Next Article in Journal
Effects of Soaking on the Volatile Compounds, Textural Property, Phytochemical Contents, and Antioxidant Capacity of Brown Rice
Next Article in Special Issue
Multi-Response Optimization of Pyrrolizidine Alkaloids Removal from Chrysanthemum morifolium by High-Pressure Extraction
Previous Article in Journal
Bioaccessibility and Antioxidant Activity of Polyphenols from Pigmented Barley and Wheat
Previous Article in Special Issue
Frozen-Phase High-Pressure Destruction Kinetics of Escherichia coli as Influenced by Application Mode, Substrate, and Enrichment Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in the Synergistic Bactericidal Effect of High Pressure and Temperature Processing in Fruits and Vegetables and Related Kinetics

1
College of Biosystems Engineering and Food Science, Zhejiang University, 866 Yuhangtang Road, Hangzhou 310058, China
2
Key Laboratory of Equipment and Informatization in Environment Controlled Agriculture, Ministry of Agriculture, 866 Yuhangtang Road, Hangzhou 310058, China
3
Hangzhou Jiangnan Talent Service Co., Ltd., 681 Qingchun East Road, Hangzhou 310000, China
4
Department of Food Science and Agricultural Chemistry, McGill University, 21111 Lakeshore Road, St-Anne-de-Bellevue, QC H9X 3V9, Canada
*
Author to whom correspondence should be addressed.
Foods 2022, 11(22), 3698; https://doi.org/10.3390/foods11223698
Submission received: 17 October 2022 / Revised: 13 November 2022 / Accepted: 16 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Ultra-High-Pressure Processing of Fruit and Vegetable Products)

Abstract

:
Background: Traditional thermal processing is a widely used method to ensure food safety. However, thermal processing leads to a significant decline in food quality, especially in the case of fruits and vegetables. To overcome this drawback, researchers are extensively exploring alternative non-thermal High-Pressure Processing (HPP) technology to ensure microbial safety and retaining the sensory and nutritional quality of food. However, HPP is unable to inactivate the spores of some pathogenic bacteria; thus, HPP in conjunction with moderate- and low-temperature is employed for inactivating the spores of harmful microorganisms. Scope and approach: In this paper, the inactivation effect of high-pressure and high-pressure thermal processing (HPTP) on harmful microorganisms in different food systems, along with the bactericidal kinetics model followed by HPP in certain food samples, have been reviewed. In addition, the effects of different factors such as microorganism species and growth stage, process parameters and pressurization mode, and food composition on microbial inactivation under the combined high-pressure and moderate/low-temperature treatment were discussed. Key findings and conclusions: The establishment of a reliable bactericidal kinetic model and accurate prediction of microbial inactivation will be helpful for industrial design, development, and optimization of safe HPP and HPTP treatment conditions.

1. Introduction

Food safety has become a global common concern as consumers around the world are facing different degrees of food safety risks. As per the reports of the World Health Organization (WHO), approximately 600 million people (almost 1 in 10 people in the world) fall ill after consuming contaminated food, and approximately 420,000 die every year due to foodborne diseases [1]. Most foodborne diseases are more likely to be widespread and, even global due to changes in food production, supply, and widespread food distribution. New foodborne pathogens continue to be found as a result of altered food production conditions and improved laboratory detection methods. In particular, a significant increase in antimicrobial drug-resistant bacteria and several viruses was observed, which were not previously recognized [2]. The food industry all over the world is also facing huge economic losses due to the contamination of raw materials. Therefore, food safety remains a huge global challenge as foodborne diseases obstruct socioeconomic progress by straining healthcare systems and harming national economies, international trade, and tourism. Therefore, sterilization has become the utmost important step in food processing industries. It was also observed that the detection of food pathogens, food contaminants, and toxins has also received significant attention since the last decade to ensure food safety [3].
Conventionally, several chemical and physical methods have been used for the decontamination of food material. However, these methods are associated with several drawbacks such as chemical residue in resultant food products and poor sensory and nutritional quality of food. Thus, these conventional decontamination methods were not able to meet consumers’ increasing demand for high-quality food. At present, thermal sterilization is the mainstream sterilization method in the food processing industry. Although thermal processing ensures product safety, it causes undesirable nutritional and organoleptic quality in food [4,5,6]. Alicyclobacillus acidoterrestris is a type of bacillus that can grow in pasteurized beverages. It has been mentioned that the use of high temperatures during heat treatment can inactivate A. acidoterrestris inoculated in apple juice, but also deteriorate the nutritional and sensory qualities of apple juice [7]. Among various food products, fruit and vegetable products, the majority of which contain heat-sensitive ingredients, exhibit significant degradation in their nutritional value, appearance, taste, and flavor following heat treatment [8,9]. Some fruits, such as carambola [10] and muskmelon [11], contain volatile sulfur compounds, which are responsible for their unique flavor. It was also mentioned that degradation of these compounds following heat treatment resulted in off-flavor and the low market price of these food products. That, in turn, limits the commercial production of these fruit and resultant products. In addition, the thermal processing of food materials also resulted in a significant decrease in their antioxidant capacity [10,12,13]. Therefore, it is urgent to adopt an alternative method that can not only enhance the microbial safety of food but also maximize the retention of the original physical and biochemical properties of food.
High-pressure processing (HPP) sterilization is one of the promising techniques among recent technologies that employs low temperature and high pressure, which can kill the vast majority of harmful microorganisms in the food in just a few minutes [14]. HPP is a non-thermal food processing technology that was reported to maintain the original flavor, physical parameters, and chemical properties of heat-sensitive fruit pulp to their maximum extent [15,16]. In comparison with traditional thermal processing, HPP presents better retention levels of the bioactive compounds, increased microbial safety and reduced enzyme activity [10,17]. Thus, HPP technology has been regarded as a green alternative to traditional preservation technologies. In the past 20 years, HPP technology has been industrialized in the field of food processing in some countries and regions to meet consumer demand for mild processed food without preservatives and with high nutritional quality.
HPP is governed mainly by Le Chatelier’s principle, the microscopic order principle, and the Isostatic principle. Le Chatelier’s principle states that any process in equilibrium that is accompanied by a decrease in volume can be enhanced by pressure. A change in micro-order also occurs during HPP treatment, which means that the molecules will shift to a more compact structure. The Isostatic principle states that the applied pressure is instantaneously and evenly distributed within the food, independent of the structure and geometry of the food [18,19]. The mechanisms of microbial inactivation by HPP are based on a combination of changes in the cell membranes. Structural changes in protein and membrane phospholipids can alter membrane permeability and cellular functions [20].
Along with potent advantages, HPP technology also exhibits certain limitations. In general, spoilage bacteria and pathogenic microorganisms are inactivated in the pressure range of 400 to 600 MPa, but certain microorganisms exhibit strong pressure tolerance, especially Gram-positive bacteria. In addition, pressure treatments up to 1000 MPa may not completely inactivate bacterial spores at ambient temperatures [21,22]. It is important to note that the pasteurization of foods requires at least 5–6 log cfu/g or ml reduction of key pathogenic or spoilage microorganisms. Traditional thermal processing of food products can efficiently achieve the required reduction of pathogenic or spoilage microbes. However, it also diminishes the quality of fruits and vegetables [23]. Thus, a combination of HPP and moderate temperature treatment is required for the efficient inactivation of spores while retaining the physical and nutritional quality of food.
Different fruit and vegetable products have different physical and biochemical characteristics, and during decontamination the different parameters of different processing methods have to be optimized to achieve optimum product quality, microbial safety, and other aspects. Therefore, much basic research is required to provide a theoretical basis for the practical application of a particular processing method. Through bactericidal kinetics, the lethal characteristics of microorganisms can be described, which plays an important role in predicting the number of microorganisms and ensuring the microbial safety of fruit and vegetable products [24]. These kinetic models can also be used to develop Hazard Analysis and Critical Control Point (HACCP) plans and process validation studies.
A reliable mathematical model and an accurate prediction of microbial inactivation will be helpful for the industrial design, development, and optimization of safe decontamination treatment, which in turn reduces the number of experiments [25,26]. Therefore, significant efforts are required to obtain different kinetic model parameters for various target bacteria in order to develop a database. Thus, this paper is an initiative to review the research progress related to the bactericidal kinetics of HPP and temperature–pressure synergy in fruit and vegetable products for microbial inactivation and food safety enhancement.

2. Inactivation Kinetics of Microbes by Temperature in Conjunction with High-Pressure

2.1. Primary Models

Primary models in the bactericidal kinetics of HPTP are mathematical equations that describe the changes in microbial counts induced by pressure as a function of treatment times. The first-order kinetic model is one of the common primary models employed to describe microbial log survivors in foods followed by the combination of high pressure and temperature pressure treatment [23,27]. The first-order dynamic model is explained as follows:
L o g N N 0 = k t = t D T , P
where N0 is the initial number of viable cells in the control sample (cfu/g or cfu/mL), N is the number of survivors in samples after HPP or HPTP treatment, t is the processing time, k is the inactivation rate constant of microbial number (related to the environmental conditions), and DT,P is the time required for one log reduction of the microbial population at a certain temperature and/or pressure.
The HPP and high-pressure thermal processing (HPTP) showed a linear trend for microbial inactivation in some fruit and vegetable systems, as mentioned in Table 1. However, in some cases, the inactivation curve of microorganisms may present a delayed or tailing phenomenon as the pressure treatment may cause sublethal injuries in the microorganisms and may activate dormant spores [28,29]. The most accepted hypothesis used to explain the tailing effect is “the presence of subpopulations within a microbial population that are more resistant to pressure treatments and remain viable even after prolonged pressure holding times” [30]. In addition, the non-linear behavior during pressure treatments is attributed to the cumulative damage to microbial cells [31]. In this case, the linear model of first-order kinetics can no longer fit the situation, and a nonlinear model is required to explain the delayed or tailing effect. Similar to the first-order kinetic model, the Weibull model is also a common primary model, and it is a typical representation of non-linear models. The Weibull model is not only applicable to describe microbial thermal inactivation [27] but can also be employed to explain other sterilization methods, such as pulsed electric field [32] and HPP [33]. Some studies have also mentioned that the Weibull model was more suitable to describe the microbial inactivation caused by the HPTP compared with thermal inactivation (Table 1). Hossein et al. reported that Weibull distribution was the best mathematical model to describe the inactivation of Bacillus coagulans 185A spores in tomato juice by HPTP, whereas first-order kinetics was appropriate for explaining only thermal processing [34].
The Weibull model was initially used for the determination of failure time in reliability engineering and was applied to the survival curve of microorganisms, which was the combination of the accelerated failure time model and parameter distribution [38,39]. The Weibull model (Equation (2)) was based on the principle of heterogeneity in the resistance distributed among individual cells within a population [40].
l o g N N 0 = b t n
where b is a rate parameter that is related to the rate at which the microorganism is inactivated, and n describes the degree of curvilinearity. When b < 1, the inactivation curve corresponds to concave-upwards (tailings). When b > 1, the inactivation curve corresponds to concave-downwards (shoulders). When b = 1, the model becomes a straight line, which is the first-order kinetic model [41]. This model assumes that the probability of death of a single cell or spore after treatment is dispersed according to Weibull distribution, and the survival curve of the distribution of lethal events is exponentially cumulative. The Weibull model also assumes that the microbes exhibit different resistance against treatment and these differences are permanent [42].

2.2. Secondary Models

The first-order kinetic models are established under certain temperature and pressure conditions. Unlike first-order kinetic models, the secondary models are an extension of the primary models, in which the parameters of the primary models relate to the environmental variables/conditions such as pressure or temperature [23]. Secondary models are applied to predict changes in the kinetics parameters of primary inactivation models as functions of intrinsic or extrinsic factors [43]. ZT-value (shown in Equation (3)) represents the temperature required to decrease D by one log cycle under a certain pressure, which can also reflect the sensitivity of microorganisms to temperature. This is equal to the reciprocal of the slope of the log D-values plotted against temperature [44,45]. The smaller the ZT-value, the higher the temperature sensitivity will be.
Z T = T 1 T l o g D l o g D 1
where D1 is the D-value at a reference temperature T1 (°C) and T is the temperature of the isothermal treatment (°C). Likewise, the ZP-value (shown in Equation (4)) is the pressure required to decrease the D value by one log cycle at a certain temperature and is equal to the reciprocal of the slope of the log D-values plotted against pressure. ZP-value can also reflect the sensitivity of microorganisms to pressure. A smaller ZP-value indicates greater sensitivity to pressure.
Z P = P 1 P l o g D l o g D 1
The previous studies related to high-pressure and temperature synergistic bactericidal kinetics also reflect the importance of sample temperature detection and control technology under a high-pressure environment. If the sample temperature under a high-pressure environment is not accurately detected and efficiently controlled, then the model developed to predict bacterial kinetics will be inaccurate [22,46]. For example, Lori et al. used a first-order kinetic model to accurately describe the thermal and pressure inactivation mechanics of Campylobacter coli and C. jejuni [47]. At the same time, secondary models were established to accurately predict the changes in the number of C. coli and C. jejuni under the combined effects of pressure (0.1–500 MPa), temperature (10 to 65 °C), and treatment time. In addition, the Weibull model can also relate the inactivation rate parameter (b-value) to environmental variables, particularly temperature, and can be used to predict the parameter values outside the range of the variables tested [23].

2.3. Polynomial Models

Polynomial models, also known as Response Surface Methodology (RSM), are used to analyze the effect of individual factors on inactivation parameters or their interactions [23,43]. RSM can provide an optimal fitting of polynomial models from a minimal number of experiments and enable the interaction study between factors based on the response of interest, including nonlinearities on curves [26,43]. As shown in Equation (5), described by Evelyn and Filipa [23]:
Y = B 0 + i = 1 n B i X i + i = 1 n B i i X i 2 + j 1 n B i j   X i X j + ε
where Y is the predicted response; B0 is a constant; Bi, Bii, and Bij are model coefficients; Xi and Xj are the input variables (environmental factors); and ε is the error term. By graphically translating RSM, operators can find the operating conditions that reduce the response, Y.

3. Influence of Different Factors on Microbial Inactivation during HPP or HPTP

Many factors affect the inactivation of microorganisms during HPP or HPTP. These factors include microbial species and growth stage, process parameters and pressurization method, temperature under pressure, food composition, and other factors such as food additives, water activity, and pH value [48,49]. in addition, researchers have also explained the influence of process parameters and pressure mode on microbial inactivation, and their mechanism of action has also been widely recognized. The research on high-pressure synergistic sterilization at moderate temperatures is also increasing. However, only a few studies have reported the synergistic sterilization of food products by low temperature and high pressure. It is important to note that treatment temperature exhibits a significant effect on the development of the kinetic model for sterilization [50].

3.1. Microbial Species and Growth Stages

Different microbe species have different sensitivities to high pressure, even within the same food system. In addition, microorganisms, including pathogens, can present significantly different responses toward high pressure. This variation exists not only between different species but also between strains of the same species [48]. Table 2 shows the inactivation effect of high pressure on different kinds of microorganisms in the same food systems. The pressure tolerance order of general microorganisms is highest in spores > Gram-positive (Listeria monocytogenes and Staphylococcus aureus) bacteria > fungus > Gram-negative (Escherichia coli) bacteria [48,51,52]. However, this order can be changed based on the food system. This is mainly because some macromolecules in food can play a protective role on some microorganisms. In addition, the resistance of microorganisms to high pressure is varied based on their growth stages. The pressure tolerance of microbial cells in their stable growth phase is higher compared to the cells in the logarithmic phase [53,54,55,56,57,58,59].
Until now, inactivated bacterial spores are the biggest challenge faced by HPP during food sterilization. Bacterial spores are highly resistant compared to mold and yeast spores. These bacterial spores are likely to grow as toxin-producing cells, leading to foodborne illness and disease outbreaks. Therefore, they are often used as indicators of pasteurization and sterilization processes in foods. Conventional pressures (<600 MPa) in the food industry have difficulty inactivating the majority of spores at room temperature. Even under the extreme pressure generated by laboratory equipment, some spores can survive. For example, B. subtilis spores were found to survive under processing conditions up to 1200 MPa at ambient temperature [57]. In addition, spore resistance varies significantly between species and is significantly influenced by the food substrate. In general, among all microorganisms, spores exhibit the highest resistance to HPP, thus presenting an urgent need for HPTP for their inactivation in foods [23].

3.2. Process Parameters and Pressure Mode

The influence of pressure and holding time of HPP or HPTP on the bactericidal effect has been studied extensively [48]. Generally, a higher pressure and a longer holding time correspond to a better sterilization effect [48,58,59,62]. Margaret et al. [63] studied the effect of HPP (at 20 °C for 1 min) in phosphate buffer on Leuconostoc and found a significant decrease in the number of Leu. kimchii in the phosphate buffer with an increase in the pressure level. Basak et al. [64] also mentioned that under the same level of pressure, the mortality of Leu. mesenteroides in orange juice increased with the increase of pressure hold time. However, the lethality of some microbial spores under low pressure may be higher compared to the mortality under high pressure as higher pressure may associate with spore germination [23,65].
The variation in pressure applied during HPP also exhibits a significant effect on microbial inactivation. The most commonly employed method is a single static high pressure (pressure holding time > 0 s). The whole high-pressure treatment process includes three parts: boost pressure, keep pressure, and relieve pressure. The boost and relief process without the keep pressure process during high-pressure treatment is called pulse-type high-pressure treatment. Different types of high-pressure pasteurization methods are composed of these two above-mentioned methods [66]. It has been reported that multi-stage high pressure (with or without pressure holding time) treatment can improve the microbial inactivation rate [67,68,69]. The multi-pulsed HPP is mentioned to be more effective than classical or single-pulsed HPP for the inactivation of enzymes, yeast cells, bacterial cells, and fungal and bacterial spores [70]. Aleman et al. have also found that multi-pulse HPP treatment was more effective compared to the single-pulse HPP treatment for inactivating Saccharomyces cerevisiae in pineapple juice at the same holding time [71]. Donsì et al. also indicated that the effectiveness of multiple pulses is dependent on the combination of pulse number, pressure, and temperature [62].
Chapleau et al. found a linear reduction in microbial count (Salmonella typhimurium and L. monocytogenes) with high-pressure pulses as a product of pressure and time [72]. Meanwhile, a logarithmic reduction in the microbial count was observed with increasing holding time. In addition, the pressurization and depressurization rates also have an impact on the sterilization effect [72,73]. Furthermore, Ratphitagsanti et al. have also reported the potent bactericidal effect of a low rate of pressurization during HPP compared to the high-rate of pressurization of samples containing B. amyloliquefaciens spores [73]. This study also found that double-pulse treatment presents better sterilization compared to single-pulse treatment.

3.3. Temperature

Temperature is the most important external condition for microbial growth and metabolism, and it has a significant influence on microbial survival. It is well known that the ambient temperature around the sample during pressurization will affect microbial resistance. It is also mentioned that moderate and low temperatures will increase the inactivation rate of microorganisms compared to room temperature [48]. However, there are significant differences in the sterilization mechanism of high-pressure and moderate-temperature synergistic sterilization and high-pressure and low-temperature synergistic sterilization. In the following sections, studies related to both sterilization methods are discussed in detail.

3.3.1. High-Pressure and Moderate Temperature Synergistic Sterilization in Fruits and Vegetables

The sterilization effect of high-pressure treatment will be strengthened with the increase in temperature, especially above room temperature (25 °C). It has been mentioned that the treatment at a moderate temperature (40–90 °C) increases the degree of protein denaturation; thus, an increase in the temperature during HPP treatment significantly enhances the lethality of microorganisms [74]. On the other hand, high pressure can lead to irreversible alteration in microbial cellular structure. The high pressure causes damage to both the membrane and the cell wall, which increases cell permeability and leads to an interruption in cell metabolism [49,75]. In addition, the loss of the secondary, tertiary, or quaternary structure of large molecules and the modification of complex organized structures are observed, followed by high-pressure treatment, which, in turn, leads to microbial death [76]. Furthermore, Zhang et al. have also mentioned a decrease in the thermal stability of horseradish peroxidase with an increase in the pressure [14]. High-pressure treatment alters the structure of the protein, which in turn changes the stability of the protein under pressure. In addition, moderate heating can enhance microbial inactivation under pressure, which in some cases leads to achieve desired results at lower pressure [48,77].
The spores of pathogenic microorganisms in food are extremely pressure-resistant. Thus, pressure–temperature co-treatment is considered to be one of the most effective and feasible methods for the inactivation of different pathogenic spores [46,78,79]. Table 3 shows the inactivation effect of HPP and HPTP on certain microbial spores in different fruit and vegetable systems. Previous studies have also mentioned that pressure and moderate temperature exhibit a potent synergistic effect on the inactivation of microorganisms and enzymes [68]. Evelyn et al. [80] exposed apple juice inoculated with Neosartorya fischeri JCM 1740 spores to high-pressure (600 MPa, room temperature), and the control group was thermally treated at 75 °C. It was mentioned that HPP treatment (600 MPa, 75 °C, 10 min) was the best method for inactivation of N. fischeri JCM 1740 ascospores in apple juice with a 3.3 log cfu/mL reduction compared to no reduction followed by thermal treatment. Similar results were also reported by Filipa et al. [81], that no significant inactivation of A. acidoterrestris spores was observed in orange juice followed by temperature or pressure treatment alone.
However, HPP in conjunction with thermal treatment can effectively inactivate spores with a 2.7 log cfu/mL reduction at 600 MPa, 65 °C, and 10 min. Evelyn et al. [36] also reported that the combination of temperature and pressure treatment was more effective compared to either pressure or thermal alone for inactivating Byssochlamys nivea JCM 12,806 spores in strawberry puree. In addition, the combination of temperature and pressure was mentioned to be efficient compared to traditional thermal processing. As shown in Table 1, most of the models of high-pressure and moderate-temperature synergistic sterilization are nonlinear models. Hossein et al. [34] reported that Weibull distribution was the best mathematical model to describe the non-linear inactivation of B. coagulans 185A spores in tomato juice by high-pressure and moderate temperature synergistic process, whereas first-order kinetics were appropriate to explain microbial inactivation during thermal processing alone.
Evelyn et al. [36] simulated the effect of temperature on Bys. nivea JCM 12,806 ascospores in strawberry puree at 600 MPa. Weibull model efficiently described ascospore inactivation by HPP-thermal treatment (600 MPa and 38, 50, 60, 75 °C). In another study by Shao et al. [85], high-pressure treatment (350 MPa) at 10–30 °C for 5 min presented no significant impact on Escherichia coli K-12 count, whereas HPP at 40 °C was reported to reduce E. coli log survivors. This finding is attributed to the fact that temperature and pressure exhibit opposite effects on volume expansion. The increase in temperature contributes towards volume increase, while pressure has a reversed effect on volume. For similar reasons, some studies have also shown that for microorganisms at sublethal levels, the inactivation effect of pressure was elevated at lower temperatures [86]. In the following section, the effects of a combination of low temperature and high pressure on microorganisms have been reviewed in detail.

3.3.2. High-Pressure and Low-Temperature Synergistic Sterilization in Fruits and Vegetables

High-pressure freezing and thawing has gained the interest of food researchers due to its inherent property to retain the quality of food. High-pressure freezing can improve the freezing effect in essence by applying pressure during freezing at atmospheric pressure. By adjusting the phase transition temperature of the water, this method increases the degree of supercooling and freezing rate during the freezing process and then changes the path of crystal nucleus formation and ice crystal growth [87]. This method has the advantages of rapid heat and mass transfer, formation of small ice crystals during freezing, and even distribution of ice crystals in food tissues and low juice leakage rates during thawing. That in turn protects the texture and nutrient quality of food products [35]. Thus, high-pressure freezing and thawing resolve the problem of irreversible quality damage caused by traditional freeze–thaw processing and successfully meet consumers’ demand for high-quality frozen and thawed food. Unlike high-pressure and moderate-temperature synergistic sterilization, there is the problem of metastable phase transition of water (ice) during high-pressure and low-temperature synergistic sterilization. This metastable phase transition of water plays a crucial role in high-pressure and low-temperature synergistic sterilization. The metastable properties of ice crystals have often been neglected in previous studies on microbial decontamination involving ice I–ice III phase transitions. The temperature–pressure phase diagram of pure water is shown in Figure 1.
During the pressurization process, there are phase transitions from ice I to metastable ice I, metastable ice I to ice III, and recrystallization from water to ice III. However, there are phase transitions from ice III to ice I and from water to ice I during the pressure relief process. The transition from ice I to other ice phases results in an instantaneous decrease in system volume, as shown in Figure 1. This phase transition of ice crystals causes microbial cell damage by mechanical forces [90]. Pedro et al. and Luscher et al. have also mentioned that microbial inactivation during high-pressure and low-temperature synergistic sterilization is caused due to mechanical effects associated with phase transition from ice I to ice III [91,92]. Therefore, crystal formation and phase transition play an important role in the inactivation of pathogenic microorganisms in food samples during high-pressure and low-temperature synergistic treatment. It is important to note that the pressure tolerance of the majority of microbes is decreased at low temperatures. Thus, Su et al. [86] reported a combination or interaction effect of pressure and subzero temperature for the inactivation of microorganisms in food. It was mentioned that phase transitions of Ice I/Ice III by pressurizing frozen systems above 200 MPa are responsible for bacterial destruction [91]. Zhu et al. [35] showed that the high-pressure inactivation effect on E. coli in frozen carrot juice samples was better compared to microbial inactivation in unfrozen samples (Table 1). Researchers have reported a 1.87 log cfu/mL reduction for E. coli in unfrozen carrot juice, while a 6.80 log cfu/mL reduction was mentioned for E. coli in frozen carrot juice after treatment at 330 MPa for 10 min.
Moussa et al. [93] reported an approximate 4 log cfu/mL reduction in E. coli K-12TG1 count in Luria-Bertani medium (liquid, not frozen) followed by high-pressure treatment at 350 MPa for 10 min at −20 °C. However, in the case of frozen samples, a better bactericidal effect (5 log cfu/mL reduction) was observed after HPP at 330 MPa for 10 min. Wang et al. [84] reported 3.5 log cfu/mL reductions for E. coli ATCC 25,922 in frozen bayberry juice after treatment at 300 MPa for less than 5 s, but the same treatment only resulted in 1.2 log cfu/mL reductions in the case of unfrozen sample. In addition, an increase in the pressure holding time resulted in a potent antibacterial effect as E. coli was not detected in the samples after treatment at 300 MPa for 5 min at −5 °C. Sami et al. [83] also reported a 4.88 log cfu/mL reductions for E. coli in frozen orange juice (−80 °C) after treatment at 250 MPa for 15 min. However, without freezing, the same treatment resulted only in a 0.42 log cfu/mL reduction. Moussa et al. [93] reported that high-pressure treatment (150, 250 and 350 MPa) for 10 min at −20 and −10 °C presented a potent antibacterial effect against E. coli K-12TG1 compared to when the sample was treated at 25 °C. The above results show that the microbial inactivation by HPP treatment in frozen samples was higher compared to the unfrozen samples. In addition, below 0 °C, the inactivation effect of high-pressure treatment on tested pathogenic microorganisms was obvious.
In addition to the important role of crystal formation and phase transition in microbial inactivation in frozen samples during HPP, increased sensitivity of proteins to high pressure at low temperatures also leads to the rapid denaturation of proteins. That in turn plays a significant role in the reduction of the pressure resistance of microorganisms. Moreover, the structure of the cell membrane was more vulnerable to damage at low temperatures. In general, only a few studies have reported the impact of combined high-pressure and sub-zero temperatures on food sterilization. In particular, studies on the microbial inactivation kinetics of high-pressure sterilization at sub-zero (freezing) temperatures are scarce [92,94]. Thus, it is of immense importance to study food sterilization using high pressure in conjunction with low temperature (below zero) and its related kinetics.

3.4. Composition of Fruits and Vegetables and Food Additives

The biochemical composition of fruits and vegetables, such as protein, carbohydrate content, and soluble solid content, significantly influence the impact of HPP or HPTP on microbial inactivation [95]. Studies have shown that the biochemical components in the food matrix protect microbes and increase their resistance towards pressure [48,96,97]. In the fruit processing industry, soluble solids content was one of the most important parameters that affects microbial resistance towards pressure and heat [64,98,99]. Rafael et al. [100] reported that in the case of combined high-pressure and moderate temperature treatment, the low content of soluble solids content in the medium lead to higher inactivation of A. acidoterrestris spores. The D-values were 4.17, 7.59, and 13.71 min in broths having 10, 20, and 30 Brix, respectively, showing the protective effect of the soluble solid content against the HPP-thermal process. In general, due to the complex composition of food materials, it is necessary to use the actual food materials as a medium for inoculation during high-pressure treatment. In addition, to enhance the practical application of high-pressure processing, extensive research in this field still needs to be conducted.
Researchers are also using additives in the medium/food matrix to enhance the efficiency of HPP to inactivate pathogenic microorganisms or spores that were difficult to kill by HPP treatment alone. When an antibacterial agent was used as an additive, the bactericidal effect of high-pressure treatment will be enhanced significantly. Julie et al. [101] reported enhanced microbial inactivation by HPP treatment with the addition of natural antimicrobials. Chung et al. [102] studied the effect of high-pressure and the addition of tert-butylhydroquinone on Listeria spp. and E. coli, and the results showed that the addition of the additive significantly improve the bactericidal effect of high-pressure treatment. HPP combined with natural antimicrobial compounds can effectively eliminate the pressure-resistant subpopulation and inhibit their revival or resuscitation [103]. Pokhrel et al. [104] found that a more than 5 log reduction of both L. innocua and E.coli in carrot juice was achieved by the combination of HPP (300 MPa/35 °C/2 min) and natural bacteriocins, compared with less than 1 log reduction in the absence of natural bacteriocins. Zhao et al. found that HPP combined with natural bacteriocins had a significant synergistic effect on reducing the total aerobic bacteria in cucumber juice [105]. The advantage of high-pressure was that there is no specific requirement to add an artificial additive while maintaining better food quality. However, it is worth considering the addition of some natural antimicrobial compounds during the HPP to enhance its antimicrobial capability.

3.5. Water Activity and pH Value

Water availability also affects the resistance of microbial spores towards high-pressure or high-pressure and moderate-temperature synergistic treatment [106]. Thus, the water activity of food materials plays a very important role to change the resistance of microorganisms towards high-pressure treatment. Pressure transfer under high-pressure treatment depends on fluid, so it is not suitable to sterilize dried food, powder, or granular food. Decreased water activity causes cell shrinkage and cell membrane thickening, thereby reducing cell volume and cell membrane fluidity and permeability. Thus, when dealing with dry food material, the addition of water will significantly enhance the bactericidal effect of high-pressure treatment [96,107]. In addition, incomplete spore germination under low water availability conditions may also be one of the reasons for the change in the pressure resistance of microbes [108]. Rodriguez et al. [109] reported that the bactericidal effect of high-pressure on E. coli was significantly weakened when the water activity of samples was reduced. Moussa et al. also reported that low water activity significantly enhanced the pressure resistance of S. cerevisiae [110]. Due to the lack of regulations on the water activity of food material by food authorities, the water activity control due to changes in moisture content, temperature, humidity, and other environmental factors are difficult to change. Unil now, measures to control water activity have not been incorporated into high-pressure sterilization technology. Thus, extensive research is required to include water activity controllers in HPP technology to enhance its microbial inactivation capability.
The pH value of the fruits and vegetables also has a great influence on the pressure resistance of microorganisms during HPP or HPTP. Although pH alone may not be enough to inactivate microorganisms, its combination with high pressure greatly enhances treatment lethality. Microbes exposed to high-pressure processing experience irreversible damage, and cells that are not presenting any physical damage may become sensitive to the high acidity of the medium [48]. The pathogenic microorganisms and their spores present in acidic foods are less resistant to stress compared to the microbes present in low-acid or alkaline foods. In other words, harmful microorganisms and spores in high-acid or acidic foods can be killed with lower pressure and temperature than those in low-acid or alkaline foods. Previously, researchers have acidified foods with citric acid or ascorbic acid to improve the food safety and sterilization efficiency of high-pressure or temperature and combined treatment. High pressure is also reported to affect the pH of the food system, but not to a large extent. Zhang et al. [10] treated carambola puree with high pressure (0–800 MPa, at 25 °C) and found a slight decrease in the pH value of carambola puree with an increase in the pressure. The pH value of the control sample was 4.57, and after 800 MPa treatment, the pH value of the sample was 4.41. The compression of food might cause the ionization of molecules such as H2O, which increases the H+ ion concentration and affect the pH variation of HPP-treated food products [111,112].
Presently, in the bactericidal kinetics of high-pressure or combined temperature and pressure treatments, certain empirical models are extensively used to fit the bactericidal curve. However, various factors affecting the bactericidal effect such as pH, water activity, and natural antibacterial in the medium are not considered. Although curve fitting methods such as first-order kinetics and the Weibull model are effective in evaluating experimental data, other factors such as food composition, pH value, and water activity will also significantly affect the accuracy of the model prediction when establishing the kinetic model of high-pressure sterilization. Therefore, different models are required to predict the kinetics of HPP-induced microbial inactivation and to estimate the effect of HPP under different conditions. In addition, whatever the model is, it needs to be validated in other real fruit and vegetable systems before it can be routinely used throughout the fruit and vegetable processing industry.

4. Conclusions and Future Outlook

High-pressure induced microbial inactivation is also influenced by the temperature, pH, water activity, nutritional composition of fruits and vegetables, and type of microorganism and their growth phase. Therefore, other factors besides pressure should be considered in future inactivation modelling, especially treatment temperature, which can significantly affect inactivation mechanics. Among them, the bactericidal kinetics of low temperature combined with high pressure is the least explored area, and is worthy of further study owing to its inherent advantage of producing high-quality food products. On the other hand, most of the existing studies only use one or two models to simply fit the bactericidal kinetic curve; however, these studies lack optimization, verification, evaluation, and application of the bactericidal model. That in turn results in the poor credibility of the research conclusions and applicability of the model. In addition, in the bactericidal kinetics study of high-pressure or combined temperature and pressure treatment, some empirical models are mainly applied to fit the bactericidal curve, and various factors affecting the bactericidal effect are not considered comprehensively. Therefore, it is necessary to carry out extensive research to obtain more data and an efficient bactericidal kinetic model. In addition, inactivation models need to be validated in other food systems before they can be routinely used throughout the fruit and vegetable processing industry.
In addition, the high cost of high-pressure equipment and high initial investment cost is one of the important reasons that hinders the application of high-pressure technology in the food processing industry. At present, most high-pressure equipment used in the food industry has a rated pressure of ≤600 MPa. If the device operates at high pressure (>500 MPa) for a long time, the maintenance cost of the device will increase. Optimizing high-pressure production processes in food processing can save costs. Therefore, the research on high-pressure food processing technology is of great significance to the high-pressure food processing industry.

Author Contributions

Conceptualization, S.Z. and Y.Y.; investigation, S.Z., M.M. and L.H.; resources, Y.Y.; data curation, S.Z. and J.R.; writing—original draft preparation, S.Z.; writing—review and editing, M.M and S.Z.; supervision, Y.Y. and H.S.R.; project administration, Y.Y.; funding acquisition, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 31871892.

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. WHO. Food Safety. 2022. Available online: https://www.who.int/news-room/fact-sheets/detail/food-safety (accessed on 1 October 2022).
  2. Gillian, H.; Martyn, D.K.; Niels, B.; Joy, E.G.; Leanne, U.; Geoffrey, M.; Russell, S.; Karin, L. Estimating foodborne gastroenteritis, Australia. Emerg. Infect. Dis. 2005, 11, 1257–1264. [Google Scholar]
  3. Lv, M.; Liu, Y.; Geng, J.H.; Kou, X.H.; Xin, Z.H.; Yang, D.Y. Engineering nanomaterials-based biosensors for food safety detection. Biosens. Bioelectron. 2018, 106, 122–128. [Google Scholar] [CrossRef] [PubMed]
  4. Lomeli-Martin, A.; Martínez, L.M.; Welti-Chanes, J.; Escobedo-Avellaneda, Z. Induced Changes in Aroma Compounds of Foods Treated with High Hydrostatic Pressure: A Review. Foods 2021, 10, 878. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, H.; Chen, B.; Wang, C. Comparison of high pressure and high temperature short time processing on quality of carambola juice during cold storage. J. Food Sci. Technol. 2018, 55, 1716–1725. [Google Scholar] [CrossRef] [PubMed]
  6. Artes-Hernandez, F.; Castillejo, N.; Martinez-Zamora, L.; Martinez-Hernandez, G.B. Phytochemical Fortification in Fruit and Vegetable Beverages with Green Technologies. Foods 2021, 10, 2534. [Google Scholar] [CrossRef]
  7. Porebska, I.; Sokolowska, B.; Skapska, S.; Rzoska, S.J. Treatment with high hydrostatic pressure and supercritical carbon dioxide to control Alicyclobacillus acidoterrestris spores in apple juice. Food Control 2017, 73, 24–30. [Google Scholar] [CrossRef]
  8. Yuan, L.; Liang, X.; Pan, X.; Lao, F.; Shi, Y.; Wu, J. Effects of High Hydrostatic Pressure Combined with Vacuum-Freeze Drying on the Aroma-Active Compounds in Blended Pumpkin, Mango, and Jujube Juice. Foods 2021, 10, 3151. [Google Scholar] [CrossRef]
  9. Bleoanca, I.; Patrașcu, L.; Borda, D. Quality and Stability Equivalence of High Pressure and/or Thermal Treatments in Peach–Strawberry Puree. A Multicriteria Study. Foods 2021, 10, 2580. [Google Scholar] [CrossRef]
  10. Zhang, S.N.; Zheng, C.Y.; Zeng, Y.X.; Zheng, Z.H.; Yao, X.S.; Zhao, Y.D.; Jiang, Z. Mechanism of colour change of carambola puree by high pressure processing and its effect on flavour and physicochemical properties. Int. J. Food Sci. Technol. 2021, 56, 5853–5860. [Google Scholar] [CrossRef]
  11. Pang, X.L.; Zhang, Y.Z.; Qiu, J.; Cao, J.M.; Sun, Y.Q.; Li, H.H.; Kong, F.Y. Coupled Multidimensional GC and Odor Activity Value Calculation to Identify Off-Odors in Thermally Processed Muskmelon Juice. Food Chem. 2019, 301, 125307. [Google Scholar] [CrossRef]
  12. Li, X.S.; Zhang, L.; Peng, Z.Y.; Zhao, Y.Q.; Wu, K.Y.; Zhou, N.; Yan, Y.; Ramaswamy, H.S.; Sun, J.X.; Bai, W.B. The Impact of Ultrasonic Treatment on Blueberry Wine Anthocyanin Color and Its In-Vitro Anti-Oxidant Capacity. Food Chem. 2020, 333, 127455. [Google Scholar] [CrossRef]
  13. Paciulli, M.; Meza, I.G.M.; Rinaldi, M.; Ganino, T.; Pugliese, A.; Rodolfi, M.; Barbanti, D.; Morbarigazzi, M.; Chiavaro, E. Improved Physicochemical and Structural Properties of Blueberries by High Hydrostatic Pressure Processing. Foods 2019, 8, 272. [Google Scholar] [CrossRef] [Green Version]
  14. Zhang, S.N.; Zheng, Z.H.; Zheng, C.Y.; Zhao, Y.D.; Jiang, Z. Effect of High Hydrostatic Pressure on Activity, Thermal Stability and Structure of Horseradish Peroxidase. Food Chem. 2022, 379, 132142. [Google Scholar] [CrossRef]
  15. Skegro, M.; Putnik, P.; Kovacevic, D.B.; Kovac, A.P.; Salkic, L.; Canak, I.; Frece, J.; Zavadlav, S.; Jezek, D. Chemometric Comparison of High-Pressure Processing and Thermal Pasteurization: The Nutritive, Sensory, and Microbial Quality of Smoothies. Foods 2021, 10, 1167. [Google Scholar] [CrossRef]
  16. Gao, G.; Zhao, L.; Ma, Y.; Wang, Y.; Sun, Z.; Liao, X. Microorganisms and Some Quality of Red Grapefruit Juice Affected by High Pressure Processing and High Temperature Short Time. Food Bioprocess Technol. 2015, 8, 2096–2108. [Google Scholar] [CrossRef]
  17. Patras, A.; Brunton, N.; Pieve, S.D.; Butler, F.; Downey, G. Effect of Thermal and High Pressure Processing on Antioxidant Activity and Instrumental Colour of Tomato and Carrot Purees. Innov. Food Sci. Emerg. Technol. 2009, 10, 16–22. [Google Scholar] [CrossRef]
  18. Rachna, S.; Barjinder, P.K.; Prabhat, K.N.; Somya, T.; Lokesh, K. Microbial Inactivation by High Pressure Processing: Principle, Mechanism and Factors Responsible. Food Sci. Biotechnol. 2020, 30, 19–35. [Google Scholar]
  19. Aganovic, K.; Hertel, C.; Vogel, R.F.; Johne, R.; Schlüter, O.; Schwarzenbolz, U.; Jäger, H.; Holzhauser, T.; Bergmair, J.; Roth, A.; et al. Aspects of High Hydrostatic Pressure Food Processing: Perspectives on Technology and Food Safety. Compr. Rev. Food Sci. Food Saf. 2021, 20, 3225–3266. [Google Scholar] [CrossRef]
  20. Barba, F.J.; Koubaa, M.; Do, P.L.; Orlien, V.; Sant’Ana, A.D.S. Mild Processing Applied to the Inactivation of the Main Foodborne Bacterial Pathogens: A Review. Trends Food Sci. Technol. 2017, 66, 20–35. [Google Scholar] [CrossRef]
  21. Margosch, D.; Ehrmann, M.A.; Gänzle, M.G.; Vogel, R.F. Comparison of Pressure and Heat Resistance of Clostridium botulinum and Other Endospores in Mashed Carrots. J. Food Prot. 2004, 67, 2530–2537. [Google Scholar] [CrossRef]
  22. Margosch, D.; Ehrmann, M.A.; Buckow, R.; Heinz, V.; Vogel, R.F.; Gänzle, M.G. High-Pressure-Mediated Survival of Clostridium botulinum and Bacillus amyloliquefaciens Endospores at High Temperature. Appl. Environ. Microbiol. 2006, 72, 3476–3481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Evelyn; Filipa, V.M. Heat Assisted HPP for the Inactivation of Bacteria, Moulds and Yeasts Spores in Foods: Log Reductions and Mathematical Models. Trends Food Sci. Technol. 2019, 88, 26–31. [Google Scholar]
  24. Sencer, B.; Hami, A. Modeling Inactivation Kinetics of Food Borne Pathogens at a Constant Temperature. LWT Food Sci. Technol. 2006, 40, 632–637. [Google Scholar]
  25. Koseki, S.; Yamamoto, K. Modelling the Bacterial Survival/Death Interface Induced by High Pressure Processing. Int. J. Food Microbiol. 2007, 116, 136–143. [Google Scholar] [CrossRef] [PubMed]
  26. Chien, S.Y.; Sheen, S.; Sommers, C.H.; Sheen, L.-Y. Modeling the Inactivation of Intestinal Pathogenic Escherichia coli O157:H7 and Uropathogenic, E. coli in Ground Chicken by High Pressure Processing and Thymol. Front. Microbiol. 2016, 7, 920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Peleg, M.; Cole, M.B. Reinterpretation of Microbial Survival Curves. Crit. Rev. Food Sci. Nutr. 1998, 38, 353–380. [Google Scholar] [CrossRef]
  28. Munoz-Cuevas, M.; Guevara, L.; Aznar, A.; Martinez, A.; Periago, P.M.; Fernandez, P.S. Characterisation of the Resistance and the Growth Variability of Listeria monocytogenes after High Hydrostatic Pressure Treatments. Food Control 2013, 29, 409–415. [Google Scholar] [CrossRef]
  29. Buzrul, S.; Alpas, H. Modeling the synergistic effect of high pressure and heat on inactivation kinetics of Listeria innocua: A preliminary study. FEMS Microbiol. Lett. 2004, 238, 29–36. [Google Scholar]
  30. Elisa, G.; Torres, J.A.; Daniel, P.S. Hurdle Approach to Increase the Microbial Inactivation by High Pressure Processing: Effect of Essential Oils. Food Eng. Rev. 2012, 4, 141–148. [Google Scholar]
  31. Tay, A.; Shellhammer, T.H.; Yousef, A.E.; Chism, G.W. Pressure Death and Tailing Behavior of Listeria monocytogenes Strains Having Different Barotolerances. J. Food Prot. 2003, 66, 2057–2061. [Google Scholar] [CrossRef]
  32. Micha, P.A. Model of Microbial Survival after Exposure to Pulsed Electric Fields. J. Sci. Food Agric. 1995, 67, 93–99. [Google Scholar]
  33. Volker, H.; Dietrich, K. High Pressure Inactivation Kinetics of Bacillus subtilis Cells by a Three-State-Model Considering Distributed Resistance Mechanisms. Food Biotechnol. 1996, 10, 149–161. [Google Scholar]
  34. Hossein, D.; Balasubramaniam, V.M. Kinetics of Bacillus coagulans Spore Inactivation in Tomato Juice by Combined Pressure–Heat Treatment. Food Control 2013, 30, 168–175. [Google Scholar]
  35. Zhu, S.M.; Wang, C.F.; Ramaswamy, H.S.; Yu, Y. Phase Transitions during High Pressure Treatment of Frozen Carrot Juice and Influence on Escherichia coli Inactivation. LWT Food Sci. Technol. 2017, 79, 119–125. [Google Scholar] [CrossRef]
  36. Evelyn; Silva, F.V. 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]
  37. Dogan, C.; Erkmen, O. High Pressure Inactivation Kinetics of Listeria monocytogenes Inactivation in Broth, Milk, and Peach and Orange Juices. J. Food Eng. 2004, 62, 47–52. [Google Scholar] [CrossRef]
  38. Smith, R.L. Weibull Regression Models for Reliability Data. Reliab. Eng. Syst. Saf. 1991, 34, 55–77. [Google Scholar] [CrossRef]
  39. Martinus, A.J.S.V. On the Use of the Weibull Model to Describe Thermal Inactivation of Microbial Vegetative Cells. Int. J. Food Microbiol. 2002, 74, 139–159. [Google Scholar]
  40. Pin, C.; Baranyi, J. Kinetics of Single Cells: Observation and Modeling of a Stochastic Process. Appl. Environ. Microbiol. 2006, 72, 2163–2169. [Google Scholar] [CrossRef] [Green Version]
  41. Rodrigo, D.; Ruız, P.; Barbosa, G.V.; Martınez, A.; Rodrigo, M. Kinetic Model for the Inactivation of Lactobacillus plantarum by Pulsed Electric Fields. Int. J. Food Microbiol. 2003, 81, 223–229. [Google Scholar] [CrossRef]
  42. Chen, H. Use of Linear, Weibull, and Log-Logistic Functions to Model Pressure Inactivation of Seven Foodborne Pathogens in Milk. Food Microbiol. 2007, 24, 197–204. [Google Scholar] [CrossRef] [PubMed]
  43. Possas, A.; Perez-Rodriguez, F.; Valero, A.; Garcia-Gimeno, R.M. Modelling the Inactivation of Listeria monocytogenes by High Hydrostatic Pressure Processing in Foods: A Review. Trends Food Sci. Technol. 2017, 70, 45–55. [Google Scholar] [CrossRef]
  44. Li, X.; Farid, M. A Review on Recent Development in Non-Conventional Food Sterilization Technologies. J. Food Eng. 2016, 182, 33–45. [Google Scholar] [CrossRef]
  45. Ramaswamy, H.S.; Shao, Y.W.; Zhu, S.M. High-Pressure Destruction Kinetics of Clostridium sporogenes ATCC 11437 Spores in Milk at Elevated Quasi-Isothermal Conditions. J. Food Eng. 2009, 96, 249–257. [Google Scholar] [CrossRef]
  46. Ahn, J.; Balasubramaniam, V.M.; Yousef, A.E. Inactivation Kinetics of Selected Aerobic and Anaerobic Bacterial Spores by Pressure-Assisted Thermal Processing. Int. J. Food Microbiol. 2007, 113, 321–329. [Google Scholar] [CrossRef]
  47. Lori, S.; Buckow, R.; Knorr, D.; Heinz, V.; Lehmacher, A. Predictive Model for Inactivation of Campylobacter spp. by Heat and High Hydrostatic Pressure. J. Food Prot. 2007, 70, 2023–2029. [Google Scholar] [CrossRef]
  48. Syed, Q.A.; Buffa, M.; Guamis, B.; Saldo, J. Factors Affecting Bacterial Inactivation during High Hydrostatic Pressure Processing of Foods: A Review. Crit. Rev. Food Sci. Nutr. 2016, 56, 474–483. [Google Scholar] [CrossRef]
  49. Agregán, R.; Munekata, P.E.S.; Zhang, W.; Zhang, J.; Pérez-Santaescolástica, C.; Lorenzo, J.M. High-Pressure Processing in Inactivation of Salmonella Spp. in Food Products. Trends Food Sci. Technol. 2021, 107, 31–37. [Google Scholar] [CrossRef]
  50. Koseki, S.; Yamamoto, K. A Novel Approach to Predicting Microbial Inactivation Kinetics during High Pressure Processing. Int. J. Food Microbiol. 2007, 116, 275–282. [Google Scholar] [CrossRef]
  51. Zagorska, J.; Galoburda, R.; Raita, S.; Liepa, M. Inactivation and Recovery of Bacterial Strains, Individually and Mixed, in Milk after High Pressure Processing. Int. Dairy J. 2021, 123, 105147. [Google Scholar] [CrossRef]
  52. 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]
  53. McClements, J.M.; Patterson, M.F.; Linton, M. The Effect of Growth Stage and Growth Temperature on High Hydrostatic Pressure Inactivation of Some Psychrotrophic Bacteria in Milk. J. Food Prot. 2001, 64, 514–522. [Google Scholar] [CrossRef]
  54. Mañas, P.; Mackey, B.M. Morphological and Physiological Changes Induced by High Hydrostatic Pressure in Exponential- and Stationary-Phase Cells of Escherichia coli: Relationship with Cell Death. Appl. Environ. Microbiol. 2004, 70, 1545–1554. [Google Scholar] [CrossRef] [Green Version]
  55. Benito, A.; Ventoura, G.; Casadei, M.; Robinson, T.; Mackey, B. Variation in Resistance of Natural Isolates of Escherichia coli O157 to High Hydrostatic Pressure, Mild Heat, and Other Stresses. Appl. Environ. Microbiol. 1999, 65, 1564–1569. [Google Scholar] [CrossRef] [Green Version]
  56. Mackey, B.M.; Forestière, K.; Isaacs, N. Factors Affecting the Resistance of Listeria monocytogenes to High Hydrostatic Pressure. Food Biotechnol. 1995, 9, 1–11. [Google Scholar] [CrossRef]
  57. Larson, W.P.; Hartzell, T.B.; Diehl, H.S. The Effect of High Pressures on Bacteria. J. Natl. Dent. Assoc. 1918, 5, 271–279. [Google Scholar]
  58. Morales, P.; Calzada, J.; Rodríguez, B.; De Paz, M.; Nuñez, M. Inactivation of Salmonella enteritidis in Chicken Breast Fillets by Single-Cycle and Multiple-Cycle High Pressure Treatments. Foodborne Pathog. Dis. 2009, 6, 577–581. [Google Scholar] [CrossRef] [Green Version]
  59. Fioretto, F.; Cruz, C.; Largeteau, A.; Sarli, T.A.; Demazeau, G.; Moueffak, A.E. Inactivation of Staphylococcus aureus and Salmonella enteritidis in Tryptic Soy Broth and Caviar Samples by High Pressure Processing. Braz. J. Med. Biol. Res. 2005, 38, 1259–1265. [Google Scholar] [CrossRef] [Green Version]
  60. Nikhil, D.H.; Hosahalli, S.R. High-pressure destruction kinetics of spoilage and pathogenic microorganisms in mango juice. J. Food Process. Preserv. 2012, 36, 113–125. [Google Scholar]
  61. Gouvea, F.S.; Padilla-Zakour, O.I.; Worobo, R.W.; Xavier, B.M.; Walter, E.H.; 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]
  62. Donsì, G.; Ferrari, G.; Maresca, P. Pasteurization of Fruit Juices by Means of a Pulsed High Pressure Process. J. Food Sci. 2010, 75, E169–E177. [Google Scholar] [CrossRef] [PubMed]
  63. Margaret, F.P.; Alan, M.M.; Malachy, C.; Mark, L. The Effect of High Hydrostatic Pressure on the Microbiological Quality and Safety of Carrot Juice during Refrigerated Storage. Food Microbiol. 2012, 30, 205–212. [Google Scholar]
  64. Basak, S.; Ramaswamy, H.S.; Piette, J.P.G. High Pressure Destruction Kinetics of Leuconostoc mesenteroides and Saccharomyces cerevisiae in Single Strength and Concentrated Orange Juice. Innov. Food Sci. Emerg. Technol. 2003, 3, 223–231. [Google Scholar] [CrossRef]
  65. Luu, S.; Cruz-Mora, J.; Setlow, B.; Feeherry, F.E.; Doona, C.J.; Setlow, P. The Effects of Heat Activation on Bacillus Spore Germination, with Nutrients or under High Pressure, with or without Various Germination Proteins. Appl. Environ. Microbiol. 2015, 81, 2927–2938. [Google Scholar] [CrossRef] [PubMed]
  66. Sencer, B. Multi-pulsed high hydrostatic pressure inactivation of microorganisms: A review. Innov. Food Sci. Emerg. Technol. 2014, 26, 1–11. [Google Scholar]
  67. Rigaldie, Y.; Largeteau, A.; Demazeau, G.; Lemagnen, G.; Grislain, L. Inactivation of Staphylococcus aureus Using High Hydrostatic Pressure. High Press. Res. 2007, 27, 125–128. [Google Scholar] [CrossRef]
  68. Ramaswamy, H.S.; Riahi, E.; Idziak, E. High-Pressure Destruction Kinetics of E. coli (29055) in Apple Juice. J. Food Sci. 2003, 68, 1750–1756. [Google Scholar] [CrossRef]
  69. Buzrul, S.; Alpas, H.; Largeteau, A.; Demazeau, G. Modeling High Pressure Inactivation of Escherichia coli and Listeria innocua in Whole Milk. Eur. Food Res. Technol. 2008, 227, 443–448. [Google Scholar] [CrossRef]
  70. Sencer, B. Multi-Pulsed High Hydrostatic Pressure Treatment of Foods. Foods 2015, 4, 173–183. [Google Scholar]
  71. Aleman, G.D.; Ting, E.Y.; Mordre, S.C.; Hawes, A.C.O.; Walker, M.; Farkas, D.F.; Torres, J.A. Pulsed Ultra High Pressure Treatments for Pasteurization of Pineapple Juice. J. Food Sci. 1996, 61, 388–390. [Google Scholar] [CrossRef]
  72. Chapleau, N.; Ritz, M.; Delépine, S.; Jugiau, F.; Federighi, M.; de Lamballerie, M. Influence of Kinetic Parameters of High Pressure Processing on Bacterial Inactivation in a Buffer System. Int. J. Food Microbiol. 2006, 106, 324–330. [Google Scholar] [CrossRef]
  73. Ratphitagsanti, W.; Ahn, J.; Balasubramaniam, V.M.; Yousef, A.E. Influence of Pressurization Rate and Pressure Pulsing on the Inactivation of Bacillus amyloliquefaciens Spores during Pressure-Assisted Thermal Processing. J. Food Prot. 2009, 72, 775–782. [Google Scholar] [CrossRef]
  74. Alpas, H.; Kalchayanand, N.; Bozoglu, F.; Ray, B. Interactions of High Hydrostatic Pressure, Pressurization Temperature and PH on Death and Injury of Pressure-Resistant and Pressure-Sensitive Strains of Foodborne Pathogens. Int. J. Food Microbiol. 2000, 60, 33–42. [Google Scholar] [CrossRef]
  75. Houška, M.; Silva, F.V.M.; Evelyn; Buckow, R.; Terefe, N.S.; Tonello, C. High Pressure Processing Applications in Plant Foods. Foods 2022, 11, 223. [Google Scholar] [CrossRef]
  76. Zhi, H.Z.; Lang, H.W.; Xin, A.Z.; Zhong, H.; Charles, S.B. Non-thermal Technologies and Its Current and Future Application in the Food Industry: A Review. Int. J. Food Sci. Technol. 2019, 54, 1–13. [Google Scholar]
  77. Carlez, A.; Rosec, J.P.; Richard, N.; Cheftel, J.C. High Pressure Inactivation of Citrobacter freundii, Pseudomonas fluorescens and Listeria innocua in Inoculated Minced Beef Muscle. LWT Food Sci. Technol. 1993, 26, 357–363. [Google Scholar] [CrossRef]
  78. Scurrah, K.J.; Robertson, R.E.; Craven, H.M.; Pearce, L.E.; Szabo, E.A. Inactivation of Bacillus Spores in Reconstituted Skim Milk by Combined High Pressure and Heat Treatment. J. Appl. Microbiol. 2006, 101, 172–180. [Google Scholar] [CrossRef]
  79. Reddy, N.R.; Tetzloff, R.C.; Solomon, H.M.; Larkin, J.W. Inactivation of Clostridium botulinum Nonproteolytic Type B Spores by High Pressure Processing at Moderate to Elevated High Temperatures. Innov. Food Sci. Emerg. Technol. 2006, 7, 169–175. [Google Scholar] [CrossRef]
  80. 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]
  81. Filipa, V.M.S.; Eng, K.T.; Mohammed, F. Bacterial Spore Inactivation at 45–65 °C Using High Pressure Processing: Study of Alicyclobacillus acidoterrestris in Orange Juice. Food Microbiol. 2012, 32, 206–211. [Google Scholar]
  82. Morgana, Z.; Donald, W.S.; Gláucia, M.F.A. Modeling the Inactivation Kinetics of Bacillus coagulans Spores in Tomato Pulp from the Combined Effect of High Pressure and Moderate Temperature. LWT Food Sci. Technol. 2013, 53, 107–112. [Google Scholar]
  83. Sami, B.; Kimon, A.G.K. Inactivation of Escherichia coli K12 in Phosphate Buffer Saline and Orange Juice by High Hydrostatic Pressure Processing Combined with Freezing. LWT Food Sci. Technol. 2021, 136, 110313. [Google Scholar]
  84. Wang, C.; Lin, Y.; Ramaswamy, H.S.; Ge, L.; Hu, F.; Zhu, S.; Yu, Y. Storage Stability of Chinese Bayberry Juice After High Pressure or Thermal Treatment. J. Food Process. Preserv. 2015, 39, 2259–2266. [Google Scholar] [CrossRef]
  85. Shao, Y.; Ramaswamy, H.S.; Zhu, S.M. High-pressure destruction kinetics of spoilage and pathogenic bacteria in raw milk cheese. J. Food Process. Eng. 2007, 30, 357–374. [Google Scholar] [CrossRef]
  86. Su, G.M.; Yu, Y.; Ramaswamy, H.S.; Hu, F.F.; Xu, M.L.; Zhu, S.M. Kinetics of Escherichia coli Inactivation in Frozen Aqueous Suspensions by High Pressure and Its Application to Frozen Chicken Meat. J. Food Eng. 2014, 142, 23–30. [Google Scholar] [CrossRef]
  87. Cheng, L.; Sun, D.W.; Zhu, Z.; Zhang, Z. Emerging Techniques for Assisting and Accelerating Food Freezing Processes: A Review of Recent Research Progresses. Crit. Rev. Food Sci. Nutr. 2017, 57, 769–781. [Google Scholar] [CrossRef]
  88. Bridgman, P.W. Water, in the Liquid and Five Solid Forms, under Pressure. Proc. Am. Acad. Arts Sci. 1912, 47, 441–558. [Google Scholar] [CrossRef]
  89. Xiao, T.; Li, Y.; Hu, L.; Nie, P.; Ramaswamy, H.S.; Yu, Y. Demonstration of Escherichia Coli Inactivation in Sterile Physiological Saline under High Pressure (HP) Phase Transition Conditions and Analysis of Probable Contribution of HP Metastable Positions Using Model Solutions and Apple Juice. Foods 2022, 11, 1080. [Google Scholar] [CrossRef]
  90. Edebo, L.; Heden, C.G. Disruption of Frozen Bacteria as a Consequence of Changes in the Crystal Structure of Ice. J. Biochem. Microbiol. Technol. Eng. 1960, 2, 113–120. [Google Scholar] [CrossRef]
  91. Pedro, P.F.; Pedro, D.S.; Antonio, D.M.G.; Laura, O.; Bérengère, G.; Sergio, R.V. Conventional Freezing plus High Pressure–Low Temperature Treatment: Physical Properties, Microbial Quality and Storage Stability of Beef Meat. Meat Sci. 2007, 77, 616–625. [Google Scholar]
  92. Luscher, C.; Balasa, A.; Fröhling, A.; Ananta, E.; Knorr, D. Effect of High-Pressure-Induced Ice I-to-Ice III Phase Transitions on Inactivation of Listeria innocua in Frozen Suspension. Appl. Environ. Microbiol. 2004, 70, 4021–4029. [Google Scholar] [CrossRef] [Green Version]
  93. Moussa, M.; Perrier-Cornet, J.M.; Gervais, P. Damage in Escherichia coli Cells Treated with a Combination of High Hydrostatic Pressure and Subzero Temperature. Appl. Environ. Microbiol. 2007, 73, 6508–6518. [Google Scholar] [CrossRef]
  94. Laetitia, P.; Eliane, D.; Joseph-Pierre, G.; Claude, C. Combined High Pressure–Sub-Zero Temperature Processing of Smoked Salmon Mince: Phase Transition Phenomena and Inactivation of Listeria innocua. J. Food Eng. 2004, 68, 43–56. [Google Scholar]
  95. Ramaswamy, H.S.; Hong, J.; Songming, Z. Effects of Fat, Casein and Lactose on High-Pressure Destruction of Escherichia coli K12 (ATCC-29055) in Milk. Food Bioprod. Process. Trans. Inst. Chem. Eng. Part C 2009, 87, 1–6. [Google Scholar] [CrossRef]
  96. Molina-Höppner, A.; Doster, W.; Vogel, R.F.; Gänzle, M.G. Protective Effect of Sucrose and Sodium Chloride for Lactococcus lactis during Sublethal and Lethal High-Pressure Treatments. Appl. Environ. Microbiol. 2004, 70, 2013–2020. [Google Scholar] [CrossRef] [Green Version]
  97. Elaine, P.B.; Thom, H.; Gerald, F.F.; Alan, L.K. Baroprotection of Vegetative Bacteria by Milk Constituents: A Study of Listeria innocua. Int. Dairy J. 2006, 17, 104–110. [Google Scholar]
  98. Palou, E.; López-Malo, A.; Barbosa-Cánovas, G.V.; Welti-Chanes, J.; Davidson, P.M.; Swanson, B.G. Effect of Oscillatory High Hydrostatic Pressure Treatments on Byssochlamys nivea Ascospores Suspended in Fruit Juice Concentrates. Lett. Appl. Microbiol. 1998, 27, 375–378. [Google Scholar] [CrossRef]
  99. Filipa, M.S.; Paul, G.; Margarida, C.V.; Cristina, L.M.S. Thermal Inactivation of Alicyclobacillus acidoterrestris Spores under Different Temperature, Soluble Solids and PH Conditions for the Design of Fruit Processes. Int. J. Food Microbiol. 1999, 51, 95–103. [Google Scholar]
  100. Rafael, U.; Filipa, V.M.S. Alicyclobacillus acidoterrestris Spore Inactivation by High Pressure Combined with Mild Heat: Modeling the Effects of Temperature and Soluble Solids. Food Control 2017, 73, 426–432. [Google Scholar]
  101. Julie, F.; Gil, R.; Jorinde, C.; Chris, W.M. Thiol-Reactive Natural Antimicrobials and High Pressure Treatment Synergistically Enhance Bacterial Inactivation. Innov. Food Sci. Emerg. Technol. 2015, 27, 26–34. [Google Scholar]
  102. Chung, Y.K.; Yousef, A.E. Inactivation of Barotolerant Strains of Listeria monocytogenes and Escherichia coli O157:H7 by Ultra High Pressure and Tert-Butylhydroquinone Combination. J. Microbiol. 2008, 46, 289–294. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, P.; Rao, L.; Zhao, L.; Wu, X.; Wang, Y.; Liao, X. High Pressure Processing Combined with Selected Hurdles: Enhancement in the Inactivation of Vegetative Microorganisms. Compr. Rev. Food Sci. Food Saf. 2021, 20, 1800–1828. [Google Scholar] [CrossRef] [PubMed]
  104. Pokhrel, P.R.; Toniazzo, T.; Boulet, C.; Oner, M.E.; Sablani, S.S.; Tang, J.; Barbosa-Cánovas, G.V. Inactivation of Listeria Innocua and Escherichia Coli in Carrot Juice by Combining High Pressure Processing, Nisin, and Mild Thermal Treatments. Innov. Food Sci. Emerg. Technol. 2019, 54, 93–102. [Google Scholar] [CrossRef]
  105. Zhao, L.; Wang, S.; Liu, F.; Dong, P.; Huang, W.; Xiong, L.; Liao, X. Comparing the Effects of High Hydrostatic Pressure and Thermal Pasteurization Combined with Nisin on the Quality of Cucumber Juice Drinks. Innov. Food Sci. Emerg. Technol. 2013, 17, 27–36. [Google Scholar] [CrossRef]
  106. Lee, S.Y.; Chung, H.J.; Kang, D.H. Combined Treatment of High Pressure and Heat on Killing Spores of Alicyclobacillus acidoterrestris in Apple Juice Concentrate. J. Food Prot. 2006, 69, 1056–1060. [Google Scholar] [CrossRef] [PubMed]
  107. Oxen, P.; Knorr, D. Baroprotective Effects of High Solute Concentrations Against Inactivation of Rhodotorula rubra. LWT Food Sci. Technol. 1993, 26, 220–223. [Google Scholar] [CrossRef]
  108. Elaine, P.B.; Peter, S.; Ailsa, D.H.; Cynthia, M.S.; Alan, L.K.; Dallas, G.H. Response of Spores to High-Pressure Processing. Compr. Rev. Food Sci. Food Saf. 2007, 6, 103–119. [Google Scholar]
  109. Rodriguez, E.; Arques, J.L.; Nuñez, M.; Gaya, P.; Medina, M. Combined Effect of High-Pressure Treatments and Bacteriocin-Producing Lactic Acid Bacteria on Inactivation of Escherichia coli O157:H7 in Raw-Milk Cheese. Appl. Environ. Microbiol. 2005, 71, 3399–3404. [Google Scholar] [CrossRef]
  110. Moussa, M.; Espinasse, V.; Perrier-Cornet, J.M.; Gervais, P. Pressure Treatment of Saccharomyces cerevisiae in Low-Moisture Environments. Appl. Microbiol. Biotechnol. 2010, 85, 165–174. [Google Scholar] [CrossRef]
  111. Kaushik, N.; Kaur, B.P.; Rao, P.S. Application of High Pressure Processing for Shelf Life Extension of Litchi Fruits (Litchi chinensis cv. Bombai) during Refrigerated Storage. Food Sci. Technol. Int. 2014, 20, 527–541. [Google Scholar] [CrossRef]
  112. Morales-de, L.P.M.; Salinas-Roca, B.; Escobedo-Avellaneda, Z.; Martin-Belloso, O.; Welti-Chanes, J. Effect of High Hydrostatic Pressure and Temperature on Enzymatic Activity and Quality Attributes in Mango Puree Varieties (Cv. Tommy Atkins and Manila). Food Bioprocess Technol. 2018, 11, 1211–1221. [Google Scholar] [CrossRef]
Figure 1. The volume of ice crystals (pure water) in individual phases; the letter V in the figure is the corresponding volume of different phases (g/cm3). The image was drawn concerning the data of Bridgman et al. [88] and Xiao et al. [89].
Figure 1. The volume of ice crystals (pure water) in individual phases; the letter V in the figure is the corresponding volume of different phases (g/cm3). The image was drawn concerning the data of Bridgman et al. [88] and Xiao et al. [89].
Foods 11 03698 g001
Table 1. Application of HPP and HPTP bactericidal kinetic models in different fruit and vegetable systems.
Table 1. Application of HPP and HPTP bactericidal kinetic models in different fruit and vegetable systems.
SpeciesFood SystemPressure
(MPa)
Temperature (°C)Model TypeModel ParametersReference
R2
E. coliFrozen carrot juice200−20First-orderD-value (min)4.030.963[35]
2503.970.954
3002.620.974
3502.210.991
4002.120.994
Unfrozen carrot juice3004First-order28.530.850
3509.200.940
4005.320.827
Bys. nivea
JCM 12,806 ascospores
Strawberry
puree
60050Weibullb = 0.16n = 0.570.959[36]
60060b = 0.19n = 0.650.993
60075b = 0.29n = 0.660.997
B. coagulans 185A sporesTomato juice0.1100First-orderD-value (min)1.660.960[34]
0.11050.590.970
60075Weibullb = 0.87n = 0.790.950
60085b = 1.28n = 0.700.940
60095b = 1.93n = 0.680.950
L. monocytogenes 4a KUEN 136Orange juice30025First-orderD-value (min)2.870.970[37]
4001.800.980
6000.870.940
Peach juice30025First-order6.170.950
4003.390.960
6001.520.970
MPa, megapascal; D-value, decimal reduction time; b and n are the Weibull scale and shape factors (Equation (2)), respectively; R2, coefficient of determination.
Table 2. Inactivation effect of HPP on different kinds of microorganisms in the same food systems.
Table 2. Inactivation effect of HPP on different kinds of microorganisms in the same food systems.
SpeciesMicrobial CollectionFood
System
Pressure (MPa)Temperature (°C)Holding Time (min)Log
Reduction
Reference
Zygosaccharomyces bailii ATCC 2333American Type Culture Collection (ATCC) Manassas, VA, USAMango juice3002052.7–2.9[60]
Pichia membranaefaciens ATCC 20853.7–4.3
Leu. mesenteroides ATCC 8293<0.5
L. monocytogenesFood Microbiology Laboratory, New York State
Agricultural Experiment Station (Geneva) and Food Safety Laboratory, Department of Food Science, Cornell University, Ithaca, NY, USA
Açaí juices400516–7[61]
E. coli O157:H73–4
Salmonella spp.3–4
B. coagulans 185A sporesDepartment of Animal & Food Sciences, University of Delaware, Newark, DE, USATomato juice6009513.3[34]
B. coagulans 186A spores3.9
B. coagulans ATCC 7050 spores4.5
Table 3. Inactivation effect of HPP and HPTP on harmful microorganisms in different food systems.
Table 3. Inactivation effect of HPP and HPTP on harmful microorganisms in different food systems.
SpeciesMicrobial CollectionFood
System
Pressure
(MPa)
Holding Time (min)Temperature (°C)Log
Reduction
Reference
B. coagulans ATCC 7050 sporesAndré Tosello Foundation, in Campinas, SP, BrazilTomato pulp30020602.5[82]
500604.5
600606.5
30050<2
500504.0
600504.0
A. acidoterrestris NZRM 4098 sporesNew Zealand Reference Culture; Collection, Medical Section, Fort Richard
Laboratories, New Zealand
Orange juice6001045<1[81]
600551.2
600652.7
200651.7
0.165-
N. fischeri JCM 1740 ascosporesJapan Collection of Microorganisms, Tsukuba, Ibarak, JapanApple juice0.13075-[80]
600381.2
600501.4
600602.8
600754.8
Bys. nivea JCM 12,806 ascosporesJapan Collection of Microorganisms, Tsukuba, Ibarak, JapanStrawberry puree60030380.7[36]
600501.1
600601.8
600702.8
E. coli K12Culture collection of the Department of Food and Nutritional Sciences, University of Reading, Reading, UKOrange juice2501540.42[83]
250−804.88
E. coli ATCC
25,922
-Frozen bayberry juice170<5 s
<5 s
−203[84]
250−203.5
1705−20ND
Unfrozen bayberry juice170<5 s
<5 s
250.5
250251.2
1705251.5
E. coli ATCC 25,922China General Microbiological Culture Collection Center, Beijing, ChinaUnfrozen carrot juice300104<0.2[35]
40041.7
Frozen carrot juice300−204
400−205
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, S.; Meenu, M.; Hu, L.; Ren, J.; Ramaswamy, H.S.; Yu, Y. Recent Progress in the Synergistic Bactericidal Effect of High Pressure and Temperature Processing in Fruits and Vegetables and Related Kinetics. Foods 2022, 11, 3698. https://doi.org/10.3390/foods11223698

AMA Style

Zhang S, Meenu M, Hu L, Ren J, Ramaswamy HS, Yu Y. Recent Progress in the Synergistic Bactericidal Effect of High Pressure and Temperature Processing in Fruits and Vegetables and Related Kinetics. Foods. 2022; 11(22):3698. https://doi.org/10.3390/foods11223698

Chicago/Turabian Style

Zhang, Sinan, Maninder Meenu, Lihui Hu, Junde Ren, Hosahalli S. Ramaswamy, and Yong Yu. 2022. "Recent Progress in the Synergistic Bactericidal Effect of High Pressure and Temperature Processing in Fruits and Vegetables and Related Kinetics" Foods 11, no. 22: 3698. https://doi.org/10.3390/foods11223698

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