Next Article in Journal
Evaluating the Effectiveness of an Augmented Reality Game Promoting Environmental Action
Next Article in Special Issue
Occupational Risk Assessment in School Food Services: Instruments’ Construction and Internal Validation
Previous Article in Journal
Evaluation of the Consumer Perception of Sharing Economy: Cases of Latvia, Russia, Ukraine and Belarus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

High-Pressure Processing for Sustainable Food Supply

by
Brera Ghulam Nabi
1,†,
Kinza Mukhtar
1,†,
Rai Naveed Arshad
2,
Emanuele Radicetti
3,*,
Paola Tedeschi
3,
Muhammad Umar Shahbaz
4,
Noman Walayat
5,
Asad Nawaz
6,
Muhammad Inam-Ur-Raheem
1 and
Rana Muhammad Aadil
1,*
1
National Institute of Food Science and Technology, University of Agriculture, Faisalabad 38000, Pakistan
2
Institute of High Voltage and High Current (IVAT), School of Electrical Engineering, Faculty of Engineering, Universiti Teknologi Malaysia, Johor Bahru 81310, Johor, Malaysia
3
Department of Chemical, Pharmaceutical and Agricultural Sciences (DOCPAS), University of Ferrara, Via Luigi Borsari 46, 44121 Ferrara, Italy
4
AARI, Plant Pathology Research Institute, Faisalabad 38850, Pakistan
5
College of Food Science and Technology, Zhejiang University of Technology, Hangzhou 310014, China
6
Shenzhen Key Laboratory of Marine Microbiome Engineering, Institute for Advanced Study, Shenzhen University, Shenzhen 518060, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this paper.
Sustainability 2021, 13(24), 13908; https://doi.org/10.3390/su132413908
Submission received: 4 November 2021 / Revised: 7 December 2021 / Accepted: 7 December 2021 / Published: 16 December 2021

Abstract

:
Sustainable food supply has gained considerable consumer concern due to the high percentage of spoilage microorganisms. Food industries need to expand advanced technologies that can maintain the nutritive content of foods, enhance the bio-availability of bioactive compounds, provide environmental and economic sustainability, and fulfill consumers’ requirements of sensory characteristics. Heat treatment negatively affects food samples’ nutritional and sensory properties as bioactives are sensitive to high-temperature processing. The need arises for non-thermal processes to reduce food losses, and sustainable developments in preservation, nutritional security, and food safety are crucial parameters for the upcoming era. Non-thermal processes have been successfully approved because they increase food quality, reduce water utilization, decrease emissions, improve energy efficiency, assure clean labeling, and utilize by-products from waste food. These processes include pulsed electric field (PEF), sonication, high-pressure processing (HPP), cold plasma, and pulsed light. This review describes the use of HPP in various processes for sustainable food processing. The influence of this technique on microbial, physicochemical, and nutritional properties of foods for sustainable food supply is discussed. This approach also emphasizes the limitations of this emerging technique. HPP has been successfully analyzed to meet the global requirements. A limited global food source must have a balanced approach to the raw content, water, energy, and nutrient content. HPP showed positive results in reducing microbial spoilage and, at the same time, retains the nutritional value. HPP technology meets the essential requirements for sustainable and clean labeled food production. It requires limited resources to produce nutritionally suitable foods for consumers’ health.

1. Introduction

Sustainability is a vital need of fulfilling the basic societal requirements by improving social, environmental, and economic systems [1]. With the world’s increasing population, people mainly focus on healthy, nutritious, and safe foods [2]. For this reason, the main target of the United Nations 2030 programs is the implementation of the latest food production for the sustainability of food. Sustainability is a major feasible way to access the upcoming basic requirements for a safe, nutritious, and healthy diet [3]. Sustainable food processing demands the exclusive use of natural resources to minimize food waste [4]. Many foods, especially fruits, vegetables, and grains, are wasted during storage, transportation, and processing [5]. Nowadays, the food industry is forced to develop sustainable food systems [6]. Some limitations of conventional technologies that lead to the adaptation of non-thermal technologies are off-flavors due to chemical reactions, fluctuations in nutrients, and losses of food quality [7].
The food industry has adopted different effective non-thermal techniques such as a pulsed electric field [8] (PEF), ultra-sonication [9] (US), cold plasma [10], high-pressure processing [11] (HPP), and ultraviolet radiation. All of these techniques maintain the original food quality, overcome nutrient loss, and have a low energy consumption compared with conventional systems [12]. The effects on the flavor, viscosity, appearance, nutritional value, and practical function in inactivating pathogens and vegetative micro-organisms are reduced compared with many conventional therapies [13]. HPP technology is considered environmentally friendly and kills bacterial cells, yeasts, and molds without heat [14]. It lessens the need for chemical additives and increases bioavailability [15,16]. HPP is perceived as a replacement for thermal techniques because it improves food safety, increases shelf life, and preserves the sensorial, physiochemical, and nutritional content of food [17]. The basic goal of HPP of meat and meat items is to minimize the infectious and spoilage microbes and to extend the shelf life [18].
HPP is also called high hydrostatic pressure, ultra-HPP, pascalisation or cold pasteurization. In this technique, high pressure is applied to solid or liquid foods to enhance their safety, organoleptic attributes, and quality. Firstly, high-pressure treatment was applied in the late 1890s for the inactivation of microbes in dairy products, i.e., milk [19]. HPP works on the isostatic principle, which states that different media such as water or oil apply constant pressure to the sample. The combination of water with oil lubricates the equipment and reduces corrosion. Thus, it helps eliminate the microorganisms in food samples and is not affected by the size and shape of the food samples [20]. According to the isostatic principle, HPP does not depend on volume; therefore, pressure is transmitted instantly and constantly throughout a sample, and pressure gradients do not exist. HPP has been utilized in food sterilization and preservation processes to enhance the shelf life of food and maintain high quality by following Le Chatelier’s and Pascal’s principles for specific purposes [21]. Le Chatelier’s Principle states that phenomena lead to a decrease in volume, such as state changes, chemical reactions, and alterations in molecular arrangement, enhanced by pressure. In contrast, volume increases by reduced pressure. HPP will perform an essential role as an alternative approach in food sterilization when the transmission medium is water [22].
It has been proven that HPP allows shelf-life extension [23], pathogen removal [24], and clean-label [25] convenience while providing natural and safe food to consumers. HPP has been utilized in different food categories such as vegetables, fruits, dairy, meat, sea foods [26], jams, fruit jellies, sauces, purees, juices [27], ready to eat meat, Deli Salads, Dips, Salsa markets [28], infants food, and the fish industry for different purposes. Table 1 describes the applications of HPP in different industries of food.
HPP operated at 350 MPa for 4 min and combined with 0.05% allyl isothiocyanate (AITC), showed a 5-log reduction of Salmonella and at 350 MPa for 12 min with 0.075% AITC, showed more than 7-log inactivation of Salmonella in ground chicken meat [47]. HPP as a pretreatment maintained the color and texture of albacore steaks when operated at 200 MPa for 6 min [48]. Nunez-Mancilla et al. [49] studied the HPP impact on strawberries at 100–500 MPa for 10 min. The results clearly showed higher antioxidant activity and a total phenolic content at 400 MPa; HPP also retained the vitamin C content in strawberries. This review mainly focuses on describing HPP sustainability, advantages, and limitations. HPP is an efficient, non-thermal, and the most accepted technique by the consumer [50]. Furthermore, the continuous demand for healthy and safe foods has discovered HPP for mild food preservation without chemicals and preservatives.
The current study was done based on research articles found on 14 March 2021, using the internet database Web of Science (WoS). The authors discovered some of the most highly cited articles referencing the term “high-pressure processing”. The search revealed papers that referred to “high-pressure processing” in their titles, abstracts, or author keywords. Furthermore, the authors established two criteria to limit the relevant research papers that were selected: either the most cited or the most current research.

2. Working Mechanism of HPP

HPP equipment mainly consists of the following components: a pressure vessel composed of high-tensile strength steel determined by the process’s lethality, closures or plugs, a yoke/wire-wound steel frame, a pressure pump, and a control system [51] as demonstrated in Figure 1. The food sample is kept in a pressure chamber using water as a pressure-transfer medium after being filled and sealed in flexible pouches. Additional fluid may be injected into the pressure chamber to create pressure. Because of the isostatic principle, pressure is uniformly distributed across the whole product in the container, ensuring that all food parts are under equal pressure. Surprisingly, the product’s size and shape have minimal impact on the HPP treatment.
A pressure vessel may also be used to create a pressure intensifier. The pressure vessel, which holds a packed food product and pressurization fluid, is intended to endure a certain pressure and temperature. The quantity of fluid that a pressure vessel can contain determines its size, varying from a few milliliters in research units to several hundred liters in commercial units [52].
Both batch and continuous HPP systems are utilized in food industries. Batch process systems are used for most solid food products; however, liquids and other pumpable goods have the option of being treated in a continuous system [53]. Semi-continuous process systems comprise two or more pressure vessels, a small pressure pump to fill the vessels, a high-pressure transmission pump, holding and sterile tanks, and controlling valves [54]. Pressure vessels contain a separating piston that moves freely to separate the product from the pressurizing fluid. Controlling valves prevent cross-contamination of the treated product with the upcoming untreated product [55]. At the start of the process, the pressure vessel using a low-pressure pump fills with the liquid product. When the food section is filled, the inlet valve is closed, and pressurizing fluid moves the free piston to compress the liquid. The vessel is depressurized after a suitable holding time, which will decompress the food and the piston returns to its original state. Finally, the processed liquid product is aseptically packed in sterilized containers [53]. Vessels are connected so that when one vessel discharges the product, the second system is pressurized, and the third is loaded with a food sample. Therefore, a continuous product is maintained. In a commercial-scale industry, three pressure vessels are used to provide a continuous product output. This production is generated by the operation of these three vessels simultaneously; one is loading, the second is compressing, and the last is discharging at any point in time [56].
Pressure, such as heat, is a fundamental thermodynamic variable. In a strict sense, the effects of temperature and pressure cannot be isolated during HPP. This is because each temperature has an associated pressure. Thermal effects result in volume and energy changes during pressure treatment. However, pressure mainly influences the product volume. The combined net impact may be synergistic, antagonistic, or additive [57]. Hence, the HPP process is characterized by three parameters: temperature (T), pressure (p), and exposure time (t) compared to heat preservation. Figure 2 shows the pressure, temperature, and time relationship in HPP.
The pressure is built up to a particular point until it reaches the desired pressure. The target pressure (pressure established as a standard for the product) is maintained for a specific time before being released. The pressure at which the sample food is held in the pressure vessel is termed the process pressure. The temperature at which the product obtains its process pressure is called the process temperature. The process temperature depends on the initial temperature, pressure transmitting fluid, and heat compression values. The time required to increase pressure from 0.1 MPa to the required process pressure is called the pressure come-up time [52]. However, the pump power (horsepower) and target process pressure increase or decrease the pressure come-up time. Strain gauges that function on the principle of a change in resistance under pressure strain (compression-decompression) and displacement transducers mounted on the external frame can be used to measure the pressure [58]. The pressure holding time is the time interval over which the product needs to be held under the process pressure. It is the period that occurs between the end of compression and the start of decompression.
In pulsed-HPP (p-HPP), pressure is applied with pulses to the sample with a total holding time [59]. The effectiveness of the p-HPP process has been reported to be more effective than an equivalent single pulse of equal time [60].
A pressure holding time of 3–10 min is used for economically viable commercial processes. After the pressurization, the temperature of the product is increased due to molecular arrangements. Therefore, a decompression period is needed to return the food sample from the process pressure to air pressure (0.1 MPa), and the product temperature is also reduced during depressurization [58]. HPP is applied at 400 to 600 MPa in most processes at ambient temperature. Most HPP processes operate at low temperatures, depending on the pressure of food processing. Depending on the product’s composition, the water temperature of the product in the pressure chamber may increase by 3–6 °C for every 100 MPa increase in pressure. Table 2 shows the target pressure for different HPP applications in food.
The target pressure must be varied for different food applications. For example, the target pressure for the extraction of xyloglucan from tamarind was 250–500 MPa [35]. Moreover, for pasteurization and sterilization, the target pressure was 300–600 MPa [64] and 500–900 MPa [63].

3. HPP Processing towards Sustainability

The Food and Agriculture Organization (FAO) and World Health Organization (WHO) reported on world food day 2019 the “Sustainable healthy diets–guiding principles” (FAO and WHO, 2019) and have described three pillars of sustainability: social, economic, and environmental. It was clearly defined that sustainable healthy diets are complete dietary patterns that improve an individual’s health and wellbeing [65]. Furthermore, sustainable diets are accessible, feasible, safe, and nutritious and are mostly accepted by consumers and have less impact on the environment.
Food product quality and safety are the most crucial factors for society and industry [66]. HPP is an environmentally friendly technology, which provides sustainable food without compromising the safety or quality of products and improves the food sector economically [67]. Moreover, HPP is an all-natural process that assures clean labeling of food products. According to consumer demands, HPP allows natural food ingredients without the addition of chemicals and preservatives [68]. Table 3 demonstrates the HPP contributions towards sustainability in different food applications. The following part demonstrates each part of this contribution.
HPP enhanced the extraction rate of bioactive compounds such as polyphenols, carotenoids, isothiocyanates, fatty acids, essential oils, and nutrients in food components [77,78]. For example, HPP enhanced the extraction of ferulic acid from Angelica sinensis [79] and the extraction of phenolic compounds from Chilean papaya (Vasconcellea pubescens) seeds [80]. Moreover, the extraction of astaxanthin carotenoid from shrimp shells was done efficiently by HPP at 200 MPa for 5 min [81]. Figure 3 illustrates the role of HPP in sustainable food production.
HPP was found to be capable of destroying vegetative as well as pathogenic bacteria and enzymes under certain circumstances; thus, it has been increasingly employed to compensate for the limitations of conventional thermal pasteurization and sterilization methods [82]. Microbes have a role in food spoilage and food losses. The inactivation of microbes helps to make food safe and extend the shelf life. HPP increases the product quality, sensory attributes, and shelf life and decreases microbial shelf life [64,83]. Woolf et al. [84] stated that the fragile sensory characteristics of avocado could be preserved with a long shelf life. HPP can also improve the product texture and structure [85].

3.1. Food Safety

Food safety has been a significant issue focused on since 1963 by the United Nations through an international forum. The FAO stated food safety as the “assurance that food will not cause harm to the consumer when it is prepared and/or eaten according to its intended use” [86]. According to another definition, food safety refers to food free from hazards [87]. Therefore, the number of interferences involving food safety can be diminished by assessing the possible hazards to forestall them and their adverse consequences. Nowadays, HPP plays a role in reducing microbes, enhancing shelf life, preserving nutrients, and assuring the safety and quality of raw and processed food items [88]. For example, in the meat industry, HPP is helpful to increase the shelf life and safety of ready-to-eat meat products [89].

3.1.1. Food Preservation

Preservation of food reduces microbial growth. Survival and tendency to grow depend on the response of microbes. Spoilage and pathogenic microbes affect food quality and safety. The main requirement is the production of high-quality foods with high nutritional value. The latest preservation technologies were accepted based on their efficiency to reduce pathogenic and spoilage foodborne microorganisms [90]. HPP slows down the activation of microorganisms by controlling their reproduction and survival rate [91]. HPP reduced L. monocytogenes by about 0.91 log10 colony forming unit (cfu)/g at 600 MPa in different food items such as guacamole, cheese, fruit juices, meats, and seafood [92]. HPP has been proved as an effective food preservation technique to replace thermal techniques, especially in the meat industry. [93]:
i.
Liquid Food Preservation:
According to Food and Drug Administration (FDA) terms, liquid food demands processes that show at least a 5-log reduction during preservation [6]. HPP successfully removes pathogenic and spoilage microorganisms in liquid foods [23]. In addition, a study on cactus juice was performed at 600 MPa, 15 °C for 10 min by HPP, and a 3-log reduction of yeast/mold was observed [94]. Table 4 explains the microbial load in some liquid foods after HPP treatment.
In 2013, orange juice was treated at 200–600 MPa, 20–60 °C by HPP. The result showed a 2-log reduction in Alicyclobacillus acidoterrestris [103]. Hiremath and Ramaswamy [104] concluded that applying 400 MPa for 10 min caused a 6-log inactivation of E. coli O157:H7 in mango juice using HPP. According to Huang et al. [105] E. coli O157:H7 showed 6-log CFU/g inactivation in frozen strawberry puree, under HPP conditions of 450 MPa for 2 min at 21 °C. Shahbaz [106] stated that applying HPP at 600 MPa and 25 °C for 1 min with 8.45 J/cm2 of TiO2-UV photocatalysis caused 7.1-log CFU/mL and 7.2-log CFU/mL reductions of E. coli and Salmonella Typhimurium in apple juice. Syed et al. [107] treated orange juice by HPP at 700 MPa for 5 min for complete inactivation of S. aureus. HPP successfully reduced the microbial content (p ≤ 0.05) in Sabah Snake Grass juice at 400 and 500 MPa [11].
ii.
Solid Food Preservation
HPP is a beneficial non-thermal process utilized to inactivate microbiota in certain foods [108]. HPP has been used for more than 100 years as a well-known pasteurization technique to make food products safe and contaminant-free [109]. HPP at 300–600 MPa effectively stopped the growth of decaying microbes in meat products [110].
HPP with antimicrobials increased the decontamination of foodborne pathogens and microbial safety of meat and meat products [111]. For example, parma ham treated with HPP showed a reduction of L. monocytogenes and was easily packaged and sold worldwide [112]. The same results were shown in the case of dry-cured ham by enhancing the shelf life of ham [113].
Water plays a major role in microorganism growth and enzymatic reactions in food [114] and therefore must be removed. HPP plays a role in food dehydration. For example, the combined effect of HPP and osmotic stress on the dehydration of ginger slices were studied [115]. At 600 MPa, the moisture content decreased, and the solid content increased.
HPP induced freezing and thawing, which formed ice crystals, resulting in a uniform crystal structure within the food matrices [116]. Thus, HPP preserves the structure and texture of solid food products through freezing. For example, peach and mango were treated at 200 MPa and −20 °C; high-pressure shift freezing of peach and mango with high supercooling formed uniform rapid ice crystals throughout the sample. As a result, large ice crystal formation and quality losses due to freezing cracking could be reduced [117].
HPP-assisted thawing caused less quality damage than traditional freezing techniques in meat by applying 100–200 MPa for 10 min [118]. According to another study, broccoli was treated at about 180–210 MPa pressure and −16 °C to −20 °C, which decreased the protein content and did not inactivate peroxidase and polyphenol oxidase enzymes. After one month, unpleasant changes in the flavor of broccoli occurred, but the texture remained the same. It destroyed the vacuole membrane and disorganized internal cells [119].

3.1.2. Effect of HPP on Food Contaminants

Food contaminants consist of environmental contaminants, food processing contaminants, unapproved adulterants, food additives, and migrants from packaging materials [120]. Environmental contaminants include impurities that are introduced by humans or naturally arise in water, air, or soil. Food processing contaminants contain unwanted components such as nitrosamines, chloropropanols, acrylamide, furanes, and polycyclic aromatic hydrocarbons (PAH) during baking, roasting, canning, heating, fermentation, or hydrolysis [121]. Moreover, chemical supplies found in food, animal feed, or antibiotic injections injected into poultry animals are examples of food contamination from chemicals [122]. HPP reduces the activity rate of pectin methylesterase (PME) isoenzymes. They mainly cause pectin degradation and cloud loss in most citrus juices. HPP showed less food quality loss, flavor retention, and eliminated spoilage microflora, i.e., yeasts, molds, and lactic acid bacteria, in citrus juice [123].
Over the last 10 years, HPP has become the most reliable and fastest decontamination technology, so many commercial large-scale units have been introduced in western countries [124]. Research has shown the combined effect of pressure and temperature (high-pressure thermal sterilization) in the inactivation of spores [125]. Therefore, high-pressure thermal sterilization (HPTS) was certified in 2015 and it reduces food processing contaminants (FPCs), e.g., furan [126], acrylamide, hydroxymethylfurfural (HMF), and 3-MCPD/-esters etc. HPP induced microbial inactivation and structural alterations in food components [127]. Therefore, it has been proved that it is a beneficial tool for the reduction of pesticides and mycotoxins. For example, studies showed that the common pesticide chlorpyrifos in fruits and vegetables, e.g., tomatoes and Brussel sprouts [128] was reduced under optimal conditions to 75 MPa, 5 °C, by HPP.

3.1.3. Reduction of Pesticides

Pesticides are helpful to eliminate pests, enhance shelf-life, and maintain quality. Specific laws enable their production, marketing, and usage [129]. About two million tons of pesticides are utilized every year worldwide. Sharma et al. [130] reported a serious risk by contaminating raw sources of food. The pesticide residue levels in food are controlled by law to lessen exposure to the consumer [131,132].
HPP has been utilized as a phytosanitary treatment for reducing insect pests in fresh or processed fruits and vegetables to enhance their shelf-life [133]. Moreover, HPP applied with pre-sterilized packaging reduced chemicals in liquid effluents [134].

3.1.4. Degradation of Toxins

Due to the increasing demand for safe food, it is essential to explore new innovative techniques for reducing pesticides and mycotoxins in food products [135]. Studies showed that most foods are contaminated by pesticides and mycotoxins [136]. The major sources of contamination are polluted raw materials [137] that are contaminated before coming into food factories [138]. Cereals utilized as staple food have a high risk of mycotoxins, such as aflatoxin and ochratoxin. HPP has a serious effect on mycotoxins and minimizes their toxicity in the environment [13]. Kalagatur et al. [61] found that when 550 MPa pressure was applied to maize grains at 45 °C for 20 min, the maximum reduction in ZEA and DON levels occurred. Another experiment was done to reduce two heat-resistant fungi, Aliivibrio fischeri and Talaromyces macrospores, in strawberry puree by combining pressure and heat [139].
HPP is an efficient technique utilized in the food sector to inactivate spore fungi and stop their growth. According to research, it was found that HPP (600 MPa) with US treatment (24 kHz/0.33 W/mL) at 75 °C for 30 min eliminated the ascospores of Byssochlamys nivea in strawberry puree [140]. In another study, the combination of HPP (600 MPa) and US (24 kHz/0.33 W/mL) at 75 °C helped to remove ascospores of Neosartorya fischeri and B. nivea in apple juice [62,141]. Moreover, the combination of HPP with ultrasound can enhance decontamination. A study by Huang et al. [142] concluded that 600–800 MPa pressure inactivated the mycotoxigenic fungus A. flavus. Tokusoglu, Alpas, and Bozoglu [143] proved that mold flora and the level of citrinin mycotoxin in black table olives were successfully reduced by HPP treatment.

3.2. Food Security

Food security is a multifaceted concept, and it can be achieved at the individual, household, national, regional, and global levels when all people consistently have physical and economic access to adequate, safe, and nutritious food to fulfill their intake needs and food choices for an active and healthy life. Food insecurity, malnutrition, and poverty are the most serious global challenges of the 21st century. Food insecurity is chronic. Xie et al. [144] showed an alarming food insecurity situation. The period from 2013 to 2019 was shown as an active period of food insecurity. According to one national nutrition survey, food insecurity has reached its highest level, it was 58% in 2005–2006 and about 72.8% in 2013, causing troubles in some regions.

3.2.1. Sustainable Nutrition Security

During 2014–2016, almost 795 million people worldwide were undernourished [145]. Therefore, there is a need for non-thermal techniques that preserve the nutrients and the quality of food products. Non-thermal preservation techniques result in minimal changes in flavors, nutrients, and vitamins due to low temperatures [146].
HPP generally causes no marked nutritional loss in foods compared with conventional thermal processes [147]. In addition, some studies showed no significant changes in carotenes immediately [148]. Some authors suggest that antioxidant activity increases by HPP [149], but others have come to the opposite conclusion [150]. Table 5 shows the effect of HPP on food nutrients in different food samples.
Additionally, it was considered that the HPP technique at 100, 200, and 400 MPa efficiently enhanced the functional properties of pine nut protein isolates with enhanced heat-induced gel strength [162]. Compared with conventional treatments, HPP enhanced nutrient retention [163] and reduced the effects on antioxidant activity [164]. As a result, HPP enhances the food shelf life and decreases nutritional losses [165].
HPP preserves the value of the vitamins in food; for example, it was observed that heat, light, pH, metals, and free radicals are prominent factors that affect vitamin E activity. In an anaerobic environment, tocopherols and trienols are stable to heat and alkaline conditions. A pressure of 400 to 600 MPa caused no marked reduction in the amount of tocopherol in milk [166]. However, HPP increased the concentration of phenolic compounds [167]. It was concluded that the polyphenol content of apple juice (600 MPa, 50 min) was 409.2 mg, which was 14% higher than that in thermal treatments. In addition, HPP application influences the stability of chlorophyll at a high temperature and low pressure. Medina-Meza et al. [168] showed high stability of chlorophyll at lower pressures from 400 to 650 MPa in spinach sauce.
HPP enhanced the solubility of myofibrillar protein at low pressure. Ziye et al. [169] stated that myofibrillar proteins are sensitive to high pressure. At 100–200 MPa, the quaternary structure dissociates, while above 200 MPa, tertiary structures are affected, and at 300–700 MPa, secondary structures are affected. HPP improves the gelation properties of myofibrillar proteins. Chan et al. [170] found that at low-pressure (≤200 MPa), the solubility of myofibrillar proteins improved; at a pressure higher than 300 MPa, the solubility of myofibrillar proteins decreased and large aggregates formed. Hence, it was proved that HPP successfully improved the texture of food products [171].
HPP plays a role in the tenderness of meat, which depends on myofibrillar and connective tissue. The mechanisms of meat tenderization that occur in the HPP of pre-rigour muscles and chill ageing of post-rigour muscles are different. HPP causes alterations in the muscle microstructure, sarcomere contraction, muscle fibre damages, and myofibril fragmentation, such as hydrolyzing the proteins in the muscle fibres, weakening the cell structure, releasing the ions, and activating calcium-activating enzymes [172]. HPP treatment (300 MPa, 20 min, 20 °C) reduced (34.78%) the sheer force of goose breast [173]. An HPP of almost 175 MPa decreased the sheer force of hot-boned beef and improved the eating quality. Thus, the moderate pressure treatment of pre-rigour meat seems to have potential since the meat was tender and looked normal [31]. According to Ma and Ledward [174], the tenderness of pre-rigour meat was significantly improved at about 100–150 MPa.

3.2.2. Food Yield

Limited agro-processing technologies have been a major concern, i.e., that many countries, especially developing ones, lack a nutritious food supply. HPP can tackle this issue. In earlier studies, it was observed that HPP had been used to increase oil yields. Jung and Mahfuz [76] reported that HPP treatment at 200 and 500 MPa enhanced the yield (3.20% and 1.30%) of soybean oil in the presence of aqueous enzymatic extraction (AEE) (protease enzyme). Moreover, when HPP was applied before AEE, a higher oil yield was obtained with ground kernels than with whole kernels at a similar pressure and time [44]. Table 6 illustrates the impact of HPP on food yield.
In research, polyphenols and anthocyanins are extracted from blueberry pomace. The combined US, microwave, and HPP effect was compared with the conventional extraction method [175]. The recovery of polyphenols and anthocyanin increased by about 70% and 40% under HPP, while in the case of conventional treatment, the values were about 53% and 32% [179]. Other research was performed under HPP conditions of 470 MPa, 30 min, with a 55% ethanol solution to extract phenolic compounds from pomegranate (Punica granatum); again, the results showed a marked increase in the yields [73].
HPP treatment enhanced the extraction yields of carotenes from plant materials [180]. HPP at 200–600 MPa for 5 min increased the total carotenoid content [181]. Furthermore, HPP as a pretreatment [182] is highly feasible to enhance the recovery of valuable compounds by being implemented with the solvent [183].

3.3. Environmental Sustainability

Food product quality and safety is the main and most important factor for society and manufacturers and distributors [180]. Commercial food production systems have limitations, such as food losses and quality losses due to a high environmental impact [184]. HPP has less impact on the environment as it utilizes less energy and water and has lower CO2 emissions than conventional thermal processes [185]. Therefore, HPP treatment is considered environmentally friendly compared to conventional processing. HPP contributes to environmental sustainability by providing nutritious, safe, and healthy food with an acceptable shelf life in developed countries [186]. HPP has been proven more effective than MAP (modified atmosphere packaging) from an environmental and an economic perspective [187]. However, due to recent food loss and food waste, environmental sustainability may be threatened [187]. A study of thermal pasteurization of orange juice compared with HPP concluded that it is more environmentally friendly than thermal pasteurization [188].

3.3.1. Reduction of Food Waste

The FAO defined food waste as: “The decrease in quantity or quality of food. Thus, food waste is part of food loss and refers to discarding or alternative (non-food) use of food that is safe and nutritious for human consumption along the entire food supply chain, from primary production to the end household consumer-level” [189].
Recently, food waste elimination has gained more attention worldwide [190]. European countries have waste reduction plans to lessen their food waste by almost 30% and 50% by 2025 and 2030 (European Commission, 2018). Food loss and food waste affect food security in two ways. Economically, food loss and food waste reduction allow easy access to food by decreasing food prices. Some studies showed that higher food waste increases food prices [191]. Economic losses from food loss and food waste have no significant impact on the consumer lifestyle in developed countries but are a threat to human livelihood [191]. Although decreasing food waste and food loss contributes to improving food security, it has not been proven to have a clear role in dietary shifts and yield increases [192].
From food waste, different types of compounds such as phenolics and antioxidants are extracted; for example, fruits and vegetables usually contain sufficient bioactive compounds, which are mostly present in peels, seeds, and kernels [193]. It is necessary to rupture the cell wall to obtain natural extracts from the target product [194]. As an effective non-thermal technique, HPP is used for improving the extraction of natural extracts from food waste [195]. Many researchers have discussed the possibility of using HPP to lessen food waste in the food sector. Table 7 shows the valuable compounds extracted from food waste by HPP.
Pumpkin contains carotenoids that can be lost during harsh thermal treatments [200]. HPP can reduce carotenoid losses and natural color losses in pumpkins [199]. In addition, HPP tends to preserve carotenoid contents in vegetables such as pumpkin and other foods [201]. Pumpkin contains different valuable compounds that are extracted, e.g., polysaccharides, pectin proteins, fixed oils, and sterols [202]. HPP-assisted extraction more successfully extracted polyphenols from orange and lemon peels [71]. HPP operated at 200 MPa for 5–10 min to extract phenolic compounds from grape pomace improved the extraction rate up to 16 times compared with enzyme-assisted extraction [198].
Food pollutants indicate the presence of chemical or biological contaminants in natural food. The chemicals commonly related to food pollution are metals, chlorodane, pesticides, and agrochemicals [203]. The pollution load can be reduced by using electricity produced from a clean, renewable energy source. In addition, HPP reduces the usage of cooling systems, which saves 50% electricity consumption [204] and therefore also reduces pollutant emissions.

3.3.2. Effects on Food Packaging

To achieve the best pressure transfer within the food product, food prepared for HPP processing has no gas inclusions, no headspace in the package, and high moisture content. In addition, the type of material used to pack food products must also be appropriate. HPP technology is used in different packaging methods, e.g., whether the product is processed in a container or packaged after processing. The physical and mechanical properties of the packaging material affect the HPP capacity. Packaging materials should have a good pressure tolerance and sealing ability to prevent quality losses after the pressure is applied. The packaging materials must be flexible enough to transmit the pressure without structural damage. Due to pressurization, the food is compressed, and the package must allow this reversible deformation. HPP treatment does not apply to rigid materials such as metal and glass, as they will not withstand the treatment. The space should be less due to efficient utilization of packaging material, and it takes less time to reach the target pressure.
Different packaging materials such as soft polymeric bags, cans, trays, and bottles are commonly used for HPP food products. Vacuum packaging is mostly used for HPP products. HPP may lead to uneven processing. Packaging can reduce unnecessary physical pressure exerted on the external package by gases within the package [205]. Changes occur in the packaging properties of pressure-assisted thermally processed packaging materials used for extended storage time, but how these changes in packages affect product quality are not well understood [206].
In one study, the impact of HPP (680 MPa at 105 °C for 3 min and 680 MPa at 100 °C for 5 min) on polyethylene terephthalate (PET), ethylene-vinyl alcohol copolymer (EVOH), and Nylon6 was studied. There were no changes shown in the melting temperature and the enthalpy of fusion of the EVOH layer [207]. However, Ayvaz et al. [206] applied HPP (600 MPa, 110 °C, 10 min) to baby carrots in three different pouches (nylon, EVOH, ethylene-vinyl acetate (EVA)) and stored the pouches in the dark at suitable conditions (25 and 37 °C for four weeks). High-barrier packaging material, i.e., nnlon/EVOH/EVA, showed less impact.
Dhawan et al. [208] stated that pressure-assisted thermal sterilization (PATS) affected the polymer morphological properties and free volume distributions, which caused a reduction in the gas-barrier properties of polymer packaging materials and maintained the original properties of PATS-processed foods. Therefore, they investigated the effect of HPP at 680 MPa for 5 min at 100 °C on two multilayer EVOH films. HPP enhanced the oxygen and water vapor transmission rates in the two films.
Fleckenstein [209] experimented on several common single layer films (PE-HD, PE-LD, PP-BO, PA6-BO, and PET-BO) and multilayer (PS/PE, PPBO/ PE peel, and PET-BO/PE) films at 600 MPa at 80 °C/90 °C. The impact of HPP on the surface roughness of biaxial-oriented polymer films as single-layer films was studied. There were no changes shown in these films after treatment. HPP hardly affected the surface energy of any stretched, non-stretched, single, or laminated films.
Arfat et al. [210] determined the impact of HPP on polylactic acid (PLA) and polyethylene glycol (PEG) at 300 MPa at 23 °C for 10 min. HPP significantly increased the percentage crystallinity. HPP did not result in any marked alterations to the mechanical properties of the structure. In one study in 2013, the impact of HPP on the thermal and mechanical properties of low-density polyethene (LDPE) films was examined at 200–800 MPa for 5–10 min at 25 and 75 °C [211]. The storage modulus increased with increasing pressure intensity, and the tensile strength increased, but elongation decreased with increasing pressure treatment.

3.4. Economic Sustainability

Economic sustainability includes the efficient utilization of energy and water, with high profits through highly competitive food processing. This will be further explained in terms of HPP treatment.

3.4.1. Yield Efficiency

Product yield is of substantial economic importance to food manufacturers, and HPP improves the yield of products, with effects depending on product type and treatment intensity [212]. HPP was successfully used in the treatment of oysters. HPP denatures the adductor muscle, which enables easy opening of the oyster shell without causing knife damage to the product, so it reduces labor costs and risks associated with hand-shucking [213]. Different food production systems involve the extraction of valuable components [214]. Generally, there are two essential parameters for the extraction of metabolites from cells: the solvent’s diffusion into the cell and the mass transfer of metabolites into the bulk of the extraction medium [215]. Table 8 demonstrates that HPP was helpful to enhance production and profit margins. Nincevic et al. [216] observed that HPP enhanced pectin and polyphenol recovery from tomato peel waste by about 15% compared with conventional techniques.
Phosvitin was successfully extracted from egg yolk at 400 and 600 MPa for 5 and 10 min by HPP [218]. According to Limsangouan et al. [35], the xyloglucan extraction rate was between 51.6% and 53% from tamarind seed using HPP at 250–500 MPa. Strati et al. [221] extracted lycopene from tomato waste using HPP; at 700 MPa for 30 min, the maximum increase in production, about 89.4 mg/g, was observed.

3.4.2. Water Efficiency

Limited water reserves force the government and water authorities to improve water utilization efficiency. The water consumption of every food industry varies. Some food industries, including meat, dairy, and fruit and vegetable processors utilize more water. Bakeries and grain producers, primarily involved in dry processes, are small water users [223]. There are always some benefits of higher water efficiency. For example, the Australian Government surveyed manufacturing groups and reported large savings in total water usage of up to 25, 30, and 60% using basic initiatives such as behavioral changes, water recycling (without conditioning treatment), and water use monitoring, respectively. HPP is used in food industries for energy saving and water saving purposes. It enhances reliability, decreases emissions, results in higher product quality, improves productivity [74], and has less influence on the environment.

3.4.3. Energy Efficiency

For the last 30 years, energy saving has been focused on automating production processes and increasing demand for food safety. The higher levels of hygiene consecutively established as goals subsequently lead to more significant consumption of cold and hot water and an increased number of cleaning cycles in production [204].
Non-thermal technologies utilize less energy than conventional technologies [224]. During HPP, momentary pressure of about 300–700 MPa is transmitted throughout the food products, reducing the processing time and, consequently, energy consumption [75]. Bull et al. [225] reported that HPP (600 MPa, 20 °C, 60 s) reduced microbes in 12 weeks and enhanced the shelf life compared to pasteurization (65 °C, 1 min; 85 °C, 25 s). Another study concluded that HPP (400 and 600 MPa for 5 min at 20 °C) was a better alternative for apple processing than thermal pasteurization (75 °C, 10 min) [226].
HPP requires less energy than chilling and freezing. The energy phase is changed during the HPP process; the heat of vaporization of water is nearly 30% lower than that at atmospheric pressure [227]. Due to water expansion during freezing, pressure increases can lead to lower freezing points [228]. Research indicated that HPP consumed less energy per kg of food than conventional thermal processes. During HPP, heat is conveyed to the foods through convection, conduction, or radiation, causing slow heat transfer rates and increased processing times. In HPP, there is no need for energy to generate heat, and isotactic pressure is applied by utilizing a specific amount of energy throughout the product. Therefore, it is considered that HPP reduces the total energy consumption in the food industry.

4. Limitations of HPP Processing

HPP is a feasible non-thermal technique that preserves food quality, preserves food safety, and increases the shelf life nutritional value but has some limitations in its introduction to the food sector.
The main disadvantage is that it is a costly process due to the high amount of power consumed [229]. Sampedro et al. [230] estimated that HPP was seven times more costly than conventional thermal processing. HPP has been utilizing for more than 100 years in the food sector, but it has limited use due to the high cost and requirement of skilled expertise [231].
In addition, HPP has limitations in killing spores; the combination of heat and HPP (PATS) would solve this problem [232]. Therefore, food with a higher acidic content is preferred for HPP compared with low acid food as it is less stable [233].
Moreover, in a modified HPP treatment method, the product is subjected to compression–decompression cycles with a fixed pressure holding time, but rapid compression and decompression increase the number of cycles in the vessel and subject the vessel material to significant stress and the risk of premature failure [60].

5. Conclusions

Consumers’ demand for fresh and nutritional food has resulted in an increase in non-thermal processing techniques, which can retain authenticity and, at the same time, can ensure the safety of food. The combination of mild heat treatment with some of these non-thermal methods also delivered promising results. HPP used alone or in combination with other treatments stops microbial growth and improves macromolecules, increasing the chemical reactions and shelf life of food products. HPP responds to the demands of consumers for higher sensory and quality attributes, such as taste, extended storage, highly nutritious, healthy, and eco-friendly processing. HPP produces less waste than thermal processing. It also enhances energy efficiency and water efficiency in food products and causes no toxic gas emissions. HPP is a beneficial and sustainable technique for food production. However, despite the ability of HPP to reduce foodborne pathogens and enhance nutritional qualities and shelf life, the high fixed costs limit its use in food processing industries. Further research is required in terms of cost optimization and scalability of HPP to fulfill the needs of both the industry and the consumer.

Author Contributions

Conceptualization: E.R. and R.M.A. Draft Preparation: B.G.N., K.M. and R.N.A. Tables and Figures preparation: P.T., M.U.S., N.W. and A.N. Methodology: R.N.A. Writing-review and editing: N.W., A.N., P.T., E.R., M.I.-U.-R. and R.M.A. Supervision: R.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Naveed, R.; Abdul-Malek, Z.; Roobab, U.; Abdullah, M.; Naderipour, A.; Imran, M.; Bekhit, A.E.; Liu, Z.; Aadil, R.M. Pulsed Electric Field: A Potential Alternative towards a Sustainable Food Processing. Trends Food Sci. Technol. 2021, 111, 43–54. [Google Scholar] [CrossRef]
  2. Loveday, S.M. Food Proteins: Technological, Nutritional, and Sustainability Attributes of Traditional and Emerging Proteins. Annu. Rev. Food Sci. Technol. 2019, 10, 311–339. [Google Scholar] [CrossRef]
  3. Mak, T.M.W.; Xiong, X.; Tsang, D.C.W.; Yu, I.K.M.; Poon, C.S. Sustainable Food Waste Management towards Circular Bioeconomy: Policy Review, Limitations and Opportunities. Bioresour. Technol. 2020, 297, 122497. [Google Scholar] [CrossRef]
  4. Castro-Muñoz, R.; Boczkaj, G.; Gontarek, E.; Cassano, A.; Fíla, V. Membrane Technologies Assisting Plant-Based and Agro-Food by-Products Processing: A Comprehensive Review. Trends Food Sci. Technol. 2020, 95, 219–232. [Google Scholar] [CrossRef]
  5. Arshad, R.N.; Abdul-Malek, Z.; Roobab, U.; Qureshi, M.I.; Khan, N.; Ahmad, M.H.; Liu, Z.W.; Aadil, R.M. Effective Valorization of Food Wastes and By-Products through Pulsed Electric Field: A Systematic Review. J. Food Process Eng. 2021, 44, e13629. [Google Scholar] [CrossRef]
  6. Režek Jambrak, A.; Vukušić, T.; Donsi, F.; Paniwnyk, L.; Djekic, I. Three Pillars of Novel Nonthermal Food Technologies: Food Safety, Quality, and Environment. J. Food Qual. 2018, 2018, 8619707. [Google Scholar] [CrossRef]
  7. Mújica-Paz, H.; Valdez-Fragoso, A.; Samson, C.T.; Welti-Chanes, J.; Torres, A. High-Pressure Processing Technologies for the Pasteurization and Sterilization of Foods. Food Bioprocess Technol. 2011, 4, 969–985. [Google Scholar] [CrossRef]
  8. 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]
  9. Aadil, R.M.; Khalil, A.A.; Rehman, A.; Khalid, A.; Inam-ur-Raheem, M.; Karim, A.; Gill, A.A.; Abid, M.; Afraz, M.T. Assessing the Impact of Ultra-Sonication and Thermo-Ultrasound on Antioxidant Indices and Polyphenolic Profile of Apple-Grape Juice Blend. J. Food Process. Preserv. 2020, 44, e14406. [Google Scholar] [CrossRef]
  10. Liu, Z.W.; Liu, L.J.; Zhou, Y.X.; Tan, Y.C.; Cheng, J.H.; Bekhit, A.E.D.; Inam-Ur-Raheem, M.; Aadil, R.M. Dielectric-Barrier Discharge (DBD) Plasma Treatment Reduces IgG Binding Capacity of β-Lactoglobulin by Inducing Structural Changes. Food Chem. 2021, 358, 129821. [Google Scholar] [CrossRef]
  11. Nor Hasni, H.; Koh, P.C.; Noranizan, M.A.; Megat Mohd Tahir, P.N.F.; Mohamad, A.; Limpot, N.; Hamid, N.; Aadil, R.M. High-Pressure Processing Treatment for Ready-to-Drink Sabah Snake Grass Juice. J. Food Process. Preserv. 2020, 44, e14508. [Google Scholar] [CrossRef]
  12. 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] [PubMed]
  13. Woldemariam, H.W.; Emire, S.A. High Pressure Processing of Foods for Microbial and Mycotoxins Control: Current Trends and Future Prospects. Cogent Food Agric. 2019, 5, 1622184. [Google Scholar] [CrossRef]
  14. Jermann, C.; Koutchma, T.; Margas, E.; Leadley, C.; Ros-Polski, V. Mapping Trends in Novel and Emerging Food Processing Technologies around the World. Innov. Food Sci. Emerg. Technol. 2015, 31, 14–27. [Google Scholar] [CrossRef]
  15. Roobab, U.; Khan, A.W.; Lorenzo, J.M.; Arshad, R.N.; Chen, B.-R.; Zeng, X.-A.; Bekhit, A.E.-D.; Suleman, R.; Aadil, R.M. A Systematic Review of Clean-Label Alternatives to Synthetic Additives in Raw and Processed Meat with a Special Emphasis on High-Pressure Processing (2018–2021). Food Res. Int. 2021, 150, 110792. [Google Scholar] [CrossRef]
  16. Huang, H.W.; Wu, S.J.; Lu, J.K.; Shyu, Y.T.; Wang, C.Y. Current Status and Future Trends of High-Pressure Processing in Food Industry. Food Control 2017, 72, 1–8. [Google Scholar] [CrossRef]
  17. Raghubeer, E.V.; Phan, B.N.; Onuoha, E.; Diggins, S.; Aguilar, V.; Swanson, S.; Lee, A. The Use of High-Pressure Processing (HPP) to Improve the Safety and Quality of Raw Coconut (Cocos Nucifera L.) Water. Int. J. Food Microbiol. 2020, 331, 108697. [Google Scholar] [CrossRef]
  18. Sazonova, S.; Galoburda, R.; Gramatina, I. Application of High-Pressure Processing for Safety and Shelf-Life Quality of Meat—A Review. In Proceedings of the 11th Baltic Conference on Food Science and Technology, Jelgava, Latvia, 27–28 April 2017. [Google Scholar] [CrossRef]
  19. Hite, B.H. The Effect of Pressure in the Preservation of Milk: A Preliminary Report; West Virginia University: Morgantown, WV, USA, 1899. [Google Scholar] [CrossRef]
  20. Sukmanov, V.; Hanjun, M.; Li, Y. Effect of High Pressure Processing on Meat and Meat Products. A Review. Ukr. Food J. 2019, 8, 448–469. [Google Scholar] [CrossRef]
  21. Barba, F.J.; Esteve, M.J.; Frígola, A. High Pressure Treatment Effect on Physicochemical and Nutritional Properties of Fluid Foods During Storage: A Review. Compr. Rev. Food Sci. Food Saf. 2012, 11, 307–322. [Google Scholar] [CrossRef]
  22. Misra, N.N.; Koubaa, M.; Roohinejad, S.; Juliano, P.; Alpas, H.; Inácio, R.S.; Saraiva, J.A.; Barba, F.J. Landmarks in the Historical Development of Twenty First Century Food Processing Technologies. Food Res. Int. 2017, 97, 318–339. [Google Scholar] [CrossRef]
  23. de Oliveira, T.L.C.; Ramos, A.L.S.; Ramos, E.M.; Piccoli, R.H.; Cristianini, M. Preservation of Foods by High Pressure Processing. Trends Food Sci. Technol. 2015, 45, 60–85. [Google Scholar] [CrossRef]
  24. Zhang, H.; Tikekar, R.V.; Ding, Q.; Gilbert, A.R.; Wimsatt, S.T. Inactivation of Foodborne Pathogens by the Synergistic Combinations of Food Processing Technologies and Food-Grade Compounds. Compr. Rev. Food Sci. Food Saf. 2020, 19, 2110–2138. [Google Scholar] [CrossRef] [PubMed]
  25. Roobab, U.; Inam-Ur-Raheem, M.; Khan, A.W.; Arshad, R.N.; Zeng, X.; Aadil, R.M. Innovations in High-Pressure Technologies for the Development of Clean Label Dairy Products: A Review. Food Rev. Int. 2021. [Google Scholar] [CrossRef]
  26. Chauhan, O.P. Non-Thermal Processing of Foods; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  27. Daher, D.; Le Gourrierec, S.; Pérez-Lamela, C. Effect of High Pressure Processing on the Microbial Inactivation in Fruit Preparations and Other Vegetable Based Beverages. Agriculture 2017, 7, 72. [Google Scholar] [CrossRef] [Green Version]
  28. Modi, M.T.K. Emerging Technologies in Food Science. Available online: http://worldfoodscience.com/content/sib-emerging-and-new-technologies-food-science-and-technology (accessed on 8 August 2021).
  29. Campus, M. High Pressure Processing of Meat, Meat Products and Seafood. Food Eng. Rev. 2010, 2, 256–273. [Google Scholar] [CrossRef]
  30. Bolumar, T.; Orlien, V.; Bak, K.H.; Aganovic, K.; Sikes, A.; Guyon, C.; Stübler, A.-S.; de Lamballerie, M.; Hertel, C.; Brüggemann, D.A. High-Pressure Processing (HPP) of Meat Products: Impact on Quality and Applications. In Present and Future of High Pressure Processing; Elsevier: Amsterdam, The Netherlands, 2020; pp. 221–244. [Google Scholar] [CrossRef]
  31. Morton, J.D.; Pearson, R.G.; Lee, H.Y.Y.; Smithson, S.; Mason, S.L.; Bickerstaffe, R. High Pressure Processing Improves the Tenderness and Quality of Hot-Boned Beef. Meat Sci. 2017, 133, 69–74. [Google Scholar] [CrossRef] [PubMed]
  32. Balamurugan, S.; Gemmell, C.; Lau, A.T.Y.; Arvaj, L.; Strange, P.; Gao, A.; Barbut, S. High Pressure Processing during Drying of Fermented Sausages Can Enhance Safety and Reduce Time Required to Produce a Dry Fermented Product. Food Control 2020, 113, 107224. [Google Scholar] [CrossRef]
  33. Usaga, J.; Acosta, Ó.; Churey, J.J.; Padilla-Zakour, O.I.; Worobo, R.W. Evaluation of High Pressure Processing (HPP) Inactivation of Escherichia Coli O157:H7, Salmonella Enterica, and Listeria Monocytogenes in Acid and Acidified Juices and Beverages. Int. J. Food Microbiol. 2021, 339, 109034. [Google Scholar] [CrossRef]
  34. Kapoor, S.; Singh, M.P.; Vatankhah, H.; Deshwal, G.K.; Ramaswamy, H.S. Production and Quality Improvement of Indian Cottage Cheese (Paneer) Using High Pressure Processing. Innov. Food Sci. Emerg. Technol. 2021, 72, 102746. [Google Scholar] [CrossRef]
  35. Limsangouan, N.; Charunuch, C.; Sastry, S.K.; Srichamnong, W.; Jittanit, W. High Pressure Processing of Tamarind (Tamarindus Indica) Seed for Xyloglucan Extraction. LWT 2020, 134, 110112. [Google Scholar] [CrossRef]
  36. Agcam, E.; Akyıldız, A.; Kamat, S.; Balasubramaniam, V.M. Bioactive Compounds Extraction from the Black Carrot Pomace with Assistance of High Pressure Processing: An Optimization Study. Waste Biomass Valorization 2021, 12, 5959–5977. [Google Scholar] [CrossRef]
  37. Zhao, Y.; Jiang, Y.; Ding, Y.; Wang, D.; Deng, Y. High Hydrostatic Pressure-Assisted Extraction of High-Molecular-Weight Melanoidins from Black Garlic: Composition, Structure, and Bioactive Properties. J. Food Qual. 2019, 2019, 1682749. [Google Scholar] [CrossRef]
  38. Park, C.Y.; Kim, S.; Lee, D.; Park, D.J.; Imm, J.Y. Enzyme and High Pressure Assisted Extraction of Tricin from Rice Hull and Biological Activities of Rice Hull Extract. Food Sci. Biotechnol. 2016, 25, 159. [Google Scholar] [CrossRef]
  39. Uribe, E.; Delgadillo, A.; Giovagnoli-Vicunã, C.; Quispe-Fuentes, I.; Zura-Bravo, L. Extraction Techniques for Bioactive Compounds and Antioxidant Capacity Determination of Chilean Papaya (Vasconcellea Pubescens) Fruit. J. Chem. 2015, 2015, 347532. [Google Scholar] [CrossRef] [Green Version]
  40. Prego, R.; Fidalgo, L.G.; Saraiva, J.A.; Vázquez, M.; Aubourg, S.P. Impact of Prior High-Pressure Processing on Lipid Damage and Volatile Amines Formation in Mackerel Muscle Subjected to Frozen Storage and Canning. LWT 2021, 135, 109957. [Google Scholar] [CrossRef]
  41. De Ancos, B.; Rodrigo, M.J.; Sánchez-Moreno, C.; Pilar Cano, M.; Zacarías, L. Effect of High-Pressure Processing Applied as Pretreatment on Carotenoids, Flavonoids and Vitamin C in Juice of the Sweet Oranges “Navel” and the Red-Fleshed “Cara Cara”. Food Res. Int. 2020, 132, 109105. [Google Scholar] [CrossRef]
  42. Albertos, I.; Martin-Diana, A.B.; Sanz, M.A.; Barat, J.M.; Diez, A.M.; Jaime, I.; Rico, D. Effect of High Pressure Processing or Freezing Technologies as Pretreatment in Vacuum Fried Carrot Snacks. Innov. Food Sci. Emerg. Technol. 2016, 33, 115–122. [Google Scholar] [CrossRef]
  43. Wu, S.; Tong, Y.; Zhang, C.; Zhao, W.; Lyu, X.; Shao, Y.; Yang, R. High Pressure Processing Pretreatment of Chinese Mitten Crab (Eriocheir Sinensis) for Quality Attributes Assessment. Innov. Food Sci. Emerg. Technol. 2021, 73, 102793. [Google Scholar] [CrossRef]
  44. Mat Yusoff, M.; Gordon, M.H.; Ezeh, O.; Niranjan, K. High Pressure Pre-Treatment of Moringa Oleifera Seed Kernels Prior to Aqueous Enzymatic Oil Extraction. Innov. Food Sci. Emerg. Technol. 2017, 39, 129–136. [Google Scholar] [CrossRef]
  45. Kaya, E.C.; Oztop, M.H.; Alpas, H. Effect of High-Pressure Processing (HPP) on Production and Characterization of Chia Seed Oil Nanoemulsions. LWT 2021, 141, 110872. [Google Scholar] [CrossRef]
  46. Razi, S.M.; Motamedzadegan, A.; Matia-Merino, L.; Shahidi, S.A.; Rashidinejad, A. The Effect of PH and High-Pressure Processing (HPP) on the Rheological Properties of Egg White Albumin and Basil Seed Gum Mixtures. Food Hydrocoll. 2019, 94, 399–410. [Google Scholar] [CrossRef]
  47. Chai, H.E.; Sheen, S. Effect of High Pressure Processing, Allyl Isothiocyanate, and Acetic Acid Stresses on Salmonella Survivals, Storage, and Appearance Color in Raw Ground Chicken Meat. Food Control 2021, 123, 107784. [Google Scholar] [CrossRef]
  48. Cartagena, L.; Puértolas, E.; Martínez de Marañón, I. Evolution of Quality Parameters of High Pressure Processing (HPP) Pretreated Albacore (Thunnus Alalunga) during Long-Term Frozen Storage. Innov. Food Sci. Emerg. Technol. 2020, 62, 102334. [Google Scholar] [CrossRef]
  49. Nuñez-Mancilla, Y.; Pérez-Won, M.; Uribe, E.; Vega-Gálvez, A.; Di Scala, K. Osmotic Dehydration under High Hydrostatic Pressure: Effects on Antioxidant Activity, Total Phenolics Compounds, Vitamin C and Colour of Strawberry (Fragaria Vesca). LWT-Food Sci. Technol. 2013, 52, 151–156. [Google Scholar] [CrossRef]
  50. Balakrishna, A.K.; Wazed, M.A.; Farid, M. A Review on the Effect of High Pressure Processing (HPP) on Gelatinization and Infusion of Nutrients. Molecules 2020, 25, 2369. [Google Scholar] [CrossRef]
  51. Balasubramaniam, V.M.; Barbosa-Cánovas, G.V.; Lelieveld, H.L.M. High-Pressure Processing Equipment for the Food Industry. In High Pressure Processing of Food; Springer: New York, NY, USA, 2016; pp. 39–65. [Google Scholar] [CrossRef]
  52. FerstI, C.; FerstI, P. Food for Thought High Pressure Processing. Available online: https://www.academia.edu/30986916/FOOD_FOR_THOUGHT_HIGH_PRESSURE_PROCESSING_INSIGHTS_ON_TECHNOLOGY_AND_REGULATORY_REQUIREMENTS_Introduction_and_Background%22%3Ehttps:/www.academia.edu/30986916/FOOD_FOR_THOUGHT_HIGH_PRESSURE_PROCESSING_INSIGHTS_ON_TECHNOLOGY_AND_REGULATORY_REQUIREMENTS_Introduction_and_Background (accessed on 9 August 2021).
  53. Tao, Y.; Sun, D.-W.; Hogan, E.; Kelly, A.L. High-Pressure Processing of Foods: An Overview. In Emerging Technologies for Food Processing, 2nd ed.; Academic Press: Cambridge, MA, USA, 2014; pp. 3–24. [Google Scholar] [CrossRef]
  54. Cavender, G.A. Continuous High Pressure Processing of Liquid Foods: An Analysis of Physical, Structural and Microbial Effects. Ph.D. Thesis, University of Georgia, Athens, GA, USA, 2011. [Google Scholar]
  55. Chawla, R.; Patil, G.R.; Singh, A.K. High Hydrostatic Pressure Technology in Dairy Processing: A Review. J. Food Sci. Technol. 2011, 48, 260. [Google Scholar] [CrossRef] [Green Version]
  56. Balasubramaniam, V.M.; Farkas, D.; Turek, E.J. Preserving Foods through High-Pressure Processing. Food Technol. 2008, 62, 32–38. [Google Scholar]
  57. Gupta, R.; Mikhaylenko, G.; Balasubramaniam, V.M.; Tang, J. Combined Pressure–Temperature Effects on the Chemical Marker (4-Hydroxy-5-Methyl- 3(2H)-Furanone) Formation in Whey Protein Gels. LWT-Food Sci. Technol. 2011, 10, 2141–2146. [Google Scholar] [CrossRef]
  58. Balasubramaniam, V.M.; Ting, E.Y.; Stewart, C.M.; Robbins, J.A. Recommended Laboratory Practices for Conducting High-Pressure Microbial Inactivation Experiments. Innov. Food Sci. Emerg. Technol. 2004, 5, 299–306. [Google Scholar] [CrossRef]
  59. Avsaroglu, M.D.; Bozoglu, F.; Alpas, H.; Largeteau, A.; Demazeau, G. Use of Pulsed-High Hydrostatic Pressure Treatment to Decrease Patulin in Apple Juice. High Press. Res. 2015, 35, 214–222. [Google Scholar] [CrossRef]
  60. Balasubramaniam, V.M.; Martínez-Monteagudo, S.I.; Gupta, R. Principles and Application of High Pressure–Based Technologies in the Food Industry. Annu. Rev. Food Sci. Technol. 2015, 6, 435–462. [Google Scholar] [CrossRef] [PubMed]
  61. Kalagatur, N.K.; Kamasani, J.R.; Mudili, V.; Krishna, K.; Chauhan, O.P.; Sreepathi, M.H. Effect of High Pressure Processing on Growth and Mycotoxin Production of Fusarium Graminearum in Maize. Food Biosci. 2018, 21, 53–59. [Google Scholar] [CrossRef]
  62. Evelyn; Silva, F.V.M. Resistance of Byssochlamys Nivea and Neosartorya Fischeri Mould Spores of Different Age to High Pressure Thermal Processing and Thermosonication. J. Food Eng. 2017, 201, 9–16. [Google Scholar] [CrossRef]
  63. Terefe, N.S.; Buckow, R.; Versteeg, C. Quality-Related Enzymes in Fruit and Vegetable Products: Effects of Novel Food Processing Technologies, Part 1: High-Pressure Processing. Crit. Rev. Food Sci. Nutr. 2014, 54, 24–63. [Google Scholar] [CrossRef] [PubMed]
  64. Milani, E.; Silva, F.V. Comparing High Pressure Thermal Processing and Thermosonication with Thermal Processing for the Inactivation of Bacteria, Moulds, and Yeasts Spores in Foods. J. Food Eng. 2017, 214, 90–96. [Google Scholar] [CrossRef]
  65. FAO. Sustainable Healthy Diets Guiding Principles; FAO: Rome, Italy, 2020. [Google Scholar] [CrossRef]
  66. Petrescu, D.C.; Vermeir, I.; Petrescu-Mag, R.M. Consumer Understanding of Food Quality, Healthiness, and Environmental Impact: A Cross-National Perspective. Int. J. Environ. Res. Public Health 2020, 17, 169. [Google Scholar] [CrossRef] [Green Version]
  67. Muntean, M.-V.; Marian, O.; Barbieru, V.; Cătunescu, G.M.; Ranta, O.; Drocas, I.; Terhes, S. High Pressure Processing in Food Industry—Characteristics and Applications. Agric. Agric. Sci. Procedia 2016, 10, 377–383. [Google Scholar] [CrossRef] [Green Version]
  68. Roobab, U.; Shabbir, M.A.; Khan, A.W.; Arshad, R.N.; Bekhit, A.E.D.; Zeng, X.A.; Inam-Ur-Raheem, M.; Aadil, R.M. High-Pressure Treatments for Better Quality Clean-Label Juices and Beverages: Overview and Advances. LWT 2021, 149, 111828. [Google Scholar] [CrossRef]
  69. Akhmazillah, M.F.N.; Farid, M.M.; Silva, F.V.M. High Pressure Processing (HPP) of Honey for the Improvement of Nutritional Value. Innov. Food Sci. Emerg. Technol. 2013, 20, 59–63. [Google Scholar] [CrossRef]
  70. Yildiz, S.; Pokhrel, P.R.; Unluturk, S.; Barbosa-Cánovas, G.V. Identification of Equivalent Processing Conditions for Pasteurization of Strawberry Juice by High Pressure, Ultrasound, and Pulsed Electric Fields Processing. Innov. Food Sci. Emerg. Technol. 2019, 57, 102195. [Google Scholar] [CrossRef]
  71. Casquete, R.; Castro, S.M.; Martín, A.; Ruiz-Moyano, S.; Saraiva, J.A.; Córdoba, M.G.; Teixeira, P. Evaluation of the Effect of High Pressure on Total Phenolic Content, Antioxidant and Antimicrobial Activity of Citrus Peels. Innov. Food Sci. Emerg. Technol. 2015, 31, 37–44. [Google Scholar] [CrossRef]
  72. Dobiáš, J.; Voldřich, M.; Marek, M.; Chudáčková, K. Changes of Properties of Polymer Packaging Films during High Pressure Treatment. J. Food Eng. 2004, 61, 545–549. [Google Scholar] [CrossRef]
  73. Alexandre, E.M.C.; Araújo, P.; Duarte, M.F.; de Freitas, V.; Pintado, M.; Saraiva, J.A. Experimental Design, Modeling, and Optimization of High-Pressure-Assisted Extraction of Bioactive Compounds from Pomegranate Peel. Food Bioprocess Technol. 2017, 10, 886–900. [Google Scholar] [CrossRef]
  74. Masanet, E.; Masanet, E.; Worrell, E.; Graus, W.; Galitsky, C. Energy Efficiency Improvement and Cost Saving Opportunities for the Fruit and Vegetable Processing Industry. An ENERGY STAR Guide for Energy and Plant Managers; Lawrence Berkeley National Lab. (LBNL): Berkeley, CA, USA, 2008. [CrossRef] [Green Version]
  75. Barba, F.J.; Terefe, N.S.; Buckow, R.; Knorr, D.; Orlien, V. New Opportunities and Perspectives of High Pressure Treatment to Improve Health and Safety Attributes of Foods. A Review. Food Res. Int. 2015, 77, 725–742. [Google Scholar] [CrossRef]
  76. Jung, S.; Mahfuz, A.A. Low Temperature Dry Extrusion and High-Pressure Processing Prior to Enzyme-Assisted Aqueous Extraction of Full Fat Soybean Flakes. Food Chem. 2009, 114, 947–954. [Google Scholar] [CrossRef]
  77. Szczepańska, J.; Barba, F.J.; Skąpska, S.; Marszałek, K. High Pressure Processing of Carrot Juice: Effect of Static and Multi-Pulsed Pressure on the Polyphenolic Profile, Oxidoreductases Activity and Colour. Food Chem. 2020, 307, 125549. [Google Scholar] [CrossRef]
  78. Putnik, P.; Lorenzo, J.M.; Barba, F.J.; Roohinejad, S.; Jambrak, A.R.; Granato, D.; Montesano, D.; Kovačević, D.B. Novel Food Processing and Extraction Technologies of High-Added Value Compounds from Plant Materials. Foods 2018, 7, 106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Xi, J.; Luo, S. The Mechanism for Enhancing Extraction of Ferulic Acid from Radix Angelica Sinensis by High Hydrostatic Pressure. Sep. Purif. Technol. 2016, 165, 208–213. [Google Scholar] [CrossRef]
  80. Briones-Labarca, V.; Plaza-Morales, M.; Giovagnoli-Vicuña, C.; Jamett, F. High Hydrostatic Pressure and Ultrasound Extractions of Antioxidant Compounds, Sulforaphane and Fatty Acids from Chilean Papaya (Vasconcellea Pubescens) Seeds: Effects of Extraction Conditions and Methods. LWT-Food Sci. Technol. 2015, 60, 525–534. [Google Scholar] [CrossRef]
  81. Li, J.; Sun, W.; Ramaswamy, H.S.; Yu, Y.; Zhu, S.; Wang, J.; Li, H. High Pressure Extraction of Astaxanthin from Shrimp Waste (Penaeus Vannamei Boone): Effect on Yield and Antioxidant Activity. J. Food Process Eng. 2017, 40, e12353. [Google Scholar] [CrossRef]
  82. Roselló-Soto, E.; Poojary, M.M.; Barba, F.J.; Koubaa, M.; Lorenzo, J.M.; Mañes, J.; Moltó, J.C. Thermal and Non-Thermal Preservation Techniques of Tiger Nuts’ Beverage “Horchata de Chufa”. Implications for Food Safety, Nutritional and Quality Properties. Food Res. Int. 2018, 105, 945–951. [Google Scholar] [CrossRef] [PubMed]
  83. Roobab, U.; Aadil, R.M.; Madni, G.M.; Bekhit, A.E.D. Impact of nonthermal technologies on microbiological quality of juices: A review. Compr. Rev. Food Sci. Food Saf. 2018, 17, 437–457. [Google Scholar] [CrossRef]
  84. Woolf, A.B.; Wibisono, R.; Farr, J.; Hallett, I.; Richter, L.; Oey, I.; Wohlers, M.; Zhou, J.; Fletcher, G.C.; Requejo-Jackman, C. Effect of High Pressure Processing on Avocado Slices. Innov. Food Sci. Emerg. Technol. 2013, 18, 65–73. [Google Scholar] [CrossRef]
  85. Khan, M.A.; Ali, S.; Yang, H.; Kamboh, A.A.; Ahmad, Z.; Tume, R.K.; Zhou, G. Improvement of Color, Texture and Food Safety of Ready-to-Eat High Pressure-Heat Treated Duck Breast. Food Chem. 2018, 277, 646–654. [Google Scholar] [CrossRef] [PubMed]
  86. Joint FAO/WHO Codex Alimentarius Commission; Food and Agriculture Organization of the United Nations; World Health Organization. Codex Alimentarius: Food Hygiene Basic Texts; FAO: Rome, Italy, 2001. [Google Scholar]
  87. Kök, M.S. Application of Food Safety Management Systems (ISO 22000/HACCP) in the Turkish Poultry Industry: A Comparison Based on Enterprise Size. J. Food Prot. 2009, 72, 2221–2225. [Google Scholar] [CrossRef] [PubMed]
  88. Simonin, H.; Duranton, F.; de Lamballerie, M. New Insights into the High-Pressure Processing of Meat and Meat Products. Compr. Rev. Food Sci. Food Saf. 2012, 11, 285–306. [Google Scholar] [CrossRef]
  89. Teixeira, P.; Kolomeytseva, M.; Silva, J.; Castro, S.M. High Hydrostatic Pressure Applied to Ready-to-Eat Meat Products Focus on Listeria Monocytogenes Inactivation. Available online: https://www.researchgate.net/publication/266386629_High_Hydrostatic_Pressure_Applied_to_Ready-to-eat_Meat_Products_Focus_on_Listeria_monocytogenes_inactivation (accessed on 20 September 2021).
  90. Sridhar, A.; Ponnuchamy, M.; Kumar, P.S.; Kapoor, A. Food Preservation Techniques and Nanotechnology for Increased Shelf Life of Fruits, Vegetables, Beverages and Spices: A Review. Environ. Chem. Lett. 2021, 19, 1715–1735. [Google Scholar] [CrossRef] [PubMed]
  91. Bolumar, T.; Orlien, V.; Sikes, A.; Aganovic, K.; Bak, K.H.; Guyon, C.; Stübler, A.-S.; de Lamballerie, M.; Hertel, C.; Brüggemann, D.A. High-Pressure Processing of Meat: Molecular Impacts and Industrial Applications. Compr. Rev. Food Sci. Food Saf. 2021, 20, 332–368. [Google Scholar] [CrossRef] [PubMed]
  92. Ferreira, M.; Almeida, A.; Delgadillo, I.; Saraiva, J.; Cunha, Â. Susceptibility of Listeria Monocytogenes to High Pressure Processing: A Review. Food Rev. Int. 2016, 32, 377–399. [Google Scholar] [CrossRef]
  93. Chan, J.T.Y.; Omana, D.A.; Betti, M. Application of High Pressure Processing to Improve the Functional Properties of Pale, Soft, and Exudative (PSE)-like Turkey Meat. Innov. Food Sci. Emerg. Technol. 2011, 12, 216–225. [Google Scholar] [CrossRef]
  94. Moussa-Ayoub, T.E.; Jäger, H.; Knorr, D.; El-Samahy, S.K.; Kroh, L.W.; Rohn, S. Impact of Pulsed Electric Fields, High Hydrostatic Pressure, and Thermal Pasteurization Selected Characteristics of Opuntia Dillenii Cactus Juice. LWT-Food Sci. Technol. 2017, 79, 534–542. [Google Scholar] [CrossRef]
  95. González-Angulo, M.; Clauwers, C.; Harastani, R.; Tonello, C.; Jaime, I.; Rovira, J.; Michiels, C.W. Evaluation of Factors Influencing the Growth of Non-Toxigenic Clostridium Botulinum Type E and Clostridium Sp. in High-Pressure Processed and Conditioned Tender Coconut Water from Thailand. Food Res. Int. 2020, 134, 109278. [Google Scholar] [CrossRef] [PubMed]
  96. Gouvea, F.S.; Padilla-Zakour, O.I.; Worobo, R.W.; Xavier, B.M.; Walter, E.H.M.; Rosenthal, A. Effect of High-Pressure Processing on Bacterial Inactivation in Açaí Juices with Varying PH and Soluble Solids Content. Innov. Food Sci. Emerg. Technol. 2020, 66, 102490. [Google Scholar] [CrossRef]
  97. Jiao, R.; Gao, J.; Li, Y.; Zhang, X.; Zhang, M.; Ye, Y.; Wu, Q.; Fan, H. Short Communication: Effects of High-Pressure Processing on the Inactivity of Cronobacter Sakazakii in Whole Milk and Skim Milk Samples. J. Dairy Sci. 2016, 99, 7881–7885. [Google Scholar] [CrossRef] [Green Version]
  98. Liu, F.; Zhang, X.; Zhao, L.; Wang, Y.; Liao, X. Potential of High-Pressure Processing and High-Temperature/Short-Time Thermal Processing on Microbial, Physicochemical and Sensory Assurance of Clear Cucumber Juice. Innov. Food Sci. Emerg. Technol. 2016, 34, 51–58. [Google Scholar] [CrossRef]
  99. Nasiłowska, J.; Sokołowska, B.; Fonberg-Broczek, M. Long-Term Storage of Vegetable Juices Treated by High Hydrostatic Pressure: Assurance of the Microbial Safety. Biomed. Res. Int. 2018, 2018, 7389381. [Google Scholar] [CrossRef] [PubMed]
  100. Stratakos, A.C.; Inguglia, E.S.; Linton, M.; Tollerton, J.; Murphy, L.; Corcionivoschi, N.; Koidis, A.; Tiwari, B.K. Effect of High Pressure Processing on the Safety, Shelf Life and Quality of Raw Milk. Innov. Food Sci. Emerg. Technol. 2019, 52, 325–333. [Google Scholar] [CrossRef] [Green Version]
  101. Van Luong, T.S.; Moir, C.; Chandry, P.S.; Pinfold, T.; Olivier, S.; Broussolle, V.; Bowman, J.P. Combined High Pressure and Heat Treatment Effectively Disintegrates Spore Membranes and Inactivates Alicyclobacillus Acidoterrestris Spores in Acidic Fruit Juice Beverage. Innov. Food Sci. Emerg. Technol. 2020, 66, 102523. [Google Scholar] [CrossRef]
  102. van Wyk, S.; Silva, F.V.M. High Pressure Processing Inactivation of Brettanomyces Bruxellensis in Seven Different Table Wines. Food Control 2017, 81, 1–8. [Google Scholar] [CrossRef]
  103. Hartyáni, P.; Dalmadi, I.; Knorr, D. Electronic Nose Investigation of Alicyclobacillus Acidoterrestris Inoculated Apple and Orange Juice Treated by High Hydrostatic Pressure. Food Control 2013, 32, 262–269. [Google Scholar] [CrossRef]
  104. Hiremath, N.D.; Ramaswamy, H.S. High-Pressure Destruction Kinetics of Spoilage and Pathogenic Microorganisms in Mango Juice. J. Food Process. Preserv. 2012, 36, 113–125. [Google Scholar] [CrossRef]
  105. Huang, Y.; Ye, M.; Chen, H. Inactivation of Escherichia Coli O157: H7 and Salmonella Spp. in Strawberry Puree by High Hydrostatic Pressure with/without Subsequent Frozen Storage. Int. J. Food Microbiol. 2013, 160, 337–343. [Google Scholar] [CrossRef] [PubMed]
  106. Shahbaz, H.M.; Yoo, S.; Seo, B.; Ghafoor, K.; Kim, J.U.; Lee, D.-U.; Park, J. Combination of TiO2-UV Photocatalysis and High Hydrostatic Pressure to Inactivate Bacterial Pathogens and Yeast in Commercial Apple Juice. Food Bioprocess Technol. 2016, 9, 182–190. [Google Scholar] [CrossRef]
  107. Syed, Q.A.; Buffa, M.; Guamis, B.; Saldo, J. Effect of Compression and Decompression Rates of High Hydrostatic Pressure on Inactivation of Staphylococcus Aureus in Different Matrices. Food Bioprocess Technol. 2014, 7, 1202–1207. [Google Scholar] [CrossRef]
  108. Sehrawat, R.; Kaur, B.P.; Nema, P.K.; Tewari, S.; Kumar, L. Microbial Inactivation by High Pressure Processing: Principle, Mechanism and Factors Responsible. Food Sci. Biotechnol. 2021, 30, 19–35. [Google Scholar] [CrossRef]
  109. Rendueles, E.; Omer, M.K.; Alvseike, O.; Alonso-Calleja, C.; Capita, R.; Prieto, M. Microbiological Food Safety Assessment of High Hydrostatic Pressure Processing: A Review. LWT-Food Sci. Technol. 2011, 44, 1251–1260. [Google Scholar] [CrossRef]
  110. García-Gimeno, R.M.; Izquierdo, G.D.P. High Hydrostatic Pressure Treatment of Meat Products. In Food Processing; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef] [Green Version]
  111. Huang, C.Y.; Sheen, S.; Sommers, C.; Sheen, L.Y. Modeling the Survival of Escherichia Coli O157:H7 under Hydrostatic Pressure, Process Temperature, Time and Allyl Isothiocyanate Stresses in Ground Chicken Meat. Front. Microbiol. 2018, 9, 1871. [Google Scholar] [CrossRef]
  112. Koutchma, T. Adapting High Hydrostatic Pressure (HPP) for Food Processing Operations, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2014; Available online: https://www.elsevier.com/books/adapting-high-hydrostatic-pressure-hpp-for-food-processing-operations/koutchma/978-0-12-420091-3 (accessed on 3 July 2021).
  113. Pérez-Santaescolástica, C.; Carballo, J.; Fulladosa, E.; Munekata, P.E.S.; Bastianello Campagnol, P.C.; Gómez, B.; Lorenzo, J.M. Influence of High-Pressure Processing at Different Temperatures on Free Amino Acid and Volatile Compound Profiles of Dry-Cured Ham. Food Res. Int. 2019, 116, 49–56. [Google Scholar] [CrossRef]
  114. Tapia, M.S.; Alzamora, S.M.; Chirife, J. Effects of Water Activity (aw) on Microbial Stability as a Hurdle in Food Preservation. In Water Activity in Foods; Wiley: Hoboken, NJ, USA, 2020; pp. 323–355. [Google Scholar] [CrossRef]
  115. Dash, K.K.; Balasubramaniam, V.M.; Kamat, S. High Pressure Assisted Osmotic Dehydrated Ginger Slices. J. Food Eng. 2019, 247, 19–29. [Google Scholar] [CrossRef]
  116. Cheng, L.; Zhu, Z.; Sun, D.W. Impacts of High Pressure Assisted Freezing on the Denaturation of Polyphenol Oxidase. Food Chem. 2021, 335, 127485. [Google Scholar] [CrossRef]
  117. Otero, L.; Martino, M.; Zaritzky, N.; Solas, M.; Sanz, P.D. Preservation of Microstructure in Peach and Mango during High-Pressure-Shift Freezing. J. Food Sci. 2000, 65, 466–470. [Google Scholar] [CrossRef] [Green Version]
  118. Cui, Y.; Xuan, X.; Ling, J.; Liao, X.; Zhang, H.; Shang, H.; Lin, X. Effects of High Hydrostatic Pressure-Assisted Thawing on the Physicohemical Characteristics of Silver Pomfret (Pampus Argenteus). Food Sci. Nutr. 2019, 7, 1573–1583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Préstamo, G.; Palomares, L.; Sanz, P. Broccoli (Brasica Oleracea) Treated under Pressure-Shift Freezing Process. Eur. Food Res. Technol. 2004, 219, 598–604. [Google Scholar] [CrossRef] [Green Version]
  120. Mastovska, K. Modern Analysis of Chemical Contaminants in Food. Available online: https://www.food-safety.com/articles/4460-modern-analysis-of-chemical-contaminants-in-food (accessed on 3 July 2021).
  121. Nerín, C.; Aznar, M.; Carrizo, D. Food Contamination during Food Process. Trends Food Sci. Technol. 2016, 48, 63–68. [Google Scholar] [CrossRef]
  122. Martin, A.; Beutin, L. Characteristics of Shiga Toxin-Producing Escherichia Coli from Meat and Milk Products of Different Origins and Association with Food Producing Animals as Main Contamination Sources. Int. J. Food Microbiol. 2011, 146, 99–104. [Google Scholar] [CrossRef]
  123. Aadil, R.M.; Zeng, X.-A.; Jabbar, S.; Nazir, A.; Mann, A.A.; Khan, M.K.I.; Abdullah, A.; Ramzan, A. Quality Evaluation of Grapefruit Juice by Thermal and High Pressure Processing Treatment. Pak. J. Agric. Res. 2017, 30, 209–309. [Google Scholar] [CrossRef]
  124. Evert-Arriagada, K.; Hernández-Herrero, M.M.; Guamis, B.; Trujillo, A.J. Commercial Application of High-Pressure Processing for Increasing Starter-Free Fresh Cheese Shelf-Life. LWT-Food Sci. Technol. 2014, 55, 498–505. [Google Scholar] [CrossRef]
  125. Georget, E.; Sevenich, R.; Reineke, K.; Mathys, A.; Heinz, V.; Callanan, M.; Rauh, C.; Knorr, D. Inactivation of Microorganisms by High Isostatic Pressure Processing in Complex Matrices: A Review. Innov. Food Sci. Emerg. Technol. 2015, 27, 1–14. [Google Scholar] [CrossRef]
  126. Sevenich, R.; Kleinstueck, E.; Crews, C.; Anderson, W.; Pye, C.; Riddellova, K.; Hradecky, J.; Moravcova, E.; Reineke, K.; Knorr, D. High-Pressure Thermal Sterilization: Food Safety and Food Quality of Baby Food Puree. J. Food Sci. 2014, 79, M230–M237. [Google Scholar] [CrossRef]
  127. Kultur, G.; Misra, N.N.; Barba, F.J.; Koubaa, M.; Gökmen, V.; Alpas, H. Microbial Inactivation and Evaluation of Furan Formation in High Hydrostatic Pressure (HHP) Treated Vegetable-Based Infant Food. Food Res. Int. 2017, 101, 17–23. [Google Scholar] [CrossRef]
  128. Iizuka, T.; Shimizu, A. Removal of Pesticide Residue from Brussels Sprouts by Hydrostatic Pressure. Innov. Food Sci. Emerg. Technol. 2014, 22, 70–75. [Google Scholar] [CrossRef]
  129. Ionel, B. European Regulation in the Veterinary Sanitary and Food Safety Area, a Component of the European Policies on the Safety of Food Products and the Protection of Consumer Interests: A 2007 Retrospective. Part Two: Regulations. Universul Jurid. 2018, 16–19. Available online: https://www.ceeol.com/search/article-detail?id=624697 (accessed on 22 November 2021).
  130. Sharma, A.; Kumar, V.; Shahzad, B.; Tanveer, M.; Sidhu, G.P.S.; Handa, N.; Kohli, S.K.; Yadav, P.; Bali, A.S.; Parihar, R.D.; et al. Worldwide Pesticide Usage and Its Impacts on Ecosystem. SN Appl. Sci. 2019, 1, 1446. [Google Scholar] [CrossRef] [Green Version]
  131. Handford, C.E.; Elliott, C.T.; Campbell, K. A Review of the Global Pesticide Legislation and the Scale of Challenge in Reaching the Global Harmonization of Food Safety Standards. Integr. Environ. Assess. Manag. 2015, 11, 525–536. [Google Scholar] [CrossRef]
  132. Ionel, B. European Regulation in the Veterinary Sanitary and Food Safety Area, a Component of the European Policies on the Safety of Food Products and the Protection of Consumer Interests: A 2007 Retrospective. Part One: The Role of European Institutions in Laying down and Passing Laws Specific to the Veterinary Sanitary and Food Safety Area. Available online: https://www.researchgate.net/publication/316716657_EUROPEAN_REGULATION_IN_THE_VETERINARY_SANITARY_AND_FOOD_SAFETY_AREA_A_COMPONENT_OF_THE_EUROPEAN_POLICIES_ON_THE_SAFETY_OF_FOOD_PRODUCTS_AND_THE_PROTECTION_OF_CONSUMER_INTERESTS_A_2007_RETROSPECTIVE_PA (accessed on 22 November 2021).
  133. Gopal, K.R. High Pressure Processing of Fruits and Vegetable Products: A Review. Int. J. Pure Appl. Biosci. 2017, 5, 680–692. [Google Scholar] [CrossRef]
  134. Pereira, R.N.; Vicente, A.A. Environmental Impact of Novel Thermal and Non-Thermal Technologies in Food Processing. Food Res. Int. 2010, 43, 1936–1943. [Google Scholar] [CrossRef] [Green Version]
  135. Bhilwadikar, T.; Pounraj, S.; Manivannan, S.; Rastogi, N.K.; Negi, P.S. Decontamination of Microorganisms and Pesticides from Fresh Fruits and Vegetables: A Comprehensive Review from Common Household Processes to Modern Techniques. Compr. Rev. Food Sci. Food Saf. 2019, 18, 1003–1038. [Google Scholar] [CrossRef] [Green Version]
  136. Cámara, M.A.; Cermeño, S.; Martínez, G.; Oliva, J. Removal Residues of Pesticides in Apricot, Peach and Orange Processed and Dietary Exposure Assessment. Food Chem. 2020, 325, 126936. [Google Scholar] [CrossRef] [PubMed]
  137. González, N.; Marquès, M.; Nadal, M.; Domingo, J.L. Occurrence of Environmental Pollutants in Foodstuffs: A Review of Organic vs. Conventional Food. Food Chem. Toxicol. 2019, 125, 370–375. [Google Scholar] [CrossRef]
  138. Narenderan, S.T.; Meyyanathan, S.N.; Babu, B. Review of Pesticide Residue Analysis in Fruits and Vegetables. Pre-Treatment, Extraction and Detection Techniques. Food Res. Int. 2020, 133, 109141. [Google Scholar] [CrossRef] [PubMed]
  139. Timmermans, R.; Hayrapetyan, H.; Vollebregt, M.; Dijksterhuis, J. Comparing Thermal Inactivation to a Combined Process of Moderate Heat and High Pressure: Effect on Ascospores in Strawberry Puree. Int. J. Food Microbiol. 2020, 325, 108629. [Google Scholar] [CrossRef] [PubMed]
  140. None, E.; FVM, S. Inactivation of Byssochlamys Nivea Ascospores in Strawberry Puree by High Pressure, Power Ultrasound and Thermal Processing. Int. J. Food Microbiol. 2015, 214, 129–136. [Google Scholar] [CrossRef]
  141. Evelyn; Kim, H.J.; Silva, F.V.M. Modeling the Inactivation of Neosartorya Fischeri Ascospores in Apple Juice by High Pressure, Power Ultrasound and Thermal Processing. Food Control 2016, 59, 530–537. [Google Scholar] [CrossRef]
  142. Huang, H.W.; Yang, B.B.; Wang, C.Y. Effects of High Pressure Processing on Immunoreactivity and Microbiological Safety of Crushed Peanuts. Food Control 2014, 42, 290–295. [Google Scholar] [CrossRef]
  143. Tokuşoǧlu, Ö.; Alpas, H.; Bozoǧlu, F. High Hydrostatic Pressure Effects on Mold Flora, Citrinin Mycotoxin, Hydroxytyrosol, Oleuropein Phenolics and Antioxidant Activity of Black Table Olives. Innov. Food Sci. Emerg. Technol. 2010, 11, 250–258. [Google Scholar] [CrossRef]
  144. Xie, H.; Wen, Y.; Choi, Y.; Zhang, X. Global Trends on Food Security Research: A Bibliometric Analysis. Land 2021, 10, 119. [Google Scholar] [CrossRef]
  145. FAO; IFAD; WFP. The State of Food Insecurity in the World 2015. Meeting the 2015 International Hunger Targets: Taking Stock of Uneven Progress. Policy Support and Governance. Food and Agriculture Organization of the United Nations. Available online: http://www.fao.org/policy-support/tools-and-publications/resources-details/en/c/469455/ (accessed on 13 July 2021).
  146. Chacha, J.S.; Zhang, L.; Ofoedu, C.E.; Suleiman, R.A.; Dotto, J.M.; Roobab, U.; Agunbiade, A.O.; Duguma, H.T.; Mkojera, B.T.; Hossaini, S.M.; et al. Revisiting Non-Thermal Food Processing and Preservation Methods—Action Mechanisms, Pros and Cons: A Technological Update (2016–2021). Foods 2021, 10, 1430. [Google Scholar] [CrossRef] [PubMed]
  147. Escobedo-Avellaneda, Z.; Pateiro-Moure, M.; Chotyakul, N.; Torres, J.A.; Welti-Chanes, J.; Pérez-Lamela, C. Benefits and Limitations of Food Processing by High-Pressure Technologies: Effects on Functional Compounds and Abiotic Contaminants. CYTA-J. Food 2011, 9, 351–364. [Google Scholar] [CrossRef] [Green Version]
  148. García, A.F.; Butz, P.; Bognàr, A.; Tauscher, B. Antioxidative Capacity, Nutrient Content and Sensory Quality of Orange Juice and an Orange-Lemon-Carrot Juice Product after High Pressure Treatment and Storage in Different Packaging. Eur. Food Res. Technol. 2001, 213, 290–296. [Google Scholar] [CrossRef]
  149. Aaby, K.; Grimsbo, I.H.; Hovda, M.B.; Rode, T.M. Effect of High Pressure and Thermal Processing on Shelf Life and Quality of Strawberry Purée and Juice. Food Chem. 2018, 260, 115–123. [Google Scholar] [CrossRef]
  150. Barba, F.J.; Esteve, M.J.; Frigola, A. Ascorbic Acid Is the Only Bioactive That Is Better Preserved by High Hydrostatic Pressure than by Thermal Treatment of a Vegetable Beverage. J. Agric. Food Chem. 2010, 58, 10070–10075. [Google Scholar] [CrossRef]
  151. Kiełczewska, K.; Jankowska, A.; Dąbrowska, A.; Wachowska, M.; Ziajka, J. The Effect of High Pressure Treatment on the Dispersion of Fat Globules and the Fatty Acid Profile of Caprine Milk. Int. Dairy J. 2020, 102, 104607. [Google Scholar] [CrossRef]
  152. Rivas-Cañedo, A.; Martínez-Onandi, N.; Gaya, P.; Nuñez, M.; Picon, A. Effect of High-Pressure Processing and Chemical Composition Lipid Oxidation, Aminopeptidase Activity and Free Amino Acids of Serrano Dry-Cured Ham. Meat Sci. 2021, 172, 108349. [Google Scholar] [CrossRef] [PubMed]
  153. Cadesky, L.; Walkling-Ribeiro, M.; Kriner, K.T.; Karwe, M.V.; Moraru, C.I. Structural Changes Induced by High-Pressure Processing in Micellar Casein and Milk Protein Concentrates. J. Dairy Sci. 2017, 100, 7055–7070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Yang, Y.; Xia, Y.; Wang, G.; Tao, L.; Yu, J.; Ai, L. Effects of Boiling, Ultra-High Temperature and High Hydrostatic Pressure on Free Amino Acids, Flavor Characteristics and Sensory Profiles in Chinese Rice Wine. Food Chem. 2019, 275, 407–416. [Google Scholar] [CrossRef]
  155. Stinco, C.M.; Szczepańska, J.; Marszałek, K.; Pinto, C.A.; Inácio, R.S.; Mapelli-Brahm, P.; Barba, F.J.; Lorenzo, J.M.; Saraiva, J.A.; Meléndez-Martínez, A.J. Effect of High-Pressure Processing on Carotenoids Profile, Colour, Microbial and Enzymatic Stability of Cloudy Carrot Juice. Food Chem. 2019, 299, 125112. [Google Scholar] [CrossRef] [PubMed]
  156. Błaszczak, W.; Latocha, P.; Jeż, M.; Wiczkowski, W. The Impact of High-Pressure Processing on the Polyphenol Profile and Anti-Glycaemic, Anti-Hypertensive and Anti-Cholinergic Activities of Extracts Obtained from Kiwiberry (Actinidia Arguta) Fruits. Food Chem. 2021, 343, 128421. [Google Scholar] [CrossRef]
  157. da Silveira, T.F.F.; Cristianini, M.; Kuhnle, G.G.; Ribeiro, A.B.; Filho, J.T.; Godoy, H.T. Anthocyanins, Non-Anthocyanin Phenolics, Tocopherols and Antioxidant Capacity of Açaí Juice (Euterpe Oleracea) as Affected by High Pressure Processing and Thermal Pasteurization. Innov. Food Sci. Emerg. Technol. 2019, 55, 88–96. [Google Scholar] [CrossRef]
  158. Zhang, W.; Shen, Y.; Li, Z.; Xie, X.; Gong, E.S.; Tian, J.; Si, X.; Wang, Y.; Gao, N.; Shu, C.; et al. Effects of High Hydrostatic Pressure and Thermal Processing on Anthocyanin Content, Polyphenol Oxidase and β-Glucosidase Activities, Color, and Antioxidant Activities of Blueberry (Vaccinium Spp.) Puree. Food Chem. 2021, 342, 128564. [Google Scholar] [CrossRef]
  159. Marszałek, K.; Doesburg, P.; Starzonek, S.; Szczepańska, J.; Woźniak, Ł.; Lorenzo, J.M.; Skaopska, S.; Rzoska, S.; Barba, F.J. Comparative Effect of Supercritical Carbon Dioxide and High Pressure Processing on Structural Changes and Activity Loss of Oxidoreductive Enzymes. J. CO2 Util. 2019, 29, 46–56. [Google Scholar] [CrossRef]
  160. Lu, S.Y.; Chu, Y.L.; Sridhar, K.; Tsai, P.J. Effect of Ultrasound, High-Pressure Processing, and Enzymatic Hydrolysis on Carbohydrate Hydrolyzing Enzymes and Antioxidant Activity of Lemon (Citrus Limon) Flavedo. LWT 2021, 138, 110511. [Google Scholar] [CrossRef]
  161. Zhang, L.; Dai, S.; Brannan, R.G. Effect of High Pressure Processing, Browning Treatments, and Refrigerated Storage on Sensory Analysis, Color, and Polyphenol Oxidase Activity in Pawpaw (Asimina Triloba L.) Pulp. LWT-Food Sci. Technol. 2017, 86, 49–54. [Google Scholar] [CrossRef]
  162. Cao, B.; Fang, L.; Liu, C.; Min, W.; Liu, J. Effects of High Hydrostatic Pressure on the Functional and Rheological Properties of the Protein Fraction Extracted from Pine Nuts. Food Sci. Technol. Int. 2018, 24, 53–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Xia, Q.; Wang, L.; Li, Y. Exploring High Hydrostatic Pressure-Mediated Germination to Enhance Functionality and Quality Attributes of Wholegrain Brown Rice. Food Chem. 2018, 249, 104–110. [Google Scholar] [CrossRef] [PubMed]
  164. Ali, N.; Popović, V.; Koutchma, T.; Warriner, K.; Zhu, Y. Effect of Thermal, High Hydrostatic Pressure, and Ultraviolet-C Processing on the Microbial Inactivation, Vitamins, Chlorophyll, Antioxidants, Enzyme Activity, and Color of Wheatgrass Juice. J. Food Process Eng. 2020, 43, e13036. [Google Scholar] [CrossRef]
  165. Huang, H.W.; Hsu, C.P.; Wang, C.Y. Healthy Expectations of High Hydrostatic Pressure Treatment in Food Processing Industry. J. Food Drug Anal. 2020, 28, 1–13. [Google Scholar] [CrossRef]
  166. Moltó-Puigmartí, C.; Permanyer, M.; Castellote, A.I.; López-Sabater, M.C. Effects of Pasteurisation and High-Pressure Processing on Vitamin C, Tocopherols and Fatty Acids in Mature Human Milk. Food Chem. 2011, 124, 697–702. [Google Scholar] [CrossRef]
  167. Nayak, P.K.; Rayaguru, K.; Radha Krishnan, K. Quality Comparison of Elephant Apple Juices after High-Pressure Processing and Thermal Treatment. J. Sci. Food Agric. 2017, 97, 1404–1411. [Google Scholar] [CrossRef]
  168. Medina-Meza, I.G.; Barnaba, C.; Villani, F.; Barbosa-Cánovas, G.V. Effects of Thermal and High Pressure Treatments in Color and Chemical Attributes of an Oil-Based Spinach Sauce. LWT-Food Sci. Technol. 2015, 60, 86–94. [Google Scholar] [CrossRef]
  169. Zhang, Z.; Yang, Y.; Zhou, P.; Zhang, X.; Wang, J. Effects of High Pressure Modification Conformation and Gelation Properties of Myofibrillar Protein. Food Chem. 2017, 217, 678–686. [Google Scholar] [CrossRef]
  170. Chan, J.T.Y.; Omana, D.A.; Betti, M. Effect of Ultimate PH and Freezing on the Biochemical Properties of Proteins in Turkey Breast Meat. Food Chem. 2011, 127, 109–117. [Google Scholar] [CrossRef]
  171. Chauhan, O.P.; Kumar, S.; Nagraj, R.; Narasimhamurthy, R.; Raju, P.S. Effect of High Pressure Processing on Yield, Quality and Storage Stability of Peanut Paneer. Int. J. Food Sci. Technol. 2015, 50, 1515–1521. [Google Scholar] [CrossRef]
  172. Lowder, A.C.; Waite-Cusic, J.G.; Mireles Dewitt, C.A. High Pressure–Low Temperature Processing of Beef: Effects on Survival of Internalized E. Coli O157:H7 and Quality Characteristics. Innov. Food Sci. Emerg. Technol. 2014, 26, 18–25. [Google Scholar] [CrossRef]
  173. Gao, H.; Zeng, J.; Ma, H.; Wang, Z.; Pan, R. Improving Tenderness of Goose Breast by Ultra-High Pressure. Int. J. Food Prop. 2015, 18, 1693–1701. [Google Scholar] [CrossRef]
  174. Ma, H.; Ledward, D.A. High Pressure Processing of Fresh Meat—Is It Worth It? Meat Sci. 2013, 95, 897–903. [Google Scholar] [CrossRef]
  175. Zhang, H.; Tchabo, W.; Ma, Y. Quality of Extracts from Blueberry Pomace by High Hydrostatic Pressure, Ultrasonic, Microwave and Heating Extraction: A Comparison Study. Emir. J. Food Agric. 2017, 29, 815–819. [Google Scholar] [CrossRef] [Green Version]
  176. Volkov, A.Y.; Kruglikov, N.A.; Alexandrov, A.V.; Kotkova, V.V. Use of High Hydrostatic Pressure for Food Sterilization and Seed Treatment. IOP Conf. Ser. Mater. Sci. Eng. 2020, 1008, 012012. [Google Scholar] [CrossRef]
  177. Irna, C.; Jaswir, I.; Othman, R.; Jimat, D.N. Comparison between High-Pressure Processing and Chemical Extraction: Astaxanthin Yield From Six Species of Shrimp Carapace. J. Diet. Suppl. 2018, 15, 805–813. [Google Scholar] [CrossRef]
  178. Huppertz, T.; Hinz, K.; Zobrist, M.R.; Uniacke, T.; Kelly, A.L.; Fox, P.F. Effects of High Pressure Treatment on the Rennet Coagulation and Cheese-Making Properties of Heated Milk. Innov. Food Sci. Emerg. Technol. 2005, 6, 279–285. [Google Scholar] [CrossRef]
  179. Chemat, F.; Abert Vian, M.; Fabiano-Tixier, A.S.; Nutrizio, M.; Režek Jambrak, A.; Munekata, P.E.S.; Lorenzo, J.M.; Barba, F.J.; Binello, A.; Cravotto, G. A Review of Sustainable and Intensified Techniques for Extraction of Food and Natural Products. Green Chem. 2020, 22, 2325–2353. [Google Scholar] [CrossRef] [Green Version]
  180. Skripnuk, D.F.; Davydenko, V.A.; Romashkina, G.F.; Khuziakhmetov, R.R. Consumer Trust in Quality and Safety of Food Products in Western Siberia. Agronomy 2021, 11, 257. [Google Scholar] [CrossRef]
  181. De Oliveira, C.F.; Gurak, P.D.; Marczak, L.D.; Karwe, M. Extraction of Carotenoids from Passion Fruit Peel Assisted by High Pressure. X CIGR Sect. IV Int. Tech. Symp. 2016, 51, 2108–3111. [Google Scholar]
  182. George, J.M.; Sowbhagya, H.B.; Rastogi, N.K. Effect of High Pressure Pretreatment on Drying Kinetics and Oleoresin Extraction from Ginger. Dry. Technol. 2017, 36, 1107–1116. [Google Scholar] [CrossRef]
  183. Xi, J.; Yan, L. Optimization of Pressure-Enhanced Solid-Liquid Extraction of Flavonoids from Flos Sophorae and Evaluation of Their Antioxidant Activity. Sep. Purif. Technol. 2017, 175, 170–176. [Google Scholar] [CrossRef]
  184. 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]
  185. Cacace, F.; Bottani, E.; Rizzi, A.; Vignali, G. Evaluation of the Economic and Environmental Sustainability of High Pressure Processing of Foods. Innov. Food Sci. Emerg. Technol. 2020, 60, 102281. [Google Scholar] [CrossRef]
  186. Yordanov, D.G.; Angelova, G.V. High Pressure Processing for Foods Preserving. Biotechnol. Biotechnol. Equip. 2010, 24, 1940–1945. [Google Scholar] [CrossRef] [Green Version]
  187. Awasthi, A.K.; Cheela, V.R.S.; D’Adamo, I.; Iacovidou, E.; Islam, M.R.; Johnson, M.; Miller, T.R.; Parajuly, K.; Parchomenko, A.; Radhakrishan, L.; et al. Zero Waste Approach towards a Sustainable Waste Management. Resour. Environ. Sustain. 2021, 3, 100014. [Google Scholar] [CrossRef]
  188. Atuonwu, J.C.; Leadley, C.; Bosman, A.; Tassou, S.A. High-Pressure Processing, Microwave, Ohmic, and Conventional Thermal Pasteurization: Quality Aspects and Energy Economics. J. Food Process Eng. 2020, 43, e13328. [Google Scholar] [CrossRef]
  189. FAO. Food Loss and Food Waste. Food and Agriculture Organization of the United Nations. Available online: http://www.fao.org/food-loss-and-food-waste/flw-data (accessed on 13 July 2021).
  190. Ishangulyyev, R.; Kim, S.; Lee, S.H. Understanding Food Loss and Waste—Why Are We Losing and Wasting Food? Foods 2019, 8, 297. [Google Scholar] [CrossRef] [Green Version]
  191. Timmermans, A.J.M.; Ambuko, J.; Belik, W.; Huang, J. Food Losses and Waste in the Context of Sustainable Food Systems; CFS Committee on World Food Security HLPE: Rome, Italy, 2014. [Google Scholar]
  192. Shafiee-Jood, M.; Cai, X. Reducing Food Loss and Waste to Enhance Food Security and Environmental Sustainability. Environ. Sci. Technol. 2016, 50, 8432–8443. [Google Scholar] [CrossRef]
  193. Gómez-Estaca, J.; Calvo, M.M.; Sánchez-Faure, A.; Montero, P.; Gómez-Guillén, M.C. Development, Properties, and Stability of Antioxidant Shrimp Muscle Protein Films Incorporating Carotenoid-Containing Extracts from Food by-Products. LWT-Food Sci. Technol. 2015, 64, 189–196. [Google Scholar] [CrossRef]
  194. Zhao, W.; Yang, R.; Wang, M.; Lu, R. Effects of Pulsed Electric Fields on Bioactive Components, Colour and Flavour of Green Tea Infusions. Int. J. Food Sci. Technol. 2009, 44, 312–321. [Google Scholar] [CrossRef]
  195. Plazzotta, S.; Manzocco, L. Effect of Ultrasounds and High Pressure Homogenization the Extraction of Antioxidant Polyphenols from Lettuce Waste. Innov. Food Sci. Emerg. Technol. 2018, 50, 11–19. [Google Scholar] [CrossRef]
  196. Briones-Labarca, V.; Giovagnoli-Vicuña, C.; Cañas-Sarazúa, R. Optimization of Extraction Yield, Flavonoids and Lycopene from Tomato Pulp by High. Hydrostatic Pressure-Assisted Extraction; Elsevier Ltd.: Amsterdam, The Netherlands, 2019; Volume 278. [Google Scholar] [CrossRef]
  197. Torres-Ossandón, M.J.; Vega-Gálvez, A.; López, J.; Stucken, K.; Romero, J.; Di Scala, K. Effects of High Hydrostatic Pressure Processing and Supercritical Fluid Extraction Bioactive Compounds and Antioxidant Capacity of Cape Gooseberry Pulp (Physalis Peruviana L.). J. Supercrit. Fluids 2018, 138, 215–220. [Google Scholar] [CrossRef]
  198. Cascaes Teles, A.S.; Hidalgo Chávez, D.W.; Zarur Coelho, M.A.; Rosenthal, A.; Fortes Gottschalk, L.M.; Tonon, R.V. Combination of Enzyme-Assisted Extraction and High Hydrostatic Pressure for Phenolic Compounds Recovery from Grape Pomace. J. Food Eng. 2020, 288, 110128. [Google Scholar] [CrossRef]
  199. García-Parra, J.; González-Cebrino, F.; Delgado, J.; Cava, R.; Ramírez, R. High Pressure Assisted Thermal Processing of Pumpkin Purée: Effect on Microbial Counts, Color, Bioactive Compounds and Polyphenoloxidase Enzyme. Food Bioprod. Process. 2016, 98, 124–132. [Google Scholar] [CrossRef] [Green Version]
  200. García-Parra, J.; González-Cebrino, F.; Delgado-Adámez, J.; Cava, R.; Martín-Belloso, O.; Elez-Martínez, P.; Ramírez, R. Application of Innovative Technologies, Moderate-Intensity Pulsed Electric Fields and High-Pressure Thermal Treatment, to Preserve and/or Improve the Bioactive Compounds Content of Pumpkin. Innov. Food Sci. Emerg. Technol. 2018, 45, 53–61. [Google Scholar] [CrossRef] [Green Version]
  201. Gabrić, D.; Barba, F.; Roohinejad, S.; Gharibzahedi, S.M.T.; Radojčin, M.; Putnik, P.; Kovačević, D.B. Pulsed Electric Fields as an Alternative to Thermal Processing for Preservation of Nutritive and Physicochemical Properties of Beverages: A Review. J. Food Process Eng. 2018, 41, e12638. [Google Scholar] [CrossRef]
  202. Caili, F.U.; Huan, S.; Quanhong, L.I. A Review on Pharmacological Activities and Utilization Technologies of Pumpkin. Plant. Foods Hum. Nutr. 2006, 61, 73–80. [Google Scholar] [CrossRef]
  203. Garvey, M. Food Pollution: A Comprehensive Review of Chemical and Biological Sources of Food Contamination and Impact on Human Health. Nutrire 2019, 44, 1. [Google Scholar] [CrossRef]
  204. Dalsgaard, H.; Food, A.A.-E. Improving Energy Efficiency. In Environmentally-Friendly Food Processing; Woodhead Publishing Ltd.: Sawston, UK, 2003. [Google Scholar]
  205. Ayvaz, H.; Balasubramaniam, V.M.; Koutchma, T. High Pressure Effects on Packaging Materials. In High Pressure Processing of Food; Springer: New York, NY, USA, 2016; pp. 73–93. [Google Scholar] [CrossRef]
  206. Ayvaz, H.; Schirmer, S.; Parulekar, Y.; Balasubramaniam, V.M.; Somerville, J.A.; Daryaei, H. Influence of Selected Packaging Materials on Some Quality Aspects of Pressure-Assisted Thermally Processed Carrots during Storage. LWT-Food Sci. Technol. 2012, 46, 437–447. [Google Scholar] [CrossRef]
  207. Dhawan, S.; Barbosa-Cànovas, G.V.; Tang, J.; Sablani, S.S. Oxygen Barrier and Enthalpy of Melting of Multilayer EVOH Films after Pressure-Assisted Thermal Processing and during Storage. J. Appl. Polym. Sci. 2011, 122, 1538–1545. [Google Scholar] [CrossRef]
  208. Dhawan, S.; Varney, C.; Barbosa-Cánovas, G.V.; Tang, J.; Selim, F.; Sablani, S.S. Pressure-Assisted Thermal Sterilization Effects on Gas Barrier, Morphological, and Free Volume Properties of Multilayer EVOH Films. J. Food Eng. 2014, 128, 40–45. [Google Scholar] [CrossRef]
  209. Fleckenstein, B.S.; Sterr, J.; Langowski, H.-C. The Influence of High Pressure Treatment and Thermal Pasteurization the Surface of Polymeric Packaging Films. Packag. Technol. Sci. 2016, 29, 323–336. [Google Scholar] [CrossRef]
  210. Ahmed, J.; Mulla, M.; Arfat, Y.A. Application of High-Pressure Processing and Polylactide/Cinnamon Oil Packaging on Chicken Sample for Inactivation and Inhibition of Listeria Monocytogenes and Salmonella Typhimurium, and Post-Processing Film Properties. Food Control 2017, 78, 160–168. [Google Scholar] [CrossRef]
  211. Yoo, S.; Holloman, C.; Tomasko, D.; Koelling, K.; Pascall, M.A. Effect of High Pressure Processing on the Thermal and Mechanical Properties of Polyethylene Films Measured by Dynamical Mechanical and Tensile Analyses. Packag. Technol. Sci. 2014, 27, 169–178. [Google Scholar] [CrossRef]
  212. Gonçalves, A.A.; de Paiva Alves, J. High Pressure Technology Improves the Quality and Yield in the Seafood Industry. Available online: https://www.researchgate.net/publication/267509875_High_pressure_technology_improves_the_quality_and_yield_in_the_seafood_industry (accessed on 21 September 2021).
  213. Torres, J.A.; Velazquez, G. Commercial Opportunities and Research Challenges in the High Pressure Processing of Foods. J. Food Eng. 2005, 67, 95–112. [Google Scholar] [CrossRef]
  214. Abenoza, M.; Benito, M.; Saldaña, G.; Álvarez, I.; Raso, J.; Sánchez-Gimeno, A.C. Effects of Pulsed Electric Field on Yield Extraction and Quality of Olive Oil. Food Bioprocess Technol. 2013, 6, 1367–1373. [Google Scholar] [CrossRef]
  215. Arshad, R.N.; Buntat, Z.B.; Dastgheib, A.M.; Jusoh, Y.M.M.; Munir, A.; Aadil, R.M.; Ahmad, M.H. Continuous Flow Treatment Chamber for Liquid Food Processing through Pulsed Electric Field. J. Comput. Theor. Nanosci. 2020, 17, 1492–1498. [Google Scholar] [CrossRef]
  216. Ninčević Grassino, A.; Ostojić, J.; Miletić, V.; Djaković, S.; Bosiljkov, T.; Zorić, Z.; Ježek, D.; Rimac Brnčić, S.; Brnčić, M. Application of High Hydrostatic Pressure and Ultrasound-Assisted Extractions as a Novel Approach for Pectin and Polyphenols Recovery from Tomato Peel Waste. Innov. Food Sci. Emerg. Technol. 2020, 64, 102424. [Google Scholar] [CrossRef]
  217. Pinela, J.; Prieto, M.A.; Barros, L.; Carvalho, A.M.; Oliveira, M.B.P.P.; Saraiva, J.A.; Ferreira, I.C.F.R. Cold Extraction of Phenolic Compounds from Watercress by High Hydrostatic Pressure: Process Modelling and Optimization. Sep. Purif. Technol. 2018, 192, 501–512. [Google Scholar] [CrossRef] [Green Version]
  218. Duffuler, P.; Giarratano, M.; Naderi, N.; Suwal, S.; Marciniak, A.; Perreault, V.; Offret, C.; Brisson, G.; House, J.D.; Pouliot, Y.; et al. High Hydrostatic Pressure Induced Extraction and Selective Transfer of β-Phosvitin from the Egg Yolk Granule to Plasma Fractions. Food Chem. 2020, 321, 126696. [Google Scholar] [CrossRef]
  219. Huang, Y.L.; Tsai, Y.H. Extraction of Chitosan from Squid Pen Waste by High Hydrostatic Pressure: Effects on Physicochemical Properties and Antioxidant Activities of Chitosan. Int. J. Biol. Macromol. 2020, 160, 677–687. [Google Scholar] [CrossRef]
  220. Jun, X. Caffeine Extraction from Green Tea Leaves Assisted by High Pressure Processing. J. Food Eng. 2009, 94, 105–109. [Google Scholar] [CrossRef]
  221. Strati, I.F.; Gogou, E.; Oreopoulou, V. Enzyme and High Pressure Assisted Extraction of Carotenoids from Tomato Waste. Food Bioprod. Process. 2015, 94, 668–674. [Google Scholar] [CrossRef]
  222. Naderi, N.; Pouliot, Y.; House, J.D.; Doyen, A. High Hydrostatic Pressure Effect in Extraction of 5-Methyltetrahydrofolate (5-MTHF) from Egg Yolk and Granule Fractions. Innov. Food Sci. Emerg. Technol. 2017, 43, 191–200. [Google Scholar] [CrossRef]
  223. Pagan, R.J.; Prasad, P. Eco-Efficiency, Water Conservation and Food Processing in Australia—UQ eSpace. 2005. Available online: https://espace.library.uq.edu.au/view/UQ:102827 (accessed on 13 July 2021).
  224. Atuonwu, J.C.; Tassou, S.A. Energy Issues in Microwave Food Processing: A Review of Developments and the Enabling Potentials of Solid-State Power Delivery. Crit. Rev. Food Sci. Nutr. 2018, 59, 1392–1407. [Google Scholar] [CrossRef] [PubMed]
  225. Bull, M.K.; Zerdin, K.; Howe, E.; Goicoechea, D.; Paramanandhan, P.; Stockman, R.; Sellahewa, J.; Szabo, E.A.; Johnson, R.L.; Stewart, C.M. The Effect of High Pressure Processing on the Microbial, Physical and Chemical Properties of Valencia and Navel Orange Juice. Innov. Food Sci. Emerg. Technol. 2004, 5, 135–149. [Google Scholar] [CrossRef]
  226. Landl, A.; Abadias, M.; Sárraga, C.; Viñas, I.; Picouet, P.A. Effect of High Pressure Processing on the Quality of Acidified Granny Smith Apple Purée Product. Innov. Food Sci. Emerg. Technol. 2010, 11, 557–564. [Google Scholar] [CrossRef]
  227. Kowalczyk, W.; Hartmann, C.; Delgado, A. Freezing and Thawing at the High Hydrostatic Pressure Conditions—Modelling and Numerical Simulation. PAMM 2003, 3, 388–389. [Google Scholar] [CrossRef]
  228. Pham, Q.T. Advances in Food Freezing/Thawing/Freeze Concentration Modelling and Techniques. Jpn. J. Food Eng. 2008, 9, 21–32. [Google Scholar] [CrossRef]
  229. Sampedro, F.; McAloon, A.; Yee, W.; Fan, X.; Zhang, H.Q.; Geveke, D.J. Cost Analysis of Commercial Pasteurization of Orange Juice by Pulsed Electric Fields. Innov. Food Sci. Emerg. Technol. 2013, 17, 72–78. [Google Scholar] [CrossRef]
  230. Sampedro, F.; McAloon, A.; Yee, W.; Fan, X.; Geveke, D.J. Cost Analysis and Environmental Impact of Pulsed Electric Fields and High Pressure Processing in Comparison with Thermal Pasteurization. Food Bioprocess Technol. 2014, 7, 1928–1937. [Google Scholar] [CrossRef]
  231. Huang, Y.; Gan, Y.; Li, F.; Yan, C.; Li, H.; Feng, Q. Effects of High Pressure in Combination with Thermal Treatment on Lipid Hydrolysis and Oxidation in Pork. LWT-Food Sci. Technol. 2015, 63, 136–143. [Google Scholar] [CrossRef]
  232. Rastogi, N.K.; Raghavarao, K.S.M.S.; Balasubramaniam, V.M.; Niranjan, K.; Knorr, D. Opportunities and Challenges in High Pressure Processing of Foods. Crit. Rev. Food Sci. Nutr. 2007, 47, 69–112. [Google Scholar] [CrossRef] [PubMed]
  233. Ramaswamy, R.; Balasubramaniam, V.; Kaletun, G. High Pressure Processing: Fact Sheet for Food Processors. Available online: https://blog.ideasinfood.com/files/hppfactsheet.pdf (accessed on 8 August 2021).
Figure 1. A standard batch HPP system.
Figure 1. A standard batch HPP system.
Sustainability 13 13908 g001
Figure 2. Presentation of temperature, pressure, and time in HPP.
Figure 2. Presentation of temperature, pressure, and time in HPP.
Sustainability 13 13908 g002
Figure 3. Contributions of HPP to sustainable food processing.
Figure 3. Contributions of HPP to sustainable food processing.
Sustainability 13 13908 g003
Table 1. Applications of HPP in the food industry.
Table 1. Applications of HPP in the food industry.
ApplicationsTreatment ConditionsFood Sample Treatment EffectsReferences
Meat processing175–600 MPa, 3–5 min,Meat, Meat Products, and Seafood Showed minimal effect on nutrients and sensorial characters. Slowed down microbial growth and reduced the activity rate of spoilage bacteria[29,30,31]
Microbial reduction300–600 MPa, 5–10 minMeat, juices, milk products Caused significant reductions in microbes in food items, i.e., about a 1.6–5-log reduction[32,33,34]
Extraction250–500 MPa, 5–15 minSeeds, fruits, vegetables, plants, cerealscarbohydrates, bioactive compoundsExtraction yields were enhanced by HPP compared with the conventional method[35,36,37,38,39]
Pretreatment200–300 MPa, 2–6 minMeat, fruits, vegetables HPP as a pretreatment improved textural, nutritional, and sensorial attributes[40,41,42,43]
Seed treatment200–400 MPa, 10–60 min
20–60 °C
Moringa oleifera kernels, Basil, chia seeds HPP enhanced the extraction of oil as well as the structure of seeds[44,45,46]
MPa: megapascal, CFU: colony-forming unit.
Table 2. Target pressure for different food processing applications in HPP.
Table 2. Target pressure for different food processing applications in HPP.
ApplicationsTarget PressureFood SamplesReferences
Pasteurization(300–600 MPa)Meat, Milk, fruit juices, cereals[61,62]
Sterilization(500–900 MPa)Fruits, macaroni, cheese[63]
Extraction(250–500 MPa)Seeds, fruits[35]
Pretreatment(200–600 MPa)Meat, vegetables[40]
Table 3. HPP Contributions to sustainable food processing.
Table 3. HPP Contributions to sustainable food processing.
ContributionsApplicationTreatment ConditionsContributions Toward SustainabilityReferences
Food safetyNutritional Improvement200–600 MPa, 5–15 minImproved nutritional value by HPP[69]
Food preservation200–400 MPa, ˃2 minMore efficiently used for pasteurization of liquid foods and dehydration of solid foods[70]
Reduction of food contaminants30 to 500 MPa, 30–50 °CHelped to decrease food contaminants and toxins[59]
Solid food pretreatment200 MPa, 0–6 minMaintained texture and color[48]
Environmental sustainabilityReduced food waste300 MPa, 3 minContribution to reducing food waste[71]
Food packagingEffects on food packaging600 a, 60 min; 800 MPa, 10 min, 60 °CMinimal effect on packaging material[72]
Economic sustainabilityProfit attained356 and 600 MPa, 30 minContributed to the recovery of valuable components from food[73]
Water efficiencyMinimal water consumption Not only energy saving but also water-saving technology[74]
Energy efficiencyMinimal energy consumption300–700 MPaUtilised less energy than conventional techniques[75]
Food securityEnhanced yield200 and 500 MPaHPP improved the food yield[76]
Table 4. Treatment of liquid foods by HPP.
Table 4. Treatment of liquid foods by HPP.
Food SampleTarget ProductTreatment ConditionsLog ReductionReferences
Coconut waterClostridium botulinum type E and Clostridium sporogenes550 MPa, 3 min, 10 °C3.0 ± 0.1[95]
Açaí juiceBacteria600 MPa, 3 min5–6[96]
Whole milk and skim milkCronobacter sakazakii300 MPa, 400 MPa, 5.0 minNo detection[97]
Cucumber juice Yeast and mold, total aerobic bacteria500 MPa 5 min1.35–1.94[98]
Beetroot and carrot juiceE. coli300–500 MPaNo detection[99]
Raw milkPathogens600 MPa, 3 min5[100]
Acidic fruit juicesAlicyclobacillus acidoterrestris600 MPa, 80 °C, 3 min2.0–2.5[101]
WinesBrettanomyces bruxellensis200 MPa, up to 3 min,1.1–5.1[102]
Table 5. HPP effects on food nutrients.
Table 5. HPP effects on food nutrients.
Food NutrientsFood SampleTreatment ConditionsTreatment EffectsReferences
Fatty acidCaprine milk200–500 MPaNo marked differences in the fatty acid profile, excluding an increase in branched-chain fatty acids[151]
Free amino acidsSerrano dry-cured ham600 MPa, 6 min, 4 °CNo effect of HPP on total FAA levels[152]
ProteinMilk150–450 MPa, 15 minThe disintegration of casein micelles at ˃250 MPa; serum proteins were denatured[153]
Free amino acidsChinese rice wine400 MPa or 600 MPa 10 minHPP increased the free amino acid content[154]
CarotenoidsCloudy carrot juice300, 450, 600 MPa, 5 min, ≈22 °CThe highest carotenoid degradation (41%) occurred at 350 MPa, whereas the lowest (26%) occurred at 600 MPa[155]
PolyphenolsKiwi berry450, 550 or 650 MPa, 5 or 15 min Caused a significant increase in the individual polyphenol content[156]
Anthocyanins, non-anthocyanin phenolics, tocopherolsAcai juice400, 450, 500 and 600 MPa, 5 min, 20 °CEfficient preservation technique for anthocyanins compared to thermal pasteurization (up to 40%) [157]
Polyphenols, EnzymesCarrot juice450 and 600 MPa, 5 min, ≈ 22 °CAt 300 MPa, maximum inactivation of (PPO) enzymes (57%) was achieved, and at 600 MPa, about 31% inactivation of peroxidase (POD) enzymes was observed. Significant changes in the color parameters and the browning index were observed[77]
Anthocyanin, polyphenol oxidase, and β-glucosidaseBlueberry300 MPaHPP treatment resulted in better puree colour retention than the conventional treatment[158]
Total phenolic compounds, vitamin CStrawberry100 and 500 MPa, 10 minAt 400 MPa, the maximum total phenolic content was obtained; preserved vitamin C content in strawberry[49]
Oxidoreductive enzymesMushroom200–900 MPa, 5–45 °C, 1–15 minHPP used in the experiments significantly (p < 0.05) decreased PPO and POD activities, with a greater decrease in the relative activity (RA) of the enzymes observed when the pressure was increased[159]
Carbohydrate hydrolysing enzymesLemon flavedo400 MPa, 10 minHPP-treated samples had high levels of carbohydrates Hydrolyzing enzyme inhibition[160]
Polyphenol oxidasePawpaw600 MPa, 4 °C, 76 sHPP significantly decreased PPO activity. HPP was proved to be an effective technique for the longer shelf-life of fresh-packaged pawpaw pulp[161]
Table 6. HPP impact on food yield.
Table 6. HPP impact on food yield.
ComponentsSourcesTreatment ConditionsYieldReferences
Polyphenols and anthocyaninsBlueberry pomace500 MPaA marked increase in the yield of polyphenols and anthocyanins was observed, about 70% and 40%, using HPP compared with the conventional heating method (53% and 32%)[175]
Tomatoes Below 100 MPaA marked increase in the yield of tomatoes occurred, i.e., >60%[176]
AstaxanthinShrimp
Carapace
210 MPaIncreased the yield from 29.44% to 59.97% by HPP[177]
CheeseMilk250–600 MPaA 15% increase in the yield of cheese was obtained[178]
M. oleifera oilM. oleifera (MO) kernels50–250 MPaFree oil (73% w/w) was recovered from ground-sieved kernels by using HPP compared with AEE alone[44]
Table 7. Extraction of different valuable compounds from food waste.
Table 7. Extraction of different valuable compounds from food waste.
Food WasteTreatment ConditionsRecovered CompoundsReferences
Orange peel300 and 500 MPaPolyphenols[71]
Tomato pulp 450 MPaLycopene, Flavonoids[196]
Cape gooseberry pulp300–500 MPaBioactive compounds[197]
Grape pomace200 MPaPhenolic compounds[198]
Pumpkin puree600 MPa, 70 °CBioactive compounds[199]
Table 8. Extraction of valuable compounds using HPP.
Table 8. Extraction of valuable compounds using HPP.
ComponentsSourcesTreatment ConditionsProductionReferences
Pectin and polyphenolsTomato peel waste30 and 45 minEnhanced pectin recovery by about 15% compared with conventional extraction[216]
Phenolic compoundsWatercress3.1 min, 600 MPa 64.68 ± 2.97 mg/g yield recovered[217]
PhosvitinEgg yolk400 and 600 MPa, 5 and 10 minPhosvitin recovery was maximum at 600 MPa[218]
ChitosanSquid pen waste500 MPa, 10 minMaximum yield of 81.9%[219]
XyloglucanTamarind seed250–500 MPa Yields were about 51.6–53.0% higher than the conventional yield (46.4%)[35]
CaffeineGreen tea leaves500 MPa, 1 minMaximum yields: 4.0 ± 0.22%[220]
LycopeneTomato waste700 MPa, 30 minMaximum yield: 89.4 mg/kg[221]
5-methyltetrahydrofolateEgg yolk400 MPa, 5 minMaximum recovery: 93%[222]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nabi, B.G.; Mukhtar, K.; Arshad, R.N.; Radicetti, E.; Tedeschi, P.; Shahbaz, M.U.; Walayat, N.; Nawaz, A.; Inam-Ur-Raheem, M.; Aadil, R.M. High-Pressure Processing for Sustainable Food Supply. Sustainability 2021, 13, 13908. https://doi.org/10.3390/su132413908

AMA Style

Nabi BG, Mukhtar K, Arshad RN, Radicetti E, Tedeschi P, Shahbaz MU, Walayat N, Nawaz A, Inam-Ur-Raheem M, Aadil RM. High-Pressure Processing for Sustainable Food Supply. Sustainability. 2021; 13(24):13908. https://doi.org/10.3390/su132413908

Chicago/Turabian Style

Nabi, Brera Ghulam, Kinza Mukhtar, Rai Naveed Arshad, Emanuele Radicetti, Paola Tedeschi, Muhammad Umar Shahbaz, Noman Walayat, Asad Nawaz, Muhammad Inam-Ur-Raheem, and Rana Muhammad Aadil. 2021. "High-Pressure Processing for Sustainable Food Supply" Sustainability 13, no. 24: 13908. https://doi.org/10.3390/su132413908

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