel, R.K.; Sanodiya, B.S.; Thakur, G.S.; Prasad, G.B. Lentinus edodes: A macrofungus with pharmacological activities. Curr. Med. Chem. 2010, 17, 2419–2430. [Google Scholar] [CrossRef]
  90. Sun, X.; Yang, C.; Ma, Y.; Zhang, J.; Wang, L. Research progress of Auricularia heimuer on cultivation physiology and molecular biology. Front. Microbiol. 2022, 13, 148249. [Google Scholar] [CrossRef]
  91. Hrudayanath, T.; Sameer, K.S. Diversity, nutritional composition and medicinal potential of Indian mushrooms: A review. Afr. J. Biotechnol. 2014, 13, 523–545. [Google Scholar] [CrossRef]
  92. Tang, C.; Hoo, P.C.; Tan, L.T.; Pusparajah, P.; Khan, T.M.; Lee, L.H.; Goh, B.H.; Chan, K.G. Golden Needle Mushroom: A Culinary Medicine with Evidenced-Based Biological Activities and Health Promoting Properties. Front. Pharmacol. 2016, 7, 474. [Google Scholar] [CrossRef] [PubMed]
  93. Chen, J.; Li, J.M.; Tang, Y.J.; Ma, K.; Li, B.; Zeng, X.; Liu, X.-B.; Li, Y.; Yang, Z.-L.; Xu, W.-N.; et al. Genome-wide analysis and prediction of genes involved in the biosynthesis of polysaccharides and bioactive secondary metabolites in high-temperature-tolerant wild Flammulina filiformis. BMC Genom. 2020, 21, 719. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, L.; Chen, J.; Xu, F.; Han, R.; Quan, M.; Wang, L. Effect of Tremella fuciformis on dough structure and rheology, noodle flavor, and quality characteristics. LWT 2022, 172, 114180. [Google Scholar] [CrossRef]
  95. Shahrajabian, M.H.; Sun, W.; Cheng, Q.; Khoshkharam, M. Exploring the quality of foods from ancient China based on traditional Chinese medicine. In Functional Foods and Nutraceuticals in Metabolic and Non-Communicable Diseases; Singh, R.B., Watanabe, S., Isaza, A.A., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 87–105. [Google Scholar] [CrossRef]
  96. Rajarathnam, S.; Shashirekha, M.N. Mushrooms and Truffles|Use of Wild Mushrooms. In Encyclopaedia of Food Sciences and Nutrition, 2nd ed.; Academic Press: Cambridge, MA, USA, 2003; pp. 4048–4054. [Google Scholar] [CrossRef]
  97. Mikola, E.; Geösel, A.; Stefanovits-Bányai, É.; Fodor, M. Quantitative determination of macro components and classification of some cultivated mushrooms using near-infrared spectroscopy. J. Food Process Preserv. 2020, 44, e14540. [Google Scholar] [CrossRef]
  98. Chauhan, G.; Prasad, S.; Himanshi, R.; Satyawati, S. Nutritional profiling and value addition of products from Hypsizygus tessellatus. Food Biol. 2017, 6, 1–6. [Google Scholar] [CrossRef]
  99. Li, W.; Chen, W.; Ma, H.; Wang, J.; Li, Z.; Wang, Q.; Zhang, Z.; Wu, D.; Zhang, J.; Yang, Y. Study on the relationship between structure and taste activity of the umami peptide of Stropharia rugosoannulata prepared by ultrasound. Ultrason. Sonochem. 2022, 90, 106206. [Google Scholar] [CrossRef] [PubMed]
  100. Hu, Y.; Kakumyan, P.; Bandara, A.; Peter, M. The nutrition, cultivation and biotechnology of Stropharia rugosoannulata. Fungal Biotec. 2021, 1, 13–25. [Google Scholar] [CrossRef]
  101. Petrović, J.; Glamočlija, J.; Stojković, D.; Ćirić, A.; Barros, L.; Ferreira, I.C.F.R.; Soković, M. Nutritional value, chemical composition, antioxidant activity and enrichment of cream cheese with chestnut mushroom Agrocybe aegerita (Brig.) Sing. J. Food Sci. Technol. 2015, 52, 6711–6718. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, M.; Gu, B.; Huang, J.; Jiang, S.; Chen, Y.; Yin, Y.; Pan, Y.; Yu, G.; Li, Y.; Wong, B.H.C.; et al. Transcriptome and Proteome Exploration to Provide a Resource for the Study of Agrocybe aegerita. PLoS ONE 2013, 8, e56686. [Google Scholar] [CrossRef]
  103. Bandura, I.; Kulyk, A.; Makohon, S.; Khareba, O.; Khareba, V. Influence of the substrate composition on the yield and nutritional value of the fruiting bodies of the edible mushrooms Pleurotus citrinopileatus and Cyclocybe aegerita. Plant Var. Stud. Prot. 2021, 17, 130–138. [Google Scholar] [CrossRef]
  104. Friedman, M. Chemistry, Nutrition, and Health-Promoting Properties of Hericium erinaceus (Lion’s Mane) Mushroom Fruiting Bodies and Mycelia and Their Bioactive Compounds. J. Agric. Food Chem. 2015, 63, 7108–7123. [Google Scholar] [CrossRef] [PubMed]
  105. Chaiyasut, C.; Sivamaruthi, B.S. Anti-hyperglycemic property of Hericium erinaceus—A mini review. Asian Pac. J. Trop. Biomed. 2017, 7, 1036–1040. [Google Scholar] [CrossRef]
  106. Sitinjak, R. The Nutritional Content of the Mushroom Phallus indusiatus Vent., which Grows in the Cocoa Plantation, Gaperta-Ujung, Medan. Der Pharma Chem. 2017, 9, 44–47. [Google Scholar]
  107. Roselló-Soto, E.; Barba, F.J.; Parniakov, O.; Galanakis, C.M.; Lebovka, N.; Grimi, N.; Vorobiev, E. High Voltage Electrical Discharges, Pulsed Electric Field, and Ultrasound Assisted Extraction of Protein and Phenolic Compounds from Olive Kernel. Food Bioprocess. Technol. 2015, 8, 885–894. [Google Scholar] [CrossRef]
  108. Antunes, F.; Marçal, S.; Taofiq, O.; Morais, A.M.M.B.; Freitas, A.C.; Ferreira, I.C.F.R.; Pintado, M. Valorization of Mushroom By-Products as a Source of Value-Added Compounds and Potential Applications. Molecules 2020, 25, 2672. [Google Scholar] [CrossRef] [PubMed]
  109. Jayachandran, M.; Xiao, J.; Xu, B.A. Critical Review on Health Promoting Benefits of Edible Mushrooms through Gut Microbiota. Int. J. Mol. Sci. 2017, 18, 1934. [Google Scholar] [CrossRef] [PubMed]
  110. Welti-Chanes, J.; Vergara-Balderas, F.; Bermúdez-Aguirre, D. Transport phenomena in food engineering: Basic concepts and advances. J. Food Eng. 2005, 67, 113–128. [Google Scholar] [CrossRef]
  111. Barba, F.J.; Parniakov, O.; Pereira, S.A.; Wiktor, A.; Grimi, N.; Boussetta, N.; Saraiva, J.A.; Raso, J.; Martin-Belloso, O.; Witrowa-Rajchert, D.; et al. Current applications and new opportunities for the use of pulsed electric fields in food science and industry. Food Res. Int. 2015, 77, 773–798. [Google Scholar] [CrossRef]
  112. Barba, F.J.; Zhu, Z.; Koubaa, M.; Sant’Ana, A.S.; Orlien, V. Green alternative methods for the extraction of antioxidant bioactive compounds from winery wastes and by-products: A review. Trends Food Sci. Technol. 2016, 49, 96–109. [Google Scholar] [CrossRef]
  113. Zhang, Q.W.; Lin, L.G.; Ye, W.C. Techniques for extraction and isolation of natural products: A comprehensive review. Chin. Med. 2018, 13, 20. [Google Scholar] [CrossRef]
  114. Chua, L.S. A review on plant-based rutin extraction methods and its pharmacological activities. J. Ethnopharmacol. 2013, 150, 805–817. [Google Scholar] [CrossRef]
  115. Chandran, A.S.; Suri, S.; Choudhary, P. Sustainable plant protein: An up-to-date overview of sources, extraction techniques and utilization. Sustain. Food Technol. 2023, 1, 466–483. [Google Scholar] [CrossRef]
  116. Nadar, S.S.; Rao, P.; Rathod, V.K. Enzyme assisted extraction of biomolecules as an approach to novel extraction technology: A review. Food Res. Int. 2018, 108, 309–330. [Google Scholar] [CrossRef]
  117. Chakka, A.K.; Sriraksha, M.S.; Ravishankar, C.N. Sustainability of emerging green non-thermal technologies in the food industry with food safety perspective: A review. LWT 2021, 151, 112140. [Google Scholar] [CrossRef]
  118. Soquetta, M.B.; Terra, L.d.M.; Bastos, C.P. Green technologies for the extraction of bioactive compounds in fruits and vegetables. CyTA—J. Food 2018, 16, 400–412. [Google Scholar] [CrossRef]
  119. Xue, D.; Farid, M.F. Pulsed Electric Field Extraction of Valuable Compounds from White Button Mushroom (Agaricus Bisporus). Innov. Food Sci. Emerg. Technol. 2015, 29, 178–186. [Google Scholar] [CrossRef]
  120. Jadhav, H.B.; Annapure, U.S.; Deshmukh, R.R. Non-thermal Technologies for Food Processing. Front. Nutr. 2021, 8, 657090. [Google Scholar] [CrossRef] [PubMed]
  121. Galanakis, C.M.; Schieber, A. Editorial. Food Res. Int. 2014, 65, 299–300. [Google Scholar] [CrossRef]
  122. Reddy, V.Y.; Koruth, J.; Jais, P.; Petru, J.; Timko, F.; Skalsky, I.; Hebeler, R.; Labrousse, L.; Barandon, L.; Kralovec, S.; et al. Ablation of Atrial Fibrillation with Pulsed Electric Fields: An Ultra-Rapid, Tissue-Selective Modality for Cardiac Ablation. JACC Clin. Electrophysiol. 2018, 4, 987–995. [Google Scholar] [CrossRef]
  123. Nuccitelli, R. Application of Pulsed Electric Fields to Cancer Therapy. Bioelectricity 2019, 1, 30–34. [Google Scholar] [CrossRef]
  124. Buchmann, L.; Brändle, I.; Haberkorn, I.; Hiestand, M.; Mathys, A. Pulsed electric field based cyclic protein extraction of microalgae towards closed-loop biorefinery concepts. Bioresour. Technol. 2019, 291, 121870. [Google Scholar] [CrossRef]
  125. Rios-Corripio, G.; Morales-de la Peña, M.; Welti-Chanes, J.; Ángel Guerrero-Beltrán, J. Pulsed electric field processing of a pomegranate (Punica granatum L.) fermented beverage. Innov. Food Sci. Emerg. Technol. 2022, 79, 103045. [Google Scholar] [CrossRef]
  126. Naliyadhara, N.; Kumar, A.; Girisa, S.; Daimary, U.D.; Hegde, M.; Kunnumakkara, A.B. Pulsed electric field (PEF): Avant-garde extraction escalation technology in food industry. Trends Food Sci. Technol. 2022, 122, 238–255. [Google Scholar] [CrossRef]
  127. Punthi, F.; Yudhistira, B.; Gavahian, M.; Chang, C.-K.; Cheng, K.-C.; Hou, C.-Y.; Hsieh, C.-W. Pulsed electric field-assisted drying: A review of its underlying mechanisms, applications, and role in fresh produce plant-based food preservation. Compr. Rev. Food Sci. Food Saf. 2022, 21, 5109–5130. [Google Scholar] [CrossRef]
  128. Navarro, A.; Ruiz-Méndez, M.-V.; Sanz, C.; Martínez, M.; Rego, D.; Pérez, A.G. Application of Pulsed Electric Fields to Pilot and Industrial Scale Virgin Olive Oil Extraction: Impact on Organoleptic and Functional Quality. Foods 2022, 11, 2022. [Google Scholar] [CrossRef] [PubMed]
  129. Parniakov, O.; Barba, F.J.; Grimi, N.; Lebovka, N.; Vorobiev, E. Extraction assisted by pulsed electric energy as a potential tool for green and sustainable recovery of nutritionally valuable compounds from mango peels. Food Chem. 2016, 192, 842–848. [Google Scholar] [CrossRef] [PubMed]
  130. Käferböck, A.; Smetana, S.; de Vos, R.; Schwarz, C.; Toepfl, S.; Parniakov, O. Sustainable extraction of valuable components from Spirulina assisted by pulsed electric fields technology. Algal Res. 2020, 48, 101914. [Google Scholar] [CrossRef]
  131. Kumari, B.; Tiwari, B.K.; Hossain, M.B.; Brunton, N.P.; Rai, D.K. Recent Advances on Application of Ultrasound and Pulsed Electric Field Technologies in the Extraction of Bioactives from Agro-Industrial By-products. Food Bioprocess. Technol. 2018, 11, 223–241. [Google Scholar] [CrossRef]
  132. Saldaña, G.; Puértolas, E.; Álvarez, I.; Meneses, N.; Knorr, D.; Raso, J. Evaluation of a static treatment chamber to investigate kinetics of microbial inactivation by pulsed electric fields at different temperatures at quasi-isothermal conditions. J. Food Eng. 2010, 100, 349–356. [Google Scholar] [CrossRef]
  133. Golberg, A.; Sack, M.; Teissie, J.; Pataro, G.; Pliquett, U.; Saulis, G.; Stefan, T.; Miklavcic, D.; Vorobiev, E.; Frey, W. Energy-efficient biomass processing with pulsed electric fields for bioeconomy and sustainable development. Biotechnol. Biofuels 2016, 9, 1–22. [Google Scholar] [CrossRef]
  134. Weaver, J.C. Electroporation of cells and tissues. IEEE Trans. Plasma Sci. 2000, 28, 24–33. [Google Scholar] [CrossRef]
  135. Wang, M.-S.; Wang, L.-H.; Bekhit, A.E.-D.A.; Yang, J.; Hou, Z.-P.; Wang, Y.-Z.; Dai, Q.-Z.; Zeng, X.-A. A review of sublethal effects of pulsed electric field on cells in food processing. J. Food Eng. 2018, 223, 32–41. [Google Scholar] [CrossRef]
  136. Napotnik, T.B.; Polajžer, T.; Miklavčič, D. Cell death due to electroporation—A review. Bioelectrochemistry 2021, 141, 107871. [Google Scholar] [CrossRef]
  137. Boussetta, N.; Lesaint, O.; Vorobiev, E. A study of mechanisms involved during the extraction of polyphenols from grape seeds by pulsed electrical discharges. Innov. Food Sci. Emerg. Technol. 2013, 19, 124–132. [Google Scholar] [CrossRef]
  138. Redondo, D.; Venturini, M.E.; Luengo, E.; Raso, J.; Arias, E. Pulsed electric fields as a green technology for the extraction of bioactive compounds from thinned peach by-products. Innov. Food Sci. Emerg. Technol. 2018, 45, 335–343. [Google Scholar] [CrossRef]
  139. Sánchez-Vega, R.; Elez-Martínez, P.; Martín-Belloso, O. Influence of high-intensity pulsed electric field processing parameters on antioxidant compounds of broccoli juice. Innov. Food Sci. Emerg. Technol. 2015, 29, 70–77. [Google Scholar] [CrossRef]
  140. Carullo, D.; Abera, B.D.; Casazza, A.A.; Donsì, F.; Perego, P.; Ferrari, G.; Pataro, G. Effect of pulsed electric fields and high pressure homogenization on the aqueous extraction of intracellular compounds from the microalgae Chlorella vulgaris. Algal Res. 2018, 31, 60–69. [Google Scholar] [CrossRef]
  141. Postma, P.R.; Pataro, G.; Capitoli, M.; Barbosa, M.J.; Wijffels, R.H.; Eppink, M.H.M.; Olivieri, G.; Ferrari, G. Selective extraction of intracellular components from the microalga Chlorella vulgaris by combined pulsed electric field–temperature treatment. Bioresour. Technol. 2016, 203, 80–88. [Google Scholar] [CrossRef] [PubMed]
  142. Carullo, D.; Donsì, F.; Ferrari, G.; Pataro, G. Extraction improvement of water-soluble compounds from Arthrospira platensis through the combination of high-shear homogenization and pulsed electric fields. Algal Res. 2021, 57, 102341. [Google Scholar] [CrossRef]
  143. Calleja-Gómez, M.; Castagnini, J.M.; Carbó, E.; Ferrer, E.; Berrada, H.; Barba, F.J. Evaluation of Pulsed Electric Field-Assisted Extraction on the Microstructure and Recovery of Nutrients and Bioactive Compounds from Mushroom (Agaricus bisporus). Separations 2022, 9, 302. [Google Scholar] [CrossRef]
  144. Moreira, S.A.; Alexandre, E.M.; Pintado, M.; Saraiva, J.A. Effect of emergent non-thermal extraction technologies on bioactive individual compounds profile from different plant materials. Food Res. Int. 2019, 115, 177–190. [Google Scholar] [CrossRef]
  145. Thongkong, S.; Klangpetch, W.; Unban, K.; Tangjaidee, P.; Phimolsiripol, Y.; Rachtanapun, P.; Jantanasakulwong, K.; Schönlechner, R.; Thipchai, P.; Phongthai, S. Impacts of Electroextraction Using the Pulsed Electric Field on Properties of Rice Bran Protein. Foods 2023, 12, 835. [Google Scholar] [CrossRef]
  146. Zhao, W.; Yang, R. Pulsed Electric Field Induced Aggregation of Food Proteins: Ovalbumin and Bovine Serum Albumin. Food Bioprocess. Technol. 2012, 5, 1706–1714. [Google Scholar] [CrossRef]
  147. Liu, Y.Y.; Zeng, X.A.; Deng, Z.; Yu, S.J.; Yamasaki, S. Effect of pulsed electric field on the secondary structure and thermal properties of soy protein isolate. Eur. Food Res. Technol. 2011, 233, 841–850. [Google Scholar] [CrossRef]
  148. Wu, L.; Zhao, W.; Yang, R.; Chen, X. Effects of pulsed electric fields processing on stability of egg white proteins. J. Food Eng. 2014, 139, 13–18. [Google Scholar] [CrossRef]
  149. Xiang, B.Y.; Simpson, M.V.; Ngadi, M.O.; Simpson, B.K. Effect of pulsed electric field on the rheological and colour properties of soy milk. Int. J. Food Sci. Nutr. 2011, 62, 787–793. [Google Scholar] [CrossRef] [PubMed]
  150. Xiang, B.Y.; Ngadi, M.O.; Ochoa-Martinez, L.A.; Simpson, M.V. Pulsed Electric Field-Induced Structural Modification of Whey Protein Isolate. Food Bioprocess Technol. 2011, 4, 1341–1348. [Google Scholar] [CrossRef]
  151. Cullen, P.J.; Tiwari, B.K.; Valdramidis, V. Novel Thermal and Non-Thermal Technologies for Fluid Foods; Academic Press: Cambridge, MA, USA, 2011; pp. 65–66. [Google Scholar] [CrossRef]
  152. El-Hag, A.H.; Rodriguez Gonzalez, O.; Jayaram, S.H.; Griffiths, M.W. A Performance Study of a Multilevel Electrode Treatment Chamber for Food Processing. IEEE Trans. Ind. Appl. 2013, 49, 1091–1097. [Google Scholar] [CrossRef]
  153. Huang, K.; Wang, J. Designs of Pulsed Electric Fields Treatment Chambers for Liquid Foods Pasteurization Process: A Review. J. Food Eng. 2009, 95, 227–239. [Google Scholar] [CrossRef]
  154. Qin, S.; Timoshkin, I.V.; Maclean, M.; Wilson, M.P.; Given, M.J.; Wang, T.; Anderson, J.G.; Macgregor, S.J. Pulsed electric field treatment of saccharomyces cerevisiae using different waveforms. IEEE Trans. Dielectr. Electr. Insul. 2015, 22, 1841–1848. [Google Scholar] [CrossRef]
  155. Qin, S.; Timoshkin, I.V.; Maclean, M.; Wilson, M.P.; Given, M.J.; Wang, T.; Anderson, J.G.; Macgregor, S.J. TiO2-Coated Electrodes for Pulsed Electric Field Treatment of Microorganisms. IEEE Trans. Plasma Sci. 2016, 44, 2121–2128. [Google Scholar] [CrossRef]
  156. Achayuthakan, P.; Wongsagonsup, R.; Sriprablom, J.; Suphantharika, M.; Intra, P. Effect of Pulsed Electric Field Treatment on the Protein, Digestibility, and Physicochemical Properties of Starch Granules in Wheat Flour. Polymers 2023, 15, 4087. [Google Scholar] [CrossRef]
  157. Raso, J.; Frey, W.; Ferrari, G.; Pataro, G.; Knorr, D.; Teissie, J.; Miklavčič, D. Recommendations guidelines on the key information to be reported in studies of application of PEF technology in food and biotechnological processes. Innov. Food Sci. Emerg. Technol. 2016, 37, 312–321. [Google Scholar] [CrossRef]
  158. Andreou, V.; Dimopoulos, G.; Dermesonlouoglou, E.; Taoukis, P. Application of pulsed electric fields to improve product yield and waste valorization in industrial tomato processing. J. Food Eng. 2020, 270, 109778. [Google Scholar] [CrossRef]
  159. Parniakov, O.; Roselló-Soto, E.; Barba, F.J.; Grimi, N.; Lebovka, N.; Vorobiev, E. New approaches for the effective valorization of papaya seeds: Extraction of proteins, phenolic compounds, carbohydrates, and isothiocyanates assisted by pulsed electric energy. Food Res. Int. 2015, 77, 711–717. [Google Scholar] [CrossRef]
  160. Qin, B.L.; Zhang, Q.; Canovas, G.V.B. Inactivation of microorganisms by Pulsed Electric Fields of different voltage waveforms. IEEE Trans. Dielectr. Electr. Insul. 1994, 1, 1047–1057. [Google Scholar] [CrossRef]
  161. Canovas, G.V.B.; Tapia, M.S.; Cano, M.P. Novel Food Processing Technologies; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  162. Brito, P.S.; Canacsinh, H.; Mendes, J.P.; Redondo, L.M.; Pereira, M.T. Comparison between monopolar and bipolar microsecond range Pulsed Electric Fields in enhancement of apple juice extraction. IEEE Trans. Plasma Sci. 2012, 40, 2348–2354. [Google Scholar] [CrossRef]
  163. Sobrino-López, Á.; Raybaudi-Massilia, R.; Martín-Belloso, O. High intensity pulsed electric field variables affecting Staphylococcus aureus inoculated in milk. J. Dairy Sci. 2006, 89, 3739–3748. [Google Scholar] [CrossRef] [PubMed]
  164. Qin, B.L.; Barbosa-Canovas, G.V.B.; Swanson, B.G.; Pedrow, P.D.; Olsen, R.G. Inactivating microorganisms using a Pulsed Electric Field continuous treatment system. IEEE Trans. Ind. Appl. 1998, 34, 43–50. [Google Scholar] [CrossRef]
  165. Parniakov, O.; Barba, F.J.; Grimi, N.; Lebovka, N.; Vorobiev, E. Impact of pulsed electric fields and high voltage electrical discharges on extraction of high-added value compounds from papaya peels. Food Res. Int. 2014, 65, 337–343. [Google Scholar] [CrossRef]
  166. Qin, B.; Zhang, Q.; Barbosa-Canovas, G.V.; Swanson, B.G.; Pedrow, P.D. Pulsed electric field treatment chamber design for liquid food pasteurization using a finite element method. Trans. ASAE 1995, 38, 557–565. [Google Scholar] [CrossRef]
  167. Ghoshal, G. Comprehensive review on pulsed electric field in food preservation: Gaps in current studies for potential future research. Heliyon 2023, 9, e17532. [Google Scholar] [CrossRef]
  168. Tamborrino, A.; Mescia, L.; Taticchi, A.; Berardi, A.; Lamacchia, C.M.; Leone, A.; Servili, M. Continuous pulsed electric field pilot plant for olive oil extraction process. Innov. Food Sci. Emerg. Technol. 2022, 82, 103192. [Google Scholar] [CrossRef]
  169. Schottroff, F.; Knappert, J.; Eppmann, P.; Krottenthaler, A.; Horneber, T.; McHardy, C.; Rauh, C.; Jaeger, H. Development of a Continuous Pulsed Electric Field (PEF) Vortex-Flow Chamber for Improved Treatment Homogeneity Based on Hydrodynamic Optimization. Front. Bioeng. Biotechnol. 2020, 8, 340. [Google Scholar] [CrossRef] [PubMed]
  170. Arshad, R.N.; Abdul Malek, Z.; Munir, A.; Ahmad, M.H.; Nawawi, Z.; Sidik, M.A.B.; Jumani, T.A.; Khan, I.; Alotabi, H.; Khan, A. An improved electroporator with continuous liquid flow and double-exponential waveform for liquid food pasteurization. IEEE Access. 2021, 9, 147732–147742. [Google Scholar] [CrossRef]
  171. Seshakamal, J.H.; Kameswara, L.R. Electric Field Fluid Treatment Chamber. CA Patent CA2560858C, 22 September 2006. [Google Scholar]
  172. Rego, D.; Costa, L.; Pereira, M.T.; Redondo, L.M. Cell Membrane Permeabilization Studies of Chlorella sp. by Pulsed Electric Fields. IEEE Trans. Plasma Sci. 2015, 43, 3483–3488. [Google Scholar] [CrossRef]
  173. Evrendilek, G.A.; Li, S.; Dantzer, W.R.; Zhang, Q.H. Pulsed Electric Field Processing of Beer: Microbial, Sensory, and Quality Analyses. J. Food Sci. 2004, 69, M228–M232. [Google Scholar] [CrossRef]
  174. Donsì, G.; Ferrari, G.; Pataro, G. Inactivation kinetics of Saccharomyces cerevisiae by pulsed electric fields in a batch treatment chamber: The effect of electric field unevenness and initial cell concentration. J. Food Eng. 2007, 78, 784–792. [Google Scholar] [CrossRef]
  175. Toepfl, S.; Volker, H.; Dietrich, K. Applications of pulsed electric fields technology for the food industry. In Pulsed Electric Fields Technology for the Food Industry; Springer: Berlin/Heidelberg, Germany, 2006; pp. 197–221. [Google Scholar]
  176. Kotnik, T.; Kramar, P.; Pucihar, G.; Miklavcic, D.; Tarek, M. Cell membrane electroporation-Part 1: The phenomenon. IEEE Electr. Insul. Mag. 2012, 28, 14–23. [Google Scholar] [CrossRef]
  177. Sugar, I.P.; Neumann, E. Stochastic model for electric field-induced membrane pores electroporation. Biophys. Chem. 1984, 19, 211–225. [Google Scholar] [CrossRef]
  178. Kandušer, M.; Belič, A.; Čorović, S.; Škrjanc, I. Modular Serial Flow Through device for pulsed electric field treatment of the liquid samples. Sci. Rep. 2017, 7, 8115. [Google Scholar] [CrossRef]
  179. Sale, A.J.H.; Hamilton, W.A. Effect of high electric fields on microorganisms. I. Killing of bacteria and yeast. Biochim. Biophys. Acta. 1967, 148, 781–788. [Google Scholar] [CrossRef]
  180. Saulis, G. Electroporation of Cell Membranes: The Fundamental Effects of Pulsed Electric Fields in Food Processing. Food Eng. Reviews. 2010, 2, 52–73. [Google Scholar] [CrossRef]
  181. Martí-Quijal, F.J.; Ramon-Mascarell, F.; Pallarés, N.; Ferrer, E.; Berrada, H.; Phimolsiripol, Y.; Barba, F.J. Extraction of Antioxidant Compounds and Pigments from Spirulina (Arthrospira platensis) Assisted by Pulsed Electric Fields and the Binary Mixture of Organic Solvents and Water. Appl. Sci. 2021, 11, 7629. [Google Scholar] [CrossRef]
  182. Vernon-Parry, K.D. Scanning Electron Microscopy: An Introduction. III-Vs Rev. 2000, 13, 40–44. [Google Scholar] [CrossRef]
  183. Sarkis, J.R.; Boussetta, N.; Blouet, C.; Tessaro, I.C.; Marczak, L.D.F.; Vorobiev, E. Effect of pulsed electric fields and high voltage electrical discharges on polyphenol and protein extraction from sesame cake. Innov. Food Sci. Emerg. Technol. 2015, 29, 170–177. [Google Scholar] [CrossRef]
  184. Jaeschke, D.P.; Mercali, G.D.; Marczak, L.D.F.; Müller, G.; Frey, W.; Gusbeth, C. Extraction of valuable compounds from Arthrospira platensis using pulsed electric field treatment. Bioresour. Technol. 2019, 283, 207–212. [Google Scholar] [CrossRef] [PubMed]
  185. Gateau, H.; Blanckaert, V.; Veidl, B.; Burlet-Schiltz, O.; Pichereaux, C.; Gargaros, A.; Marchand, J.; Schoefs, B. Application of pulsed electric fields for the biocompatible extraction of proteins from the microalga Haematococcus pluvialis. Bioelectrochemistry 2021, 137, 107588. [Google Scholar] [CrossRef] [PubMed]
  186. Yu, X.; Gouyo, T.; Grimi, N.; Bals, O.; Vorobiev, E. Pulsed electric field pretreatment of rapeseed green biomass (stems) to enhance pressing and extractives recovery. Bioresour. Technol. 2016, 199, 194–201. [Google Scholar] [CrossRef] [PubMed]
  187. Kumari, B.; Tiwari, B.K.; Walsh, D.; Griffin, T.P.; Islam, N.; Lyng, J.G.; Brunton, N.P.; Rai, D.K. Impact of pulsed electric field pre-treatment on nutritional and polyphenolic contents and bioactivities of light and dark brewer’s spent grains. Innov. Food Sci. Emerg. Technol. 2019, 54, 200–210. [Google Scholar] [CrossRef]
  188. Kronbauer, M.; Shorstkii, I.; Silva, S.; Toepfl, S.; Lammerskitten, A.; Siemer, C. Pulsed electric field assisted extraction of soluble proteins from nettle leaves (Urtica dioica L.): Kinetics and optimization using temperature and specific energy. Sustain. Food Technol. 2023, 1, 886–895. [Google Scholar] [CrossRef]
  189. Li, X.; Li, J.; Wang, R.; Rahaman, A.; Zeng, S.X.-A.; Brennan, C.S. Combined effects of pulsed electric field and ultrasound pretreatments on mass transfer and quality of mushrooms. LWT 2021, 150, 112008. [Google Scholar] [CrossRef]
  190. Boussetta, N.; Soichi, E.; Lanoisellé, J.L.; Vorobiev, E. Valorization of oilseed residues: Extraction of polyphenols from flaxseed hulls by pulsed electric fields. Ind. Crops Prod. 2014, 52, 347–353. [Google Scholar] [CrossRef]
  191. Martinchik, A.N. Nutritional value of sesame seeds. Vopr. Pitan. 2011, 80, 41–43. [Google Scholar] [PubMed]
  192. Kamboj, A.; Chopra, R.; Singh, R.; Saxena, V.; Prassana Kumar, G.V. Effect of pulsed electric field parameters on the alkaline extraction of valuable compounds from perilla seed meal and optimization by central composite design approach. Appl. Food Res. 2022, 2, 100240. [Google Scholar] [CrossRef]
  193. Pappas, V.M.; Lakka, A.; Palaiogiannis, D.; Athanasiadis, V.; Bozinou, E.; Ntourtoglou, G.; Makris, D.P.; Dourtoglou, V.G.; Lalas, S.I. Optimization of Pulsed Electric Field as Standalone “Green” Extraction Procedure for the Recovery of High Value-Added Compounds from Fresh Olive Leaves. Antioxidants 2021, 10, 1554. [Google Scholar] [CrossRef]
  194. Genovese, J.; Kranjc, M.; Serša, I.; Petracci, M.; Rocculi, P.; Miklavčič, D.; Mahnič-Kalamiza, S. PEF-treated plant and animal tissues: Insights by approaching with different electroporation assessment methods. Innov. Food Sci. Emerg. Technol. 2021, 74, 102872. [Google Scholar] [CrossRef]
  195. Saldaña, M.D.A.; Silva, E.K.; Olmos Cornejo, J.E.; Orozco Lopez, C.L. Comprehensive foodomics. In Green Processes in Foodomics: Biorefineries in the Food Industry; Elsevier: Amsterdam, The Netherlands, 2021; pp. 808–824. [Google Scholar] [CrossRef]
  196. Taha, A.; Casanova, F.; Šimonis, P.; Stankevič, V.; Gomaa, M.A.E.; Stirkė, A. Pulsed Electric Field: Fundamentals and Effects on the Structural and Techno-Functional Properties of Dairy and Plant Proteins. Foods 2022, 11, 1556. [Google Scholar] [CrossRef] [PubMed]
  197. Coustets, M.; Joubert-Durigneux, V.; Hérault, J.; Schoefs, B.; Blanckaert, V.; Garnier, J.-P.; Teissié, J. Optimization of protein electroextraction from microalgae by a flow process. Bioelectrochemistry 2015, 103, 74–81. [Google Scholar] [CrossRef] [PubMed]
  198. Segovia, F.J.; Luengo, E.; Corral-Pérez, J.J.; Raso, J.; Almajano, M.P. Improvements in the aqueous extraction of polyphenols from borage (Borago officinalis L.) leaves by pulsed electric fields: Pulsed electric fields (PEF) applications. Ind. Crops Prod. 2015, 65, 390–396. [Google Scholar] [CrossRef]
  199. Priyadarshini, A.; Rajauria, G.; O’Donnell, C.P.; Tiwari, B.K. Emerging food processing technologies and factors impacting their industrial adoption. Crit. Rev. Food Sci. Nutr. 2019, 59, 3082–3101. [Google Scholar] [CrossRef]
  200. Schoenbach, K.H.; Joshi, R.P.; Stark, R.H.; Dobbs, F.C.; Beebe, S.J. Bacterial decontamination of liquids with pulsed electric fields. IEEE Trans. Dielectr. Electr. Insul. 2000, 7, 637–645. [Google Scholar] [CrossRef]
  201. Safi, C.; Charton, M.; Pignolet, O.; Silvestre, F.; Vaca-Garcia, C.; Pontalier, P.-Y. Influence of microalgae cell wall characteristics on protein extractability and determination of nitrogen-to-protein conversion factors. J. Appl. Phycol. 2013, 25, 523–529. [Google Scholar] [CrossRef]
  202. Gad, A.; Jayaram, S.H.; Pritzker, M. Performance of electrode materials during food processing by Pulsed Electric Fields. IEEE Trans. Plasma Sci. 2014, 42, 3161–3166. [Google Scholar] [CrossRef]
  203. Moonesan, M.S.; Jayaram, S.H. Effect of pulsewidth on medium temperature rise and microbial inactivation under Pulsed Electric Field food treatment. IEEE Trans. Ind. Appl. 2013, 49, 1767–1772. [Google Scholar] [CrossRef]
  204. Joana Gil-Chávez, G.; Villa, J.A.; Fernando Ayala-Zavala, J.; Basilio Heredia, J.; Sepulveda, D.; Yahia, E.M.; González-Aguilar, G.A. Technologies for Extraction and Production of Bioactive Compounds to be Used as Nutraceuticals and Food Ingredients: An Overview. Compr. Rev. Food Sci. Food Saf. 2013, 12, 5–23. [Google Scholar] [CrossRef]
  205. Carpentieri, S.; Jambrak, A.R.; Ferrari, G.; Pataro, G. Pulsed Electric Field-Assisted Extraction of Aroma and Bioactive Compounds from Aromatic Plants and Food By-Products. Front. Nutr. 2022, 8, 792203. [Google Scholar] [CrossRef] [PubMed]
  206. Lajoie, L.; Fabiano-Tixier, A.-S.; Chemat, F. Water as Green Solvent: Methods of Solubilisation and Extraction of Natural Products—Past, Present and Future Solutions. Pharmaceuticals 2022, 15, 1507. [Google Scholar] [CrossRef] [PubMed]
  207. Wang, M.; Zhou, J.; Castagnini, J.M.; Berrada, H.; Barba, F.J. Pulsed electric field (PEF) recovery of biomolecules from Chlorella: Extract efficiency, nutrient relative value, and algae morphology analysis. Food Chem. 2023, 404, 134615. [Google Scholar] [CrossRef]
  208. Ranjha, M.M.A.N.; Kanwal, R.; Shafique, B.; Arshad, R.N.; Irfan, S.; Kieliszek, M.; Kowalczewski, P.Ł.; Irfan, M.; Khalid, M.Z.; Roobab, U.; et al. A Critical Review on Pulsed Electric Field: A Novel Technology for the Extraction of Phytoconstituents. Molecules 2021, 26, 4893. [Google Scholar] [CrossRef]
  209. Parniakov, O.; Barba, F.J.; Grimi, N.; Marchal, L.; Jubeau, S.; Lebovka, N.; Vorobiev, E. Pulsed electric field and pH assisted selective extraction of intracellular components from microalgae Nannochloropsis. Algal Res. 2015, 8, 128–134. [Google Scholar] [CrossRef]
  210. Carpentieri, S.; Mazza, L.; Nutrizio, M.; Jambrak, A.R.; Ferrari, G.; Pataro, G. Pulsed electric fields- and ultrasound-assisted green extraction of valuable compounds from Origanum vulgare L. and Thymus serpyllum L. Int. J. Food Sci. Technol. 2021, 56, 4834–4842. [Google Scholar] [CrossRef]
  211. Toepfl, S. Pulsed electric field food processing -industrial equipment design and commercial applications. Stewart Postharvest Rev. 2012, 8, 1–7. [Google Scholar] [CrossRef]
  212. Timmermans, R.A.H.; Mastwijk, H.C.; Berendsen, L.B.J.M.; Nederhoff, A.L.; Matser, A.M.; Van Boekel, M.A.J.S.; Nierop Groot, M.N. Moderate intensity Pulsed Electric Fields (PEF) as alternative mild preservation technology for fruit juice. Int. J. Food Microbiol. 2019, 298, 63–73. [Google Scholar] [CrossRef]
  213. Li, X.; Farid, M. A review on recent development in non-conventional food sterilization technologies. J. Food Eng. 2016, 182, 33–45. [Google Scholar] [CrossRef]
  214. Kempkes, M. Industrial pulsed electric field systems. In Handbook of Electroporation; Miklavcic, D., Ed.; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–22. [Google Scholar] [CrossRef]
  215. Bhat, Z.F.; Morton, J.D.; Mason, S.L.; Bekhit, A.E.-D.A. Current and future prospects for the use of pulsed electric field in the meat industry. Crit. Rev. Food Sci. Nutr. 2019, 59, 1660–1674. [Google Scholar] [CrossRef] [PubMed]
  216. Arshad, R.N.; Abdul-Malek, Z.; Munir, A.; Buntat, Z.; Ahmad, M.H.; Jusoh, Y.M.M.; Bekhit, A.E.-D.; Roobab, U.; Manzoor, M.F.; Aadil, R.M. Electrical systems for pulsed electric field applications in the food industry: An engineering perspective. Trends Food Sci. Technol. 2020, 104, 1–13. [Google Scholar] [CrossRef]
  217. Puértolas, E.; Barba, F. Electrotechnologies applied to valorization of by-products from food industry: Main findings, energy and economic cost of their industrialization. Food Bioprod. Process. 2016, 100, 172–184. [Google Scholar] [CrossRef]
  218. Toepfl, S.; Kinsella, J.; Parniakov, O. Industrial Scale Equipment, Patents, and Commercial Applications in Pulsed Electric Fields to Obtain Healthier and Sustainable Food for Tomorrow; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar] [CrossRef]
  219. Jaeger, H.; Balasa, A.; Knorr, D. Food industry applications for pulsed electric fields. In Electrotechnologies for Extraction from Food Plants and Biomaterials; Food Engineering Series; Springer: New York, NY, USA, 2009; pp. 181–216. [Google Scholar] [CrossRef]
  220. Mohammed, M.; Eiss, A.A. Pulsed electric fields for food processing technology. In Structure and Function of Food Engineering; Eissa, A.A., Ed.; Intech: Rijeka, Croatia, 2012; pp. 275–306. [Google Scholar] [CrossRef]
  221. Morren, J.; Roodenburg, B.; de Haan, S.W.H. Electrochemical reactions and electrode corrosion in pulsed electric field (PEF) treatment chambers. Innov. Food Sci. Emerg. Technol. 2003, 4, 285–295. [Google Scholar] [CrossRef]
  222. Pataro, G.; Barca, G.M.J.; Donsì, G.; Ferrari, G. On the modeling of electrochemical phenomena at the electrode-solution interface in a PEF treatment chamber: Methodological approach to describe the phenomenon of metal release. J. Food Eng. 2015, 165, 34–44. [Google Scholar] [CrossRef]
  223. Samaranayake, C.P.; Sastry, S.K.; Zhang, H. Pulsed Ohmic Heating–A Novel Technique for Minimization of Electrochemical Reactions During Processing. J. Food Sci. 2005, 70, e460–e465. [Google Scholar] [CrossRef]
  224. Kotnik, T.; Miklavčič, D.; Mir, L.M. Cell membrane electropermeabilization by symmetrical bipolar rectangular pulses: Part II. Reduced electrolytic contamination. Bioelectrochemistry 2001, 54, 91–95. [Google Scholar] [CrossRef]
  225. Gad, A.; Jayaram, S.H. Effect of electric pulse parameters on releasing metallic particles from stainless steel electrodes during PEF processing of milk. IEEE Trans. Ind. Appl. 2013, 50, 1402–1409. [Google Scholar] [CrossRef]
  226. Bocker, R.; Silva, E.K. Pulsed electric field assisted extraction of natural food pigments and colorings from plant matrices. Food Chem. X 2022, 15, 100398. [Google Scholar] [CrossRef] [PubMed]
  227. Picart-Palmade, L.; Cunault, C.; Chevalier-Lucia, D.; Belleville, M.-P.; Marchesseau, S. Potentialities and Limits of Some Non-thermal Technologies to Improve Sustainability of Food Processing. Front. Nutr. 2019, 5, 130. [Google Scholar] [CrossRef] [PubMed]
  228. Martínez, J.M.; Delso, C.; Álvarez, I.; Raso, J. Pulsed Electric Field-assisted extraction of valuable compounds from microorganisms. Compr. Rev. Food Sci. Food Saf. 2020, 19, 530–552. [Google Scholar] [CrossRef] [PubMed]
  229. Schottroff, F.; Gratz, M.; Krottenthaler, A.; Johnson, N.B.; Bédard, M.F.; Jaeger, H. Pulsed electric field preservation of liquid whey protein formulations—Influence of process parameters, pH, and protein content on the inactivation of Listeria innocua and the retention of bioactive ingredients. J. Food Eng. 2019, 243, 142–152. [Google Scholar] [CrossRef]
  230. Martínez-Beamonte, R.; Ripalda, M.; Herrero-Continente, T.; Barranquero, C.; Dávalos, A.; López de las Hazas, M.C.; Álvarez-Lanzarote, I.; Sánchez-Gimeno, A.C.; Raso, J.; Arnal, C.; et al. Pulsed electric field increases the extraction yield of extra virgin olive oil without loss of its biological properties. Front. Nutr. 2022, 9, 1065543. [Google Scholar] [CrossRef] [PubMed]
  231. Manzoor, M.F.; Zeng, X.A.; Rahaman, A.; Siddeeg, A.; Aadil, R.M.; Ahmed, Z.; Li, J.; Niu, D. Combined impact of pulsed electric field and ultrasound on bioactive compounds and FT-IR analysis of almond extract. J. Food Sci. Technol. 2019, 56, 2355–2364. [Google Scholar] [CrossRef]
  232. Steinbruch, E.; Wise, J.; Levkov, K.; Chemodanov, A.; Israel, A.; Livney, Y.D.; Golberg, A. Enzymatic cell wall degradation combined with pulsed electric fields increases yields of water-soluble-protein extraction from the green marine macroalga Ulva sp. Innov. Food Sci. Emerg. Technol. 2023, 84, 103231. [Google Scholar] [CrossRef]
  233. Zhou, J.; Wang, M.; Berrada, H.; Zhu, Z.; Grimi, N.; Barba, F.J. Pulsed electric fields (PEF), pressurized liquid extraction (PLE) and combined PEF + PLE process evaluation: Effects on Spirulina microstructure, biomolecules recovery and Triple TOF-LC-MS-MS polyphenol composition. Innov. Food Sci. Emerg. Technol. 2022, 77, 102989. [Google Scholar] [CrossRef]
Figure 1. Valorization of waste plant/mushroom resources (reproduced with permission from Elsevier, license no. 5670740416741).
Figure 1. Valorization of waste plant/mushroom resources (reproduced with permission from Elsevier, license no. 5670740416741).
Processes 12 00342 g001
Figure 2. Potential uses of PEF in food applications.
Figure 2. Potential uses of PEF in food applications.
Processes 12 00342 g002
Figure 3. Pulsed Electric Field setup.
Figure 3. Pulsed Electric Field setup.
Processes 12 00342 g003
Figure 4. Electroporation stages in a plant cell.
Figure 4. Electroporation stages in a plant cell.
Processes 12 00342 g004
Figure 5. PEF-assisted protein extraction process.
Figure 5. PEF-assisted protein extraction process.
Processes 12 00342 g005
Table 1. Protein contents of plant wastes.
Table 1. Protein contents of plant wastes.
S.NoPlant WastesProtein (%) *Protein (g/100 g) **Ref.
1Sesame Press Cake35–50X[43]
2Papaya Seeds27.3–28.3X[44]
3Olive pomace4.51 (Crude)X[45]
4Flaxseed hulls17.45–19.14X[46]
5Tomato wastes [47]
(i) Pomace15.1–22.7X
(ii) Peels5.7–20X
(iii) Seeds16.6–39.3X
6Brewer’s spent grains26–30X[48]
7Cabbage leaf residues17.0–17.2 (Crude protein on DM*)X[49]
8Potato peels13.02 (dry powder)X[50]
9Apple pomace4X[51]
10Orange peel waste7.7X[52]
11Rapeseed meal36X[53]
12Pumpkin peel1.65X[54]
13Ripe Banana peel5.5–7.87 (Crude)X[55]
14Pomegranate peel4.90–8.97X[56]
15Inodorus melon peel34.90X[57]
16Sweet potato peel [58]
(i) Primary peel6.40–6.49X
(ii) Blanched peel8.11–8.20X
17Pumpkin seeds~35 (Crude)X[59]
18Wheat Bran 13.2–18.4X[60]
19Rice bran10–16X[61]
20Outer scales of onion2.64X[62]
21Almond HuskX3.27 (Crude)[63]
22Brewer dry grainX19.96 (Crude)[63]
* Representation of protein quantity in %. ** Representation of protein quantity in g/100 g. X, not applicable. DM*, Dry Matter.
Table 2. Properties, health benefits, and protein content of commercially cultivated mushrooms.
Table 2. Properties, health benefits, and protein content of commercially cultivated mushrooms.
Mushroom SpeciesProtein Content (%)Nutritional Properties Medicinal PropertiesReference
Agaricus Bisporus29.64–39.84
  • Rich in carbohydrates, amino acids, fats, and minerals
  • Used in the treatment of cancer, bacterial and fungal infections, diabetes, heart disorders, and skin problems
[65,81]
Pleurotus:
Pleurotus ostreatus
Pleurotus eryngii
Pleurotus sajor-caju
Pleurotus giganteus

7
11
37.4
17.7
  • Rich in proteins and fiber, low in carbohydrates, fats, and sodium
  • Used to treat atherosclerosis.
Other properties:
  • Antimicrobial
  • Antiviral
  • Anticancer
  • Antioxidant
  • Hypolipidemic
  • Hypocholesterolemic
  • Antihyperglycemic
  • Immunomodulatory effects
[82,83,84]
Morchella esculenta32.7 (DW)*
  • Rich in protein, fiber, and necessary vitamins and minerals.
Low in calories and fat
  • Antioxidative
  • Immune protective
  • Anti-inflammatory
  • Anti-colon cancer effects
[85,86]
Lentinula edodes15–23
  • Low lipid contents
  • Rich in protein, vitamins, and minerals.
  • High concentrations of calcium, iron, phosphorus, potassium, zinc, manganese, phenolic compounds, and glucans
Medicinally used for the following diseases
  • Depressed immune function (including AIDS)
  • Cancer
  • Environmental allergies
  • Fungal infection
  • Frequent flu and colds
  • Bronchial inflammation
  • Heart disease
  • Hyperlipidemia (including high blood cholesterol)
  • Hypertension
  • Infectious diseases
  • Diabetes
  • Hepatitis
  • Regulating urinary inconsistencies.
[87,88,89]
Auricularia heimuer10.62
  • Plentiful in several amino acids, especially leucine and lysine.
  • High in protein, trace element, vitamin, and carbohydrate content
  • Low in fat
  • Anticancer
  • Detoxifying Anticoagulant
  • Hypoglycemic
  • Cholesterol-lowering capacity
[90]
Volvariella bombycina
Fruit Body
Mycelia

28.30
25.50
  • Rich in proteins, carbohydrates, vitamins, and minerals
  • Resistance to ovarian, lung, breast, and prostate cancer
[91]
Enoki mushrooms:
Flammulina filiformis,
Flammulina velutipes
5
  • Rich in carbohydrates, proteins, unsaturated fatty acids, many valuable micronutrients, and dietary fiber
  • Highly comparable to vegetables
  • Anti-inflammatory activities
[92,93]
Tremella fuciformis9.63
  • Contains large quantities of vitamin D, proteins, vitamins, minerals, immune-boosting polysaccharides, trace minerals, and carbohydrates
  • Low in fat
  • Used medicinally for cough, sore throat, stethalgia, constipation, and paramenia
[94,95,96]
Hypsizygus tessellatus,33.9
  • Rich in proteins, fiber, carbohydrates, potassium, calcium and phosphorous
  • Has organoleptic and pharmacological properties,
  • Anticancer properties
[97,98]
Stropharia rugosoannulata,30−50
  • Rich in proteins, peptides, and amino acids
Essential Properties:
  • Antidiabetic
  • Antimicrobial
  • Antioxidant
  • Antiproliferative
  • Antitumor
  • Immunomodulatory and osteoclast-formation-suppressing properties
[99,100]
Cyclocybe aegerita37.6
  • Rich in carbohydrates, ash, and proteins
Essential Properties:
  • Antitumor, antioxidant, and antifungal
  • Hypocholesterolemic and hypolipidemic effects
[101,102,103]
Hericium erinaceusDried powder, 20;
Mycelia, 42
  • Enriched protein and dietary fiber content and low in calories and fat
Essential Properties:
  • Antibiotic
  • Anticarcinogenic
  • Antidiabetic
  • Antifatigue
  • Antihypertensive
  • Anti-hyperlipodemic
  • Antisenescence
  • Cardioprotective
  • Hepatoprotective
  • Nephroprotective
  • Neuroprotective properties
  • Improvements in anxiety, cognitive function, and depression.
[104,105]
Phallus Indusiatus4.813
  • Contains very high water, crude fiber, protein, fat, carbohydrate, ash, and vitamin C content.
  • Improves cardiovascular health
  • Reduces the risk of cancer
  • Promotes immune function
  • Prevents viruses, bacteria, and fungi
  • Reduces inflammation
  • Combats allergies
  • Helps to balance blood sugar levels
  • Supports the body’s detoxification mechanism
[106]
DW*, Dry Weight.
Table 3. PEF-assisted protein extraction studies from waste plant parts and mushrooms.
Table 3. PEF-assisted protein extraction studies from waste plant parts and mushrooms.
Protein SourceProcess ParametersOutcomeEssential FindingsReference
Agaricus Bisporus38.4 kV/cm,
Pulse width of 2 µS
at both 400 Hz and
800 Hz frequencies
Inlet Temperature: 20 °C
Protein extraction
of 40.8%
* Significant increase in protein yield when combined with mild heating
* Protein concentration increased with an increase in the electric field, temperature, and treatment time
[119]
Agaricus Bisporus2.5 kV/cm,
Specific energy of
50 kJ/kg, and 6 h
of extraction time
Maximum protein
recovery of 140.72 +/−
15.14 mg BSA/g DM*
* PEF increased the
recovery of proteins
[143]
Papaya peels13.3 kV/cm, 400 pulses,
40 kV applied voltage,
treatment time of 2720 secs.
Protein concentration of 20 mg/L* Two stages of PEF + supplementary aqueous extraction resulted in good recovery of bioactive compounds including proteins[165]
Rapeseed stems and leaves 20 kV/cm
200 pulses, 40 kV applied voltage,
pulse duration of 10 μs
Protein extraction of
up to 80%
* PEF treatment increased the yield of proteins.
* Protein contents changed during plant development
[4]
Rapeseed green biomass (stems)8 kV/cm, Pulse duration of 10 μS, pressure of 10 bar, maximum applied voltage of 40 kV, frequency of 0.5 Hz, 200 pulses 0.14 g BSA/100 g DM* * Protein content increased by about 2 times compared with the untreated sample and 50 pulse applications[186]
Sesame cake13.3 kV/cm
Pulse duration of 10 μs
Frequency of 0.5 Hz (2 s between pulses or discharges)
0–700 pulses
Temperature of 40 °C
Variable treatment time between 1 and 7 ms
Energy inputs between 0 and 291 kJ/kg
Between 374 and 2001 mg BSA/100 g DM*.* Protein contents increased with energy inputs until 83 kJ/kg was reached.
* PEF improved the protein extraction yield.
[183]
Tomato wastes5 kV/cm
Treatment time of 1.5 ms
200 pulses
145.1 mg/100 g* Protein concentration increased with increasing the electric field strength and time[158]
Rice bran2.3 kV/cm
Treatment time of 25 min
250 pulses per minute
20.71–22.8%* Improved the extraction yield of rice bran protein
* Improved oil’s holding and emulsifying properties
* Increased foaming ability and stability
* Enhanced in vitro digestibility
[145]
Brewers spent grains (BSG)
Light
Dark
2.8 kV/cm
3000 pulses
20 μs pulse-width
20.31 ± 0.01 (%DWe*)
23.78 ± 0.16 (%DWe*)
* Increased yield of proteins from dark BSG compared with light BSG
* Light BSG exhibited higher antimicrobial activity than untreated ones
[187]
Nettle leaves3 kV/cm
Specific energy of 10–24 kJ/kg
Temperature between 70 and 78°C
Extraction time of 5 min
>60%PEF specific energy input, extraction temperature, and particle size had high impacts on solid liquid extraction kinetics[188]
DM*, Dry Matter; DWe*, Dry weight extract.
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

Ramaswamy, R.; Krishnan, S.B.; Leong, S.S.J. Pulsed Electric Field Technology for Recovery of Proteins from Waste Plant Resources and Deformed Mushrooms: A Review. Processes 2024, 12, 342. https://doi.org/10.3390/pr12020342

AMA Style

Ramaswamy R, Krishnan SB, Leong SSJ. Pulsed Electric Field Technology for Recovery of Proteins from Waste Plant Resources and Deformed Mushrooms: A Review. Processes. 2024; 12(2):342. https://doi.org/10.3390/pr12020342

Chicago/Turabian Style

Ramaswamy, Ramya, Sivaneasan Bala Krishnan, and Susanna Su Jan Leong. 2024. "Pulsed Electric Field Technology for Recovery of Proteins from Waste Plant Resources and Deformed Mushrooms: A Review" Processes 12, no. 2: 342. https://doi.org/10.3390/pr12020342

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