Next Article in Journal
Additive SMILES-Based Carcinogenicity Models: Probabilistic Principles in the Search for Robust Predictions
Next Article in Special Issue
Bax Inhibitor-1, a Conserved Cell Death Suppressor, Is a Key Molecular Switch Downstream from a Variety of Biotic and Abiotic Stress Signals in Plants
Previous Article in Journal
Effect of Seven Newly Synthesized and Currently Available Oxime Cholinesterase Reactivators on Cyclosarin-Intoxicated Rats
Previous Article in Special Issue
Protein and Metabolite Analysis Reveals Permanent Induction of Stress Defense and Cell Regeneration Processes in a Tobacco Cell Suspension Culture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bacterial Stressors in Minimally Processed Food

1
Department of Food Science, University of Foggia, via Napoli 25, 71100 Foggia, Italy
2
Department of Biomedical Sciences, University of Foggia, via L. Pinto 1, 71100 Foggia, Italy
3
Department of Production Sciences, Engineering, and Economics for Agricultural Systems (PrIME), University of Foggia, via Napoli 25, 71100 Foggia, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2009, 10(7), 3076-3105; https://doi.org/10.3390/ijms10073076
Submission received: 10 June 2009 / Revised: 29 June 2009 / Accepted: 29 June 2009 / Published: 8 July 2009
(This article belongs to the Special Issue Biotic and Abiotic Stress)

Abstract

:
Stress responses are of particular importance to microorganisms, because their habitats are subjected to continual changes in temperature, osmotic pressure, and nutrients availability. Stressors (and stress factors), may be of chemical, physical, or biological nature. While stress to microorganisms is frequently caused by the surrounding environment, the growth of microbial cells on its own may also result in induction of some kinds of stress such as starvation and acidity. During production of fresh-cut produce, cumulative mild processing steps are employed, to control the growth of microorganisms. Pathogens on plant surfaces are already stressed and stress may be increased during the multiple mild processing steps, potentially leading to very hardy bacteria geared towards enhanced survival. Cross-protection can occur because the overlapping stress responses enable bacteria exposed to one stress to become resistant to another stress. A number of stresses have been shown to induce cross protection, including heat, cold, acid and osmotic stress. Among other factors, adaptation to heat stress appears to provide bacterial cells with more pronounced cross protection against several other stresses. Understanding how pathogens sense and respond to mild stresses is essential in order to design safe and effective minimal processing regimes.

Graphical Abstract

1. Introduction

Different organizations (WHO, FAO, USDA, EFSA) recommend the regular consumption of fruit and vegetables for promoting/maintaining good health. Freshly prepared, ready-to-eat fruits and vegetables are a good example of convenient foods within the context of our modern life. Driven by consumers’ tendencies, the fresh-cut fruit and vegetable industry has expanded rapidly in recent years; as a consequence the production and manufacture of these products is at a stage of innovative dynamics. Consumers require high quality and convenience; to harmonize these demands without compromising safety, it is necessary to implement new preservation technologies. Moreover, some of these new preservation technologies aim at energy saving and being environmentally friendly. Fresh-cut are raw fruits and vegetables that have been washed, peeled, sliced, chopped or shredded prior to being packaged for consumption. These products are typically preserved within semi-permeable packages and stored at refrigeration temperatures. The unit operations in use usually led to the destruction of surface cells making available a potentially richer source of nutrients for microorganisms [1,2]. These factors combined with high aw and either close to neutral (vegetables) or low acidic (many fruits) tissue pH, make easy rapid microbial growth [3,4]. A number of important human pathogens can contaminate fresh-cut produce and there has been an augment in the number of food produce-linked foodborne outbreaks in recent years [5]. In food processing, “mild technologies” are used to describe the technologies for the storage or processing of foods that, in principle, allow to minimize the thermal damage, mechanical and oxidative and chemical and biological contamination that usually accompany such operations unit. Many of these mild preservation technologies aim at being energy saving and environmentally friendly. Ohlsson [6] suggested that minimal processing techniques have emerged to replace traditional harsher methods of food preservation as they retain nutritional and sensory quality better. “Minimal processing” describes non-thermal technologies to process food in a manner to guarantee the food safety and preservation as well as to preserve as much as possible the fresh-like attributes of fruits and vegetables (Figure 1).
This has stimulated interest in the use of mild preservation procedures and the development of the combined effects of several antimicrobial principles in multifactorial preservation approach or hurdle technology [7]. This approach is based on the observation that antimicrobial factors act co-operatively or synergistically with their combined antimicrobial effect being greater than the sum of the individual factors. Numerous reports described such observations for a range of different hurdles [8,9]. The practical outcome of this is that the combined effect of several relatively mild antimicrobial hurdles may offer the desired shelf life and safety properties while retaining many desirable sensory characteristics.
Targeted application of the hurdle concept has become more available as a result of the important improvements in our understanding of the principles of main preservative factors and their interactions [10]. On exposure to stressful conditions such as drying, cold, heat and low pH, stressed bacterial cells may lose their viability, become injured, or express adaptive mechanisms that would help them to survive or even continue growth during stress. These mechanisms begin with stress-sensing followed by producing signals that induce the development of a response that aids adaptation. On sensing stress and developing a signal, cells synthesize mechanisms to cope with the emergent hardship. These mechanisms involve modifications of gene expression and protein activities aiming at preventing or reducing injures to cellular structures and components.
In this paper, we review the molecular basis of bacterial stress response to classical and new stressors that deal with some mild technologies and minimal processing. Our goal is to stimulate the research and the development of molecular targets in order to analyse bacterial cell response in minimally processed food and, particularly, “cross protection”. This biological mechanism, might have an important role in optimizing food preservation procedures and improve process sustainability and global quality of fresh-cut produce, from safety to the healthy properties.

2. Bacterial Pathogens

Food-produce contamination can occur during agricultural production (via animals or insects, soil, water, dirty equipment and human handling), harvesting, processing (cutting, shredding, washing, contaminated work surfaces/equipment, hygiene practices of workers), packaging (contaminated packaging materials/equipment) and transportation and distribution.
Salmonella is the most common cause of disease outbreaks linked to fresh fruit and vegetables. Salmonellae are abundant in faecal material and sewage-polluted water; consequently they may contaminate soil and crops with which they come into contact. Salmonellae from a range of food-produce, including sprouted seeds, cantaloupe melons, tomatoes, unpasteurised citrus juices, rocket and lettuce, have been responsible for several food poisoning outbreaks [11].
Escherichia coli O157:H7 is a member of the enterohemorrhagic group of pathogenic E. coli that has emerged as a foodborne and waterborne pathogen of major public health concern [12]. Fresh produce was not considered a significant vector for the transmission of E. coli O157:H7 until the mid-1990s, when a series of outbreaks associated with minimally processed horticultural products clearly showed that contamination can occur by indirect routes [13,14]. The largest E. coli O157:H7 outbreak ocurred in 1996, when >6,000 school children in Japan were infected with E. coli O157:H7 from white radish seed sprouts. Since 1993, 26 reported outbreaks of E. coli O157:H7 infection have been traced to contaminated lettuce and leafy green vegetables [15]. A recent multi-state outbreak in the USA linked to bagged fresh spinaches caused approximately 205 confirmed illnesses, 31 cases of hemolytic uremic syndrome and three deaths [16,17]. This outbreak was followed by two restaurant associated outbreaks linked to the consumption of pre-washed lettuce [18]. These recent outbreaks have highlighted the dangers of centralised distribution and the great distances that fresh produce travels.
The Gram-positive bacterium Listeria monocytogenes is a food-borne pathogen of both public health and food safety significance. L. monocytogenes is widespread on plants and on agricultural environment generally, and is capable of surviving a variety of environmental stresses, including refrigeration temperatures [19], gas atmospheres commonly present within modified atmosphere packaging (MAP) produce and relatively low pH. Although there is a large amount of literature dedicated to researches concerning with the isolation, attachment, survival and growth of L. monocytogenes on food-produce [20,21], only two fresh-cut produce related listeriosis outbreaks have been documented [22,23].

3. Key Factors Affecting the Survival and Growth of Pathogens

Pathogen survival and growth on food-produce is influenced by a number of interdependent factors, principally storage temperature, product type/combinations, minimal processing operations (e.g. slicing, shredding, washing and decontamination treatments), mild technologies, package atmosphere and competition from the natural microflora present on food-produce. First, each product type has an exclusive combination of compositional and physical characteristics and will have specific growing, harvesting and processing practices, and storage conditions; pathogen growth on food-produce varies significantly with the type of product [24,25]. Storage of food-produce at adequate refrigeration temperatures is probably the single most important factor affecting survival and growth of pathogens. Washing with tap water removes soil and other debris, some of the surface microflora, cell contents and nutrients released throughout slicing that support growth of microorganisms [26]. However, while washing in tap water removes bacteria from exposed surfaces, substantial numbers will remain in hollows at the connection of epidermal cells and in folds in the epidermis [2729]. In addition, due to the re-use of wash water in industry, washing may result in bacterial enrichment and cross-contamination of products rather than decontamination [3,30]. A variety of antimicrobial wash solutions have been used to diminish populations of microorganisms on fresh produce. Chlorine added to water as a solid, liquid or gas is the most frequently used disinfectant for fresh fruits and vegetables [30,31]; however, a wide variety of other disinfectants, including acidic electrolysed water [32], peroxyacetic acid [33], chlorine dioxide [34], hydrogen peroxide [35], organic acids [36], trisodium phosphate [34] and ozone [37] have been also evaluated [38]. Natural antimicrobials from edible plants such as oilseeds, herbs, spices, fruit and vegetables, have been studied for their potential as possible replacements for chemical additives because of their safety for human consumption and broad acceptance from consumers [39,40]. Phenolic compounds present in plant essential oils (EOs) have been shown to possess antimicrobial activity and some are classified as generally recognised as safe (GRAS), and consequently may be useful to prevent post-harvest growth of spoilage and pathogenic bacteria [41]. Efforts to improve the overall effectiveness of the washing step by the use of classical physical treatments, such as mild heat or ethanol vapours, may enhance pathogen destruction. Novel decontamination techniques, including ultra-violet (UV) irradiation, high-pressure treatment, pulsed electric fields, microwave, high intensity pulsed light and thermal destruction using condensing steam, warrant supplementary investigation [42]. When a fresh-cut product is packaged, it continues to breath thereby modifying the gas atmosphere inside the package, hence the term modified atmosphere packaging (MAP). Ideally, O2 levels will fall from 21% in air to 2–5%, and CO2 levels will increase to the 3–10% range. The gas mixtures control the product’s biochemical and enzymatic reactions, inhibit the growth of microorganisms and make longer the shelf-life.
MAP produce harbour large populations of microorganisms including pseudomonads, lactic acid bacteria (LAB) and Enterobacteriaceae [28,43]. The background microflora supply indicators of temperature abuse largely by causing detectable spoilage, and levels can vary considerably for each product and during storage. LAB can exert antibacterial effects due to one or more of the following mechanisms: lowering the pH; generating H2O2 [44]; competing for nutrients; and by producing antimicrobial compounds, such as bacteriocins [45]. The bacteriocins alone were also used in biocontrol of fresh-cut microflora. One other biotic factor to control bacterial pathogens on fresh-cut produce is represented by the use of bacteriophages.

4. Stressors and Related Bacterial Stress Response

Acid adapted L. monocytogenes, Salmonella and E. coli O157:H7 were shown to survive significantly better in acidic foods such as fruit juices, than their non-acid adapted counterpart cells [46,47]. This is due to a specific stress response characteristic of acid stressor, that involved the induction of a specific subset of genes organized into regulons, constituting the acid shock stimulon. A general overview of the stress responses to each single key factor affecting pathogen survival, is shown in Figure 2.

4.1. Cold Stress

Exposure to cold temperatures following harvest in order to minimize and/or inhibit the effects of wounding stress is recognized as one of the principal factors controlling the quality of fresh-cut leafy vegetables [48,49]. Although 0 °C is usually the desirable temperature for most fresh-cut products, in practice many of them are shipped and marketed at temperatures ranging from 5 to 10 °C [50].
In response to temperature downshift, a number of changes take place in prokaryotic cellular physiology such as, (i) decrease in membrane fluidity, (ii) stabilization of secondary structures of nucleic acids leading to reduced efficiency of mRNA translation and transcription, (iii) inefficient folding of some proteins, and (iv) hampered ribosome function [51]. A number of cold shock proteins are induced to cope with these harmful effects of temperature downshift. For all organisms, maintenance of functional cell membranes is a limiting factor for survival. Upon cold shock the physical status of biological membranes is altered from being fluid to becoming rigid. In a process generally termed homeoviscous adaptation, with decreasing temperature, bacteria incorporate fatty acids of lower melting points into lipids in a species-specific mode to re-establish membrane integrity and hence function. In this picture, the introduction of double bonds into acyl chains can either be achieved anaerobically during fatty acids synthesis or aerobically by modification of readily synthesized fatty acids through fatty acid desaturase enzymes. Transcription and translation are closely coupled in bacterial cells. However, transcription machinery and ribosomes generally occupy different subcellular regions in bacteria such as Escherichia coli and Bacillus subtilis, indicating the need for (a) mechanism(s) coupling these processes. A prime function of this mechanism(s) would be ensuring the transfer of unfolded mRNA from the nucleoid to the ribosomes, which need linear mRNA for the initiation of translation. During conditions of a sudden decrease in temperature (cold shock), secondary structures in mRNA would pose an even greater problem for the initiation process. Two conserved classes of proteins, cold shock proteins (CSPs) and cold induced RNA helicases (CSHs), appear to be key players in the prevention of secondary mRNA structures and in transcription/translation coupling. CSPs are general mRNA-binding proteins, and like CSH-type RNA helicases, the presence of at least one csp gene in the cell is essential for viability [52]. E. coli contains nine CSPs (CspA to CspI), of which four (CspA, -B, -G, and -I) are cold shock inducible [53]. CspA, CspC, and CspE are RNA-binding proteins which function as transcriptional antiterminators by preventing the formation of secondary structures in the nascent RNA. Csp-induced transcriptional antitermination is responsible for the increased expression of several genes [54,55]. B. subtilis contains three CSPs (CspB, -C, and -D); CspB is essential for cellular growth in a strain lacking CspC and CspD and plays an important role for efficient protein synthesis at optimal and low temperatures [56], while CspB and CspC are major stationary phase-induced proteins [57].

4.2. Heat Stress

The use of hot water or steam may be a possibility to replace disinfection. Martín-Diana et al. [58] reported that short time exposure of fresh-cut (FC) lettuce to water steam reduced the respiration rate (RR), partially inactivated browning-related enzymes, and kept the mesophilic load as low as with a chlorine treatment. In FC fruits, mild heat pre-treatments (MHPT) (40 °C/70 min or 46 °C for 75 min) were effective in inducing firmness and avoiding browning of the cut surface while preserving their nutritional quality [59,60]. In the post-harvesting of fresh-cut vegetables hot water immersion treatment (HWT) and hot water rinsing and brushing (HWRB) technologies were successfully used [61]. HWT is applied at temperatures between 43 °C and 53 °C for periods of several minutes for the fresh cut, while HWRB is employed commercially for 10-25 s at temperatures between 48 °C and 63 °C. Additionally, oscillating magnetic fields (ohmic heating, dielectric heating, microwaves) represent alternative ways to heating food matrices. The application of high temperatures has been widely used for the elimination of foodborne pathogens. This is due to the effectiveness of heat and its ability to cause damage to diverse structures and components in microbial cells including outer and cytoplasmic membranes, RNA and DNA. It also causes protein denaturation leading to destruction of enzyme activity and enzyme-controlled metabolism in microorganisms [62,63]. Traditional heat treatment techniques widely adopted in food industry, such as pasteurization and sterilization are too harsh to be resisted by vegetative bacterial cells. An interesting concept has been also proposed on sensing environmental stresses including heat. According to this theory, cells produce extracellular proteins to sense stress and act as “alarmones” of the emergent environmental changes [64]. The production of these extracellular components are suggested to provide early warning of stress compared to the cytoplasmic membrane and ribosomes. Following stress-sensing, cells set up strategies to cope with the emergent hardship. These mechanisms involve changes of gene expression and protein activities aiming at preventing or reducing damage to cellular structures and components. An important change is the induction of the synthesis of the so-called “heat shock proteins” (HSPs) [65]. These are highly conserved proteins that act as molecular chaperones or proteases affecting protein folding, repair and degradation under normal and stress conditions [66]. For example, during heat stress, several HSPs such as DnaJ, DnaK, GrpE and GroEL function as chaperones preventing or repairing protein misfolding and thus ensuring their proper functioning. Whereas, other HSPs including ClpP, ClpX and Lon act as proteases catalyzing the degradation of misfolded proteins generated by exposure to stress. Both functions of HSPs help to provide cells with functional proteins which allow survival or growth during heat stress. In a study on heat shock proteins synthesis and heat resistance of Salmonella typhimurium, Mackey and Derrick [67] found four important heat shock proteins, with molecular weights ranging from 25 to 83 KDa, and correlated them with DnaK and GroEL reported in E. coli. The chaperone GroEL was found to diminish heat inactivation of a range of enzymes in vitro [68].

4.3. Acid and Solvent Stress

A decontamination step with lactic acid was evaluated to reduce the microbial contamination of minimally processed vegetables [69]. The antimicrobial effect of 1% and 2% lactic acid was established in potable tap water. Citric acid has been widely accepted as effective in reducing superficial pH of cut fruits [70]. Well documented is the antimicrobial effect of the treatments based on calcium salts on fruits and vegetables [58]. While calcium has a prevalent technological importance, on the other hand, organic calcium salts, with lowering pH (i.e. calcium lactate), may have antimicrobial properties [71].
In recent years, the acid stress response of several prokaryotes has been studied with both proteomics and transcriptomics approaches. A few reports describe the use of these approaches to study the organic acid stress response caused by lactate [73], acetate [74], propionate [74] and formate [75]. In other prokaryotes some enzymatic-transport systems have been found to play a role in pH control or in the maintenance of the proton motive force: a proton-translocating F1 F0 -ATPase [76]; several sodium-proton antiporters [77]; amino acid decarboxylases that use an intracellular hydrogen ion for the decarboxylation of an imported amino acid [76,78]. Several theories have been postulated to explain the toxic effect of organic acids in more detail. One of these considers organic acids as uncouplers that transport protons towards the inside of the cell, which is a pH-driven process. Eventually, this influx could lead to a complete dissipation of the proton motive force [79]. A second aspect relates to the deleterious effects of the lower intracellular pH, caused by the lactic acid. However, whereas many organisms aim at maintaining a constant intracellular pH [80], most anaerobic fermenting species avoid a basic pH keeping a lower intracellular pH, and, as a result, increasing their tolerance to organic acids [81]. A third factor explaining the inhibitory effect of organic acids is the intracellular accumulation of anions, which could lead to both end-product inhibition and a loss of water activity (aw) [82].
A link between acid, ethanol and stress response has been demonstrated in a number of Gram-positive bacteria including food-borne pathogens. The expression of some stress inducible genes identified so far in bacteria is also affected by low pH values and ethanol, suggesting the intersection of different regulatory pathways and overlapping control of gene expression. For instance, acid and solvents such as ethanol and buthanol, induces several heat shock proteins, including DnaK and GroEL. Moreover, the Clp-ATP dependent proteases which degrade aberrant and nonfunctional proteins, arising from stress conditions are also responsible for adaptation to multiple stresses and are inducible by low pH or high ethanol [83].
Ethanol kills organisms by denaturing their proteins and dissolving their lipids and is effective against most bacteria, but is ineffective against bacterial spores [84]. Additionally, ethanol causes water stress by lowering aw and thereby interferes with hydrogen bonds within and between hydrated cell components, ultimately disrupting enzyme and membrane structure and function [85].
Plotto et al. [86] tested ethanol vapours on fresh cut fruit and showed that at lower application rates (8–10 h exposure), ethanol could be used as a safe microbial control in a fresh-cut production sanitation system. Ethanol vapors applied to whole apples reduced ethylene and CO2 production of fresh-cut apples, and their shelf life was increased due to maintenance of visual quality [87].

4.4. Oxidative Stress

Chlorine dioxide is a stable dissolved gas, having a high oxidation and penetration power; ClO2 is a strong bactericide: with minimal contact time, it is highly efficient against pathogenic organisms such as Legionella, Amoebal cysts, Giardia cysts, E. coli, and Cryptosporidium [88]. Hyrdogen peroxide is a powerful bactericide (including spores) and oxidant, being able to generate other cytotoxic oxidising chemical species such as hydroxyl radicals [89]. Electrolyzed water (EW) is formed by adding a very small amount of NaCl (usually about 0.1%) to pure water, and conducting a current across an anode and cathode [90], the cathode area produces alkaline reducing water while the anode area produces acidic oxidizing water [90]; upon release of O•, O3 acts as a strong oxidizing agent being very effective in destroying microorganisms [91]. O3 destroys microorganisms by the progressive oxidation of vital cell components, preventing microbial growth and extending the shelf-life of many fruit and vegetables, and its industrial use is increasing [92].
Oxidative stress is a key stress in bacteria, caused by an imbalance between intracellular oxidant concentration, cellular antioxidant protection and oxidative change of macromolecules (membrane lipids, proteins and DNA repair enzymes) [93,94]. The reactive oxygen species (ROS) and nitrogen species (RNS) are the main causes of oxidative stress [95]. They are principally constituted by the hydroxyl radical (OH), the superoxide anion (O2), hydrogen peroxide (H2O2), organic hydroperoxide (ROOH), peroxynitrite (OONO) and nitric oxide (NO). ROS and RNS cause damages to proteins [96], DNA molecules [97], RNA and lipids leading to negative repercussions of the cellular metabolism functions [98]. The toxicity of ROS/RNS discloses the significant role of competent protection subsystems, for instance the detoxification subsystem that numbers enzymes classified with regard to their substrates, or thioredoxin that help in the cellular defense against several oxidative stresses [99]. Catalases are common enzymes found in almost all-living organisms that catalyze the decomposition of hydrogen peroxide to produce oxygen and water [100]. Peroxidases reduce hydrogen or organic peroxides into water and alcohol moiety. This class of enzymes includes a wide number of phylogenetically unrelated families such as peroxiredoxins [101], rubrerythrins [102], glutathione-peroxidases [103] or haloperoxidases [104]. Superoxide dismutases (SOD) dismute superoxide into hydrogen peroxide and oxygen [105]. An additional mechanism lately reported involves superoxide reductases (SOR), that are non-heme iron proteins [106]. The latter catalyzes the one-electron reduction of superoxide into hydrogen peroxide. Moreover, RNS-scavenging enzymes are essentially globins and nitric oxide reductases [107].

4.5. Osmotic Stress

The inner osmotic pressure of a bacterial cell is normally maintained higher than that of the surrounding medium [108,109], and this is generally referred to as “turgor pressure”. To maintain this turgor pressure in a medium with a high concentration of solutes and to mitigate the osmotic stress resulting from low water availability, microbial cells tend to increase their internal cytoplasmic solute concentration through different mechanisms [110]. Water availability in a food environment is usually assessed by measuring the water activity (aw). Water activity (aw) is defined as the ratio of the water vapour pressure of the food or solution to that of pure water at the same temperature. It indicates the quantity of water available in a material for microbial growth. The values of water activity are represented in a scale range between 0 (no available water) and 1 (pure water). While most bacteria show rapid growth at high aw (0.99), microbial growth does not occur at aw below 0.6 [111]. Several osmosensors are known to be involved in osmotic stress induced responses [112]. As mentioned above, the solute concentration in the bacterial cytoplasm is normally maintained above that of the external environment. However, immediately, following an osmotic up-shift (decrease in aw) in the environment, bacterial cells respond by activation of transporters that aid the cell increase the internal solute concentration by either uptake of inorganic ions into the cell or synthesis and concentration of specific organic solutes to counter the osmotic stress [113]. Under mild osmotic stress, only the ionic solutes are accumulated, whereas other compatible solutes become progressively more important on exposure to severe osmotic stress [114]. These accumulated solutes must not interfere with biochemical processes within the cell and they are thus termed “compatible” solutes [108]. Potassium ions (K+), glutamate (as ionic solutes), glycine betaine, trehalose and proline (as non ionic solutes) are the most important compatible solutes accumulated by bacterial cells [114,115]. Members of the Enterobactereaceae family are reported to synthesize glutamate and trehalose, while K+, glycine betaine and proline are taken up from the medium [110]. Accumulation of K+ during osmotic stress takes place in the initial response, which is then accompanied by increased synthesis of glutamate to preserve electroneutrality in the cytoplasm. This is followed by the accumulation of other compatible solutes such as glycine betaine, proline and trehalose [114,115]. The accumulation of the latter molecules was found to influence that of the originally accumulated compatible solutes since increased levels of trehalose were associated with decreased accumulation of K+ and glutamate [114]. The above mechanism of solute accumulation appears to be affected by environmental temperatures as trehalose accumulation was reported to be enhanced at higher temperatures (45°C) in S. typhimurium [116].

4.6. Irradiation

Radiations have been used both to delay ripening-associated processes and to diminish microorganism growth. Several studies have been published, and in recent times, UV-C has been used as an alternative treatment to preserve the quality of different fruits and vegetables [117]. The use of non-ionizing, germicidal and artificial UV at a wavelength of 190–280nm (UV-C) was found to be effective for surface decontamination of fresh-cut products. Lado and Yousef [118] reported that UV-C radiation from 0.5 to 20 kJm−2 inhibited microbial growth by inducing the formation of pyrimidine dimers which alter the DNA helix and block microbial cell replication. Therefore, cells which are unable to repair radiation damaged DNA die and sub-lethally injured cells are often subject to mutations. Treatment with ultraviolet light is simple to use and lethal to most types of microorganisms [119]. Intense light pulses (ILP) are an interesting decontamination method for food surfaces approved by the US Food and Drug Administration (FDA) that could be appropriate for disinfecting fresh-cut produce. ILP kills microorganisms using short time (from 85 ns to 0.3 ms) high frequency pulses (from 0.45 to 15 Hz) and energy per pulse ranging from 3 to 551 J of an intense broad spectrum, rich in UV-C light [120]. This treatment seems to induce structural changes of microbial DNA, similarly to the effect caused by continuous UV sources, although further mechanisms seem to be involved [121].
Irradiation of DNA with UV light produces a variety of photoproducts, of which the main species are cyclobutane pyrimidine dimers (CPDs) and pyrimidine-pyrimidone. Both lesions, if not repaired, provoke mutagenesis and cell death. To survive in a UV-rich environment, E. coli developed an inducible response known as the SOS response regulated by the recA-lexA regulon. The SOS response aids survival by combining increased expression of genes involved in Nucleotide Excision Repair (NER) and recombinational repair mechanisms. The genes recA for the recombination enzyme, RecA, and uvrA and uvrB for subunits of the UvrABC NER enzymes, UvrA and UvrB, have SOS boxes that are bound by the LexA repressor under physiological conditions. Upon UV irradiation, the constitutive amount of RecA protein binds single-stranded DNA resulting from replication blocks and acts as a coprotease for inactivation of LexA, consequently the levels of RecA, UvrA and UvrB increase. Upon completion of repair, the inducing signal disappears and cells return to the preinduction state [122].

4.7. High Pressure Stress

High pressure treatments of fruits and vegetables have been applied on processed products typically having been processed to some degree [123,124]. High pressures have been used to inhibit enzymes, microorganisms and spores, and to preserve aroma compounds [125]. Yanga et al. [126] indicate an undesirable effect of hyperbaric storage on the synthesis of peach volatiles immediately after storage, however, the post-storage potential for recovery of normal synthesis has not been assessed.
Pressure effects on any physiological or biochemical system basically result from the compression of the system, according to Le Chatelier’s principle, which states that at equilibrium a system tends to minimise the effects of troubling external factors. In lipid membranes, a pressure increase of 1,000 atm is equivalent to a temperature decrease of 20 °C [127]. Pressure increase and temperature decrease result in similar effects, i.e. by ordering structures and reducing flexibility in lipids, nucleic acids and carbohydrates [128]. For proteins, pressure and temperature act in synergy and promote protein denaturation and loss of function [129]. In E.coli O 157:H7, high pressure affected the transcription of many genes involved in a variety of intracellular mechanisms, including the stress response, the thioldisulfide redox system and the Fe-S cluster assembly [130].

4.8. Modified Atmosphere Packaging (MAP)

Producers mainly rely on produce sanitation, refrigeration temperatures, and, more recently MAP to extend shelf life and to reduce microbial load [131]. Modifying the internal atmosphere of a package lowers the oxygen (O2) concentration, from 20% to 0%, hence slowing down the growth of aerobic organisms and the speed of oxidation reactions. The removed oxygen can be replaced with nitrogen (N2), commonly acknowledged as an inert gas, or carbon dioxide (CO2), which can inhibit the growth of bacteria. Although there is a wide literature describing how MAP affects microbial load on various food-produce [132,133], little is known about how growth under subatmospheric oxygen partial pressures would impact the enteric pathogens’ ability to breach the gastric stomach barrier and increase the risk of disease. CO2 inhibits the growth of bacteria by (i) affecting cellular enzymes and decreasing the rate of metabolic reactions, (ii) CO2 product repression of carboxylases and decarboxylases, (iii) disrupting cell membrane structural integrity and/or specific functions, (iv) decreasing the substrate and intra-cellular pH, or by a combination of these mechanisms [134]. The extent of inhibition by CO2 varies with the microorganism, CO2 concentration, temperature of incubation, and type of food [134,135].

4.9. Biological Compounds

Greater consumer awareness and concern regarding synthetic chemical additives have led researchers and food processors to look for natural food additives with a broad spectrum of antimicrobial activity [136]. This is an heterogeneous category: some examples. Plant essential oils and natural aroma compounds are gaining interest for their potential as preservative ingredients or decontaminating treatments, as they have GRAS status and a wide acceptance from consumers [137,138]. The antimicrobial components are commonly found in the essential oil fractions and it is well established that many have a wide spectrum of antimicrobial activity, with potential for control of L. monocytogenes and spoilage bacteria within food systems [139]. Oregano (Origanum vulgare) and thyme (Thymus vulgaris) are amongst the most active EOs, while lemon balm (Melissa officinalis) and marjoram (Origanum majorana) exhibit a good antimicrobial action against Gram-positive and Gram-negative bacteria, respectively [140]. Chitosan, which is a cationic polysaccharide extracted from source of shellfish exoskeletons or the cell walls of some microorganisms and fungi, has been used to preserve the quality of post-harvest fruits and vegetables [141,142]. Martin-Diana et al. [143] tested whey permeate at different concentrations (0.5%, 1.5% and 3%) in the washing treatment of lettuce and carrots, and the results suggest that whey permeate could be a promising alternative for sanitizing fresh-cut vegetables.

4.10. Antagonistic Microflora

Packaged produce harbour large populations of microorganisms including pseudomonads, lactic and bacteria (LAB) and Enterobacteriaceae [28,43]. The background microflora provide indicators of temperature abuse largely by causing detectable spoilage, and levels can vary appreciably for each product and during storage. LAB can exert antibacterial effects due to one or more of the following mechanisms: lowering the pH; generating H2O2; competing for nutrients; and by producing antimicrobial compounds, such as bacteriocins [45]. Cai et al. [144] reported that a large portion of LAB isolates from beansprouts inhibited the growth of L. monocytogenes. Strains of LAB were reported to inhibit Aeromonas hydrophila, L. monocytogenes and S. typhimurium, on vegetable salads [145]. Various researchers have reported antagonism by the native microflora of vegetables against Listeria [146,147]. Reducing the background microflora of endive leaves and shredded lettuce resulted in enhanced growth of Listeria [148]. However, the inhibitory effects were dependent on gas atmosphere; in 3% O2 (balance N2) growth of the mixed population was inhibited while L. monocytogenes proliferated [149]. Enterobacter isolates significantly reduced L. monocytogenes growth during storage on a model lettuce medium; however, the inhibitory activities of Enterobacter decreased as the concentration of CO2 increased [149]. Competitive microflora had a significant effect on the growth of E. coli O157:H7 in broth media [150]. Little is known about the mechanism by which Salmonella manages to compete with natural microflora and survive on plant products [151]. Generally, the complex interactions with the indigenous microflora may have significant effects on survival, growth and biocontrol of pathogens.

4.11. Bacteriocins

Bacteriocins are antimicrobial peptides or proteins produced by strains of different bacterial species. The antimicrobial activity of this set of natural substances against foodborne pathogenic, as well as spoilage bacteria, has raised considerable interest for their application in food preservation [152].
Nisin is the only commercially available bacteriocin recognized as a safe and legal biological food preservative (number E234) by the Food and Agriculture Organization and World Health Organization as well as the FDA. Nisin, a broad-spectrum, pore-forming bacteriocin, is produced by lactic acid bacteria that are often found on food-produce [153]. It is active against many Gram-positive bacteria, including L. monocytogenes [154]. Nisin is particularly active at the lower pH values typical of many fruits and some vegetables [155].
The production of bacteriocin must thus be coupled with a mechanism by which the producing strain can protect itself from the lethal action of its own antimicrobial compound. This mechanism is referred to as immunity. In non-nisin-producing Lactococcus lactis, nisin resistance could be conferred by a specific nisin resistance gene (nsr), which encodes a 35-kDa nisin resistance protein (NSR). NSR is a nisin-degrading protease [156], however, the mechanism underlying NSR-mediated nisin resistance is poorly understood.

4.12. Bacteriophages

Leverentz et al. [157] reported a study on the control of Salmonella by phages on fresh-cut fruits. The use of naturally occurring lytic phages to reduce contamination of fresh-cut produce with foodborne pathogens has several advantages over the use of chemical sanitizers and washes [38]. For example, methods commonly used in industry, such as aqueous washes containing chlorine formulations or plain water, are nonspecific and can achieve a less-than-10-fold reduction in Listeria populations on cut-produce surfaces [158]. Conversely, specific phages attack the targeted pathogens only, thus preserving the competitive potential of the indigenous microflora [38]. Leverentz et al. [159] found that treatment with a Listeria-specific lytic phage cocktail alone or in combination with nisin is an effective method for reducing L. monocytogenes contamination on fresh-cut fruit.
Bacteria have evolved different sophisticated natural bacteriophage defense systems that can interfere with bacteriophage proliferation at different steps during the lytic cycle. These consist of natural phage defense mechanisms that impede the adsorption of the bacteriophage to the cell, mechanisms that inhibit the injection of DNA into the cell, restriction-modification systems, and numerous systems that abort the infection at various points in the replication cycle [160,161,162]. Additionally, Hazan and Engelberg-Kulka [163] demonstrated that E. coli mazEF-mediated cell death acts as a suicide-defense mechanism to protect the bacterial culture against the spread of P1 phage infection.

5. Hurdle Technologies and Cross Protection

The adaptation of bacterial cells to a certain stress is often associated with enhanced protection against other subsequent stresses, which is referred to as “cross protection” [164]. This has important implications in food safety and risk assessment programs, given that preservative tools (stresses) applied for different food products are designed to eliminate microbial loads that have been grown under optimal rather than stress conditions [165]. Several stresses have been shown to induce cross protection, including heat stress [166], cold stress [167], acid stress [168] and osmotic stress [169].
Instead of one robust method such as heat sterilisation, minimal processing involves the use of a number of synergic mild preservation techniques known as hurdles. According to this approach we can look to cross protection as “hurdle stresses” to better understand the effectiveness and the challenges of hurdle technologies. While individually not effective in preventing microbial growth, the right combination of hurdles is a very powerful tool in preventing microbial outgrowth and in minimising organoleptic changes in foods.
The molecular basis of cross response and cross response genes are widely analysed in model Gram-negative and Gram-positive pathogenic bacteria. For instance, CspC and CspE from E. coli regulate the expression of RpoS-regulated stress proteins, such as OsmY, Dps, ProP and KatG, possibly thorough regulation of RpoS itself. These proteins are induced in response to osmotic stress, oxidative stress, or upon stationary phase. CspE and CspC also regulate expression of Universal protein A, UspA, a protein responding to numerous stresses [170]. In addition, CspA homologues are involved in diverse phenomena, such as response to freezing conditions, stationary phase, osmotic stress, starvation, antibiotic biosynthesis, resistance to antimicrobial peptides, inhibition of replication, heat resistance of the spores, UV sensitivity etc. [171,172,173,174]. L. monocytogenes possesses three small, highly homologous protein members of the cold shock protein (Csp) family. Schmid et al. [175] used gene expression analysis and a set of mutants with single, double, and triple deletions of the csp genes to evaluate the roles of CspA, CspB, and CspD in the cold and osmotic (NaCl) stress adaptation responses of L. monocytogenes. The hierarchies of their functional importance differed, depending on the environmental stress conditions: CspA>CspD>CspB in response to cold stress versus CspD>CspA/CspB in response to NaCl salt osmotic stress. The fact that Csps are promoting L. monocytogenes adaptation against both cold and NaCl stress has significant implications in view of practical food microbial control measures, in fact the combined or sequential exposure of L. monocytogenes cells to these two stresses in food environments might inadvertently induce cross-protection responses.
Among other factors, adaptation to heat stress appears to provide bacterial cells with more pronounced cross protection against several other stresses [176]. The ability of chaperones, both protein-based and chemical, to confer cold tolerance in bacteria might be maneuvered to improve the growth rate at low temperatures of some mesophilic bacteria [177]. Heat adaptation of Salmonella is reported to provide protection against subsequent heat treatment [178] and low pH conditions. Similarly, S. typhimurium showed increased resistance to heat and salt following adaptation to acidic condition [47]. This was linked to the observation that the expression of about half the acid shock proteins induced following exposure to acidic conditions were also stimulated by subjecting cells to heat shock [179]. Rowbury [64] reported that damage to DNA is probably the most lethal event in thermal inactivation. Single strand breaks were shown in heat-treated S. typhimurium and other microorganisms. Addition of NaCl (lowering aw) markedly protected DNA against loss of biological activity after heating at 121°C for 15 minutes [180]; probably, NaCl inactivated nuclease rather than lowering aw per se. It is reported that the master stress regulator RpoS may be involved in mediating cross protection in bacteria. This is indicated by the increased level of the alternative sigma factor σs (encoded by the rpoS gene) following exposure to stresses such as osmotic stress, heat stress and starvation [64,108].
El-Sharoud [181] reported that increasing acid resistance of a given bacterium following exposure to other stressful conditions differs among bacteria species. Kim et al. [182] and Abram et al. [183] reported that sigma factor genes sigh (heat shock), sigR (oxidative stress), sigB (osmotic shock), and hrdD, which plays a major role in the secondary metabolism, were all strongly upregulated by the pH shock. A number of heat shock proteins including the DnaK family and chaperones such as GroEL were also observed to be upregulated by the pH shock, while their repressor hspR was strongly downregulated. Oxidative stress-related proteins such as thioredoxin, catalase, superoxide dismutase, peroxidase, and osmotic shock-related protein, such as vesicle synthases, were also upregulated in overall. An interlink between the cold tolerance and acid tolerance of Lactobacillus delbrueckii has been evidenced very recently by the enhanced freeze-tolerance of some cells that were acidified at pH 5.25 for 30 min at the end of fermentation [184].
Studying alternative sigma factor interactions in Salmonella during oxidative stress, Bang et al. [185] discovered that interactions between alternative sigma factors permitted the integration of diverse stress signals to produce coordinated genetic responses, suggesting the hierarchical interactions between alternative sigma factors control sequential gene expression in Gram-positive bacteria, whereas alternative sigma factors in Gram-negative bacteria are generally regarded to direct expression of discrete gene subsets. This consideration is confirmed in Mycobacterium smegmatis, where the alternative sigma factor SigF is required for survival to heat shock, acidic pH and oxidative stress [186]. The response of aerobically grown E. coli cells to the cold shock induced by the rapid lowering of growth temperature from 37 to 20 °C was found to be basically the same as the oxidative stress response [187].
Proteolysis is a powerful mechanism used by cells to control adaptation and recovery after exposure to a variety of stress conditions, first of all characteristic of heat stress response. E. coli has five ATP-dependent proteases: ClpAP, ClpXP, FtsH, HslUV and Lon [188,189,190]. A proteomic study indicated that UvrA is a substrate for degradation by ClpXP [191]. During post-UV recovery, UvrA levels decrease principally as a result of ClpXP-dependent protein degradation, revealing that a complex network of interactions contribute to tuning the level of UvrA in the cell in response to the extent of DNA damage [192].
With respect to solvent stress, among the three characterized small heat shock genes from Lactobacillus plantarum [193,194], Fiocco et al. [195] suggested a potential role for Hsp 18.55 and Hsp 19.3 (small Heat shock proteins) in solvent tolerance. In fact, overproduction of Hsp 18.55 and Hsp 19.3 led to an enhanced survival in the presence of butanol (1% v/v) or ethanol (12% v/v) treatment.
Concerning high pressure and cross-protection, under high hydrostatic pressure the syntheses of some HSPs and CSPs were found to be induced in E. coli [196]. Bacterial ribosomes seem to play the role of intracellular sensors, which integrate the adaptation of the organism to high temperatures, low temperatures and high pressure. In contrast, differences in pressure tolerance of L. monocytogenes strains are not correlated with other stress tolerances [197].
McDougald et al. [198] have found evidence for a large degree of overlap in the cell’s use of global regulators to deal with both starvation and oxidative stress. In addition, the post-transcriptional regulator CsrA (or RsmA) has been reported to play a central role in cross protection (starvation, oxidative stress, virulence) and in the adaptation of baterial pathogens to different stages of infection in animals and also in vegetable/fruit [199].
Giotis et al. [200], in L. monocytogenes, found that alkaline conditions induced cross-protection against osmotic and ethanol challenges; this phenomenon may have serious implications for food safety and human health because such stress conditions are routinely used as part of food preservation and surface cleaning processes.
With respect to biological compounds, we may only remember that: i) it is also well documented that some compounds called chemical chaperones (e.g. glycine, betaine and proline), which are known to stabilize the native conformation of cellular proteins, were found to have a protective role against cold stress, salt stress and thermal stress in bacteria [201,202,203]; ii) treating bacterial cells with two different groups of antibiotics (which all acted on ribosomes), which were found to mimic temperature upshift and downshift of E. coli cells, led to the synthesis of HSPs and cold shock proteins (CSPs), respectively [204].
With regards to biotic stressor, we only underline that psp operon induction, first depicted as a response of E. coli upon infection with filamentous phages [205], was oserved also under more general stress conditions, including extreme heat shock, hyperosmotic stress, ethanol treatment, and uncoupling of proton motive force [206,207].

6. Conclusions

Minimally processed food is easily contaminated by food borne pathogens either directly or via cross-contamination during food preparation. For instance, in addition to Salmonella, L. monocytogenes and E.coli O157:H7, fresh cut produce have been identified as a transmission vehicle for pathogens such as Campylobacter species [208]. The ability of pathogens to survive stress requires specific, co-ordinated responses, which induce resistance to the stressful conditions. The molecular mechanisms involved are complex and there are a number of genes involved in bacterial stress response. For instance, the ability of L. monocytogenes and several Gram-positive bacteria (such as B. subtilis and Staphylococcus aureus) to resist many adverse environmental conditions has been attributed in part to activation of the alternative sigma factor σB, encoded by the sigB gene. Survival under stress involves adaptive responses mediated by a set of conserved proteins (usualy called heat-shock proteins), that are upregulated upon exposure to heat shock, low pH, oxidative agents, toxic chemical compounds, starvation, and in general, any situation in which bacterial growth is arrested indicating a protective role in the general stress response.
Apart from the enhanced survival in foods and increased resistance to subsequent food processing treatments, adapted or hardened pathogens may also have enhanced virulence. Stress response and cross-protection must be considered when current processing technologies are being modified or when new preservation technologies are being developed for fresh-cut produce. These responses are particularly significant in minimal processing technologies used in preparation of fresh-cut produce, where the imposition of one sub-lethal stress may lead to the induction of multiple stress responses that may reduce the efficacy of subsequent treatments [209,210,211]. More research on how to use cumulative sub-lethal hurdles and safe practical interventions, without inducing stress response, is needed.
Finally, the combination of well designed integrated production, handling, processing and distribution chains for fresh-cut produces is crucial for achieving the high quality and safety demanded by consumers [212]. One important strategy might be studying the molecular basis of cross response of human pathogens to develop the most suitable combination of synergic “hurdles”. Moreover, we should take into account that synergic and antagonistic actions of hurdle technologies may be pathogen dependent, and that selected hurdle technology combinations may also improve the bacterial pathogens ability to survive gastric acid conditions without enhancing virulence.

Acknowledgments

Vittorio Capozzi was partially supported by a SfAM Research Development Fund grant.

References

  1. Brackett, RE. Microbiological spoilage and pathogens in minimally processed refrigerated fruits and vegetables. In Minimally processed refrigerated fruits and vegetables; Wiley, RC, Ed.; Chapman and Hall: New York, NY, USA, 1994; pp. 269–312. [Google Scholar]
  2. Barry-Ryan, C; Pacussi, JM; O’Beirne, D. Quality of shredded carrots as affected by packaging film and storage temperature. J. Food Sci 2000, 65, 726–730. [Google Scholar]
  3. Beuchat, LR. Pathogenic microorganisms associated with fresh produce. J. Food Prot 1996, 59, 204–216. [Google Scholar]
  4. Parish, ME; Beuchat, LR; Suslow, TV; Harris, LJ; Garrett, EH; Farber, JN; Busta, FF. Methods to reduce/eliminate pathogens from fresh cut produce. Comp. Rev. Food Sci. Food Safety 2003, 2, 16–173. [Google Scholar]
  5. FDA, US Food and Drug Administration. Guidance for Industry. Guide to Minimize Microbial Food Safety Hazards of Fresh-cut Fruits and Vegetables. Guidance Contains Nonbinding Recommendations. Available online: http://www.cfsan.fda.gov/~dms/prodgui4.html#appe, Last updates: February 2008, accessed February 29, 2008.
  6. Ohlsson, T. Minimal processing technologies in the food industry; Ohlsson, T, Bengtsson, N, Eds.; Woodhead Publishing: Cambridge, UK, 2002. [Google Scholar]
  7. Leistner, L; Gould, GW. Hurdle Technologies: Combination Treatment for Food Stability, Safety and Quality; Kluwer Academic/Plenum Publishers: New York, NY, USA, 2002. [Google Scholar]
  8. Thomas, LV; Wimpenny, JWT. Investigation of the effect of combined variations in temperature, pH and NaCl concentrations on nisin inhibition of Listeria monocytogenes and Staphylococcus aureus. Appl. Environ. Microbiol 1996, 62, 2006–2012. [Google Scholar]
  9. McMeekin, TA; Presser, K; Ratkowsky, D; Ross, T; Salter, M; Tienungoon, S. Quantifying the hurdle concept by modelling the bacterial growth/no growth interface. Int. J. Food Microbiol 2000, 55, 93–98. [Google Scholar]
  10. Leistner, L. Basic aspects of food preservation by hurdle technology. Int. J. Food Microbiol 2000, 55, 181–186. [Google Scholar]
  11. Jones, K; Heaton, J. Microbial contamination of fruit and vegetables: evidence and issues. Microbiol 2006, 7, 28–31. [Google Scholar]
  12. Beneduce, L; Spano, G; Massa, S. Escherichia coli O157:H7 general characteristics, isolation and identification techniques. Annal. Microbiol 2003, 53, 511–527. [Google Scholar]
  13. Delaquis, P; Bach, S; Dinu, LD. Behavior of Escherichia coli O157:H7 in leafy vegetables. J. Food Prot 2007, 70, 1966–1974. [Google Scholar]
  14. Harris, LJ; Farber, JN; Beuchat, LR; Parish, ME; Suslow, TV; Garrett, EH; Busta, FF. Outbreaks associated with fresh produce: incidence, growth and survival of pathogens in fresh and fresh-cut produce. Comp. Rev. Food Sci. Food Safety 2003, 2, 78–141. [Google Scholar]
  15. Maki, D. Don’t eat the spinach - controlling foodborne infectious disease. New Engl. J. Med 2006, 355, 1952–1955. [Google Scholar]
  16. California Food Emergency Response Team 2007. Investigation of an Escherichia coli O157:H7 outbreak associated with dole pre-packaged spinach.
  17. Thurston-Enriquez, J; Watt, P; Dowd, SE; Enriquez, R; Pepper, IL; Gerbe, CP. Detection of protozoan parasites and microsporidia in irrigation waters for crop production. J. Food Prot 2002, 65, 378–382. [Google Scholar]
  18. Iwamoto, M; Taco, Bell; Taco, Johns. Escherichia coli in lettuce outbreaks 2006. CDC Outbreak Response and Surveillance Team Enteric Diseases Epidemiology Branch. Proceedings of the International Association of Food Protection’s 94th Annual Meeting (IAFP 2007), Disney's Contemporary Resort, Lake Buena Vista, FL, USA, July 8–11, 2007.
  19. Bell, C; Kyriakides, A. Background, Listeria a practical approach to the organism and its control in foods; Bell, C, Kyriakides, A, Eds.; Chapman and Hall: London, UK, 1998; pp. 3–29. [Google Scholar]
  20. MacGowan, AP; Bowker, K; McLauchlin, J; Bennett, PM; Reeves, DS. The occurrence and seasonal changes in the isolation of Listeria spp. in shop bought food stuffs, human faeces, sewage and soil from urban sources. Int. J. Food Microbiol 1994, 21, 325–334. [Google Scholar]
  21. Ukuku, DO; Fett, W. 2002, Behaviour of Listeria monocytogenes inoculated on cantaloupe surfaces and efficacy of washing treatments to reduce transfer from rind to fresh-cut pieces. J. Food Prot 2002, 65, 924–930. [Google Scholar]
  22. Ho, JL; Shands, KN; Freidland, G; Eckind, P; Fraser, DW. An outbreak of type 4b Listeria monocytogenes infection involving patients from eight Boston hospitals. Arch. Internal. Med 1986, 146, 520–524. [Google Scholar]
  23. Schlech, WF; Lavigne, PM; Bortolussi, RA; Allen, AC; Haldane, EV; Wort, AJ; Hightower, AW; Johnson, SE; King, SH; Nicholls, ES; Broome, CV. Epidemic listeriosis-evidence for transmission by food. New Eng. J. Med 1983, 308, 203–206. [Google Scholar]
  24. Austin, JW; Dodds, KL; Blanchfield, B; Farber, JM. Growth and toxin production by Clostridium botulinum on inoculated fresh-cut packaged vegetables. J. Food Prot 1998, 61, 324–328. [Google Scholar]
  25. Jacxsens, L; Devlieghere, F; Falcato, P; Debevere, J. Behaviour of Listeria monocytogenes and Aeromonas spp. on fresh-cut produce packaged under equilibrium-modified atmosphere. J. Food Prot 1999, 62, 1128–1135. [Google Scholar]
  26. Bolin, HR; Stafford, AE; King, AD, Jr; Huxsoll, CC. Factors affecting the storage stability of shredded lettuce. J. Food Sci 1977, 42, 1319–1321. [Google Scholar]
  27. Beuchat, LR. Surface disinfection of raw produce. Dairy, Food Environ. Sanitat 1992, 12, 6–9. [Google Scholar]
  28. Nguyen-the, C; Carlin, F. The microbiology of minimally processed fresh fruits and vegetables. Crit. Rev. Food Sci 1994, 34, 371–401. [Google Scholar]
  29. Izumi, H. Electrolyzed water as a disinfectant for fresh-cut vegetables. J. Food Sci 1999, 64, 536–539. [Google Scholar]
  30. Beuchat, LR; Ryu, J-H. Produce handling and processing practices. Emerg. Infect. Dis 1997, 3, 459–465. [Google Scholar]
  31. Adams, MR; Hartley, AD; Cox, LJ. Factors affecting the efficacy of washing procedures used in the production of prepared salads. Food Microbiol 1989, 6, 69–77. [Google Scholar]
  32. Park, C-M; Hung, Y-C; Doyle, MP; Ezeike, GOI; Kim, C. Pathogen reduction and quality of lettuce treated with electrolysed oxidizing and acidified chlorinated water. J. Food Sci 2001, 66, 1368–1372. [Google Scholar]
  33. Park, C-M; Beuchat, LR. Evaluation of sanitizers for killing Escherichia coli O157:H7, Salmonella and naturally occurring microorganisms on cantaloupes, honeydew melons and asparagus. Dairy, Food Environ. Sanit 1999, 19, 842–847. [Google Scholar]
  34. Zhang, S; Farber, JM. The effects of various disinfectants against Listeria monocytogenes on fresh-cut vegetables. Food Microbiol 1996, 13, 311–321. [Google Scholar]
  35. Sapers, GM; Simmons, GF. Hydrogen peroxide disinfection of minimally processed fruits and vegetables. Food Technol 1998, 52, 48–52. [Google Scholar]
  36. Karapinar, M; Gonul, SA. Effects of sodium bicarbonate, vinegar, acetic and citric acids on growth and survival of Yersinia enterocolitica. Int. J. Food Microbiol 1992, 16, 343–347. [Google Scholar]
  37. Burrows, JP; Weber, M; Buchwitz, M; Rozanov, V; Ladsttter-Weienmayer, A; Richter, A; de Beek, R; Hoogen, R; Bramstedt, K; Eichmann, K-U; Eisinger, M; Perner, D. The Global Ozone Monitoring Experiment (GOME): Mission Concept and First Scientific Results. J. Atmos. Sci 1999, 56, 151–175. [Google Scholar]
  38. Beuchat, LR. Survival of enterohemorrhagic Escherichia coli O157:H7 in bovine feces applied to lettuce and the effectiveness of chlorinated water as a disinfectant. J. Food Prot 1999, 62, 845–849. [Google Scholar]
  39. Aureli, P; Costantini, A; Zolea, S. Antimicrobial activity of some plant essential oils against Listeria monocytogenes. J. Food Prot 1992, 55, 344–348. [Google Scholar]
  40. Skandamis, PN; Nychas, GJE. Effect of oregano essential oil on microbiological and physical-chemical attributes of minced meat stored in air and modified atmosphere. J. App. Microbiol 2001, 91, 1011–1022. [Google Scholar]
  41. Singh, N; Singh, RK; Bhunia, AK; Stroshine, RL. Efficacy of chlorine dioxide, ozone and thyme essential oil or a sequential washing in killing Escherichia coli O157:H7 on lettuce and baby carrots. Lebensm. Wiss. U. Technol 2002, 35, 720–729. [Google Scholar]
  42. James, SJ. Novel physical methods for decontamination of produce. Proceedings of IAFP’s Third European Symposium on Food Safety: Advancements in Food Safety, Rome, Italy, October 18–19, 2007.
  43. Francis, GA; Thomas, C; O’Beirne, D. Review paper: The microbiological safety of minimally processed vegetables. Int. J. Food Sci. Technol 1999, 34, 1–22. [Google Scholar]
  44. Price, RJ; Lee, JS. Inhibition of Pseudomonas species by hydrogen peroxide producing lactobacilli. J. Milk Food Technol 1970, 33, 13–18. [Google Scholar]
  45. Arihara, K; Cassens, RG; Luchansky, JB. Characterization of bacteriocins from Enterococcus faecium with activity against Listeria monocytogenes. Int. J. Food Microbiol 1993, 19, 123–134. [Google Scholar]
  46. Gahan, CGM; O’Driscoll, B; Hill, C. Acid adaptation of Listeria monocytogenes can enhance survival in acidic foods and during milk fermentation. Appl. Environ. Microbiol 1996, 62, 3128–3132. [Google Scholar]
  47. Leyer, GL; Johnson, EA. Acid adaptation promotes survival of Salmonella spp. in cheese. Appl. Environ. Microbiol 1992, 58, 2075–2080. [Google Scholar]
  48. Artés, F; Allende, A. Processing lines and alternative preservation techniques to prolong the shelf-life of minimally fresh processed leafy vegetables. Eur. J. Hortic. Sci 2005, 70, 231–245. [Google Scholar]
  49. Hodges, DM; Toivonen, PMA. Quality of fresh-cut fruits and vegetables as affected by exposure to abiotic stress. Postharvest Biol. Technol 2008, 48, 155–162. [Google Scholar]
  50. Aguayo, E; Escalona, V; Artés, F. Quality of fresh-cut tomato as affected by type of cut, packaging, temperature and storage time. Eur. Food Res. Technol 2004, 219, 492–499. [Google Scholar]
  51. Phadtare, S. Recent Developments in Bacterial Cold-Shock Response. Curr. Issues Mol. Biol 2004, 6, 125–136. [Google Scholar]
  52. El-Sharoud, WM; Graumann, PL. Cold shock proteins aid coupling of transcription and translation in bacteria. Sci. Prog 2007, 90, 15–27. [Google Scholar]
  53. Wang, N; Yamanaka, K; Inouye, M. CspI, the ninth member of the CspA family of Escherichia coli is induced upon cold shock. J. Bacteriol 1999, 181, 1603–1609. [Google Scholar]
  54. Bae, W; Xia, B; Inouye, M; Severinov, K. Escherichia coli CspA-family RNA chaperones are transcription antiterminators. Proc. Natl. Acad. Sci. USA 2000, 97, 7784–7789. [Google Scholar]
  55. Phadtare, S; Tyagi, S; Inouye, M; Severinov, K. Three amino acids in Escherichia coli CspE surface-exposed aromatic patch are critical for nucleic acid melting activity leading to transcription antitermination and cold acclimation of cells. J. Biol. Chem 2002, 277, 46706–46711. [Google Scholar]
  56. Graumann, P; Wendrich, TM; Weber, MHW; Schröder, K; Marahiel, MA. A family of cold shock proteins in Bacillus subtilis is essential for cellular growth and for efficient protein synthesis at optimal and low temperature. Mol Microbiol 1997, 25, 741–756. [Google Scholar]
  57. Graumann, P; Marahiel, MA. Cold shock proteins CspB and CspC are major stationary-phase-induced proteins in B. subtilis. Arch. Microbiol 1999, 171, 135–138. [Google Scholar]
  58. Martín-Diana, A; Rico, D; Barry-Ryan, C; Frías, J; Henehan, G; Barat, J. Efficacy of steamer jet-injection as alternative to chlorine in fresh-cut lettuce. Postharvest Biol. Technol 2007, 45, 97–107. [Google Scholar]
  59. Kim, Y; Lounds-Singleton, AJ; Talcott, ST. Antioxidant phytochemical and quality changes associated with hot water immersion treatment of mangoes (Mangifera indica L.). Food Chem 2009, 115, 989–993. [Google Scholar]
  60. Steiner, A; Abreu, M; Correia, L; Beirão-da-Costa, S; Leitão, E; Beirão-da-Costa, ML; Empis, J; Moldão-Martins, M. Metabolic response to combined mild heat pre-treatments and modified atmosphere packaging on fresh-cut peach. Eur. Food Res. Technol 2006, 222, 217–222. [Google Scholar]
  61. Fallik, E. Prestorage hot water treatments (immersion, rinsing and brushing). Postharvest Biol. Technol 2004, 32, 125–134. [Google Scholar]
  62. Ray, B. Methods to Detect Stressed Microorganisms. J. Food Prot 1979, 42, 346–355. [Google Scholar]
  63. Mackey, BM; Miles, CA; Parsons, SE; Seymour, DA. Thermal denaturation of whole cells and cell components of Escherishia coli examined by differential scanning calorimetry. J. Gen. Microbiol 1991, 137, 2361–2374. [Google Scholar]
  64. Rowbury, RJ. Basis of Stress Adaptation, with Particular Reference to the Subversion of Stress Adaptation, and to the Invovement of Extracellular Components in Adaptation. In Microbial Stress Adaptation and Food Safety; Yousef, AE, Juneja, VK, Eds.; CRC Press: Boca Raton, FL, USA, 2002; Chapter 8. [Google Scholar]
  65. Bardwell, JCA; Craig, EA. Major heat shock gene of Drosophila and the Escherichia coli heat inducible dnaK gene are homologous. Proc. Natl. Acad. Sci. USA 1984, 81, 848–852. [Google Scholar]
  66. Yura, T; Kanemori, M; Morita, MT. The heat shock response: regulation and function. In Bacterial Stress Responses; Stortz, G, Hengge-Aronis, R, Eds.; ASM Press: Washington, D.C., USA, 2000; pp. 3–18. [Google Scholar]
  67. Mackey, BM; Derrick, C. Heat shock protein synthesis and thermotolerance in Salmonella typhimurium. J. Appl. Bacteriol 1990, 69, 373–383. [Google Scholar]
  68. Mendoza, JA; Wilson, M; Joves, F; Ackermann, E. Thermostabilization of enzymes by the chaperonin GroEL. Biotechnol. Tech 1996, 10, 535–540. [Google Scholar]
  69. Uyttendaele, M; Neyts, K; Vanderswalmen, H; Notebaert, E; Debevere, J. Control of Aeromonas on minimally processed vegetables by decontamination with lactic acid, chlorinated water, or thyme essential oil solution. Int. J. Food Microbiol 2004, 90, 263–271. [Google Scholar]
  70. Soliva-Fortuny, RC; Martìn-Belloso, O. New advances in extending the shelflife of fresh-cut fruits: a review. Trends Food Sci. Tech 2003, 14, 341–353. [Google Scholar]
  71. Alandes, L; Hernando, I; Quiles, A; Perez-Munuera, I; Lluch, MA. Cell wall stability of fresh-cut Fuji apples treated with calcium lactate. J. Food Sci 2006, 71, 615–620. [Google Scholar]
  72. Pieterse, B; Leer, RJ; Schuren, FHJ; van der Werf, MJ. Unravelling the multiple effects of lactic acid stress on Lactobacillus plantarum by transcription profiling. Microbiol 2005, 151, 3881–3894. [Google Scholar]
  73. Hartke, A; Bouche, S; Giard, J-C; Benachour, A; Boutibonnes, P; Auffray, Y. The lactic acid stress response of Lactococcus lactis subsp. lactis. Curr. Microbiol 1996, 33, 194–199. [Google Scholar]
  74. Polen, T; Rittmann, D; Wendisch, VF; Sahm, H. DNA Microarray analyses of the long-term adaptive response of Escherichia coli to acetate and propionate. Appl. Environ. Microbiol 2003, 69, 1759–1774. [Google Scholar]
  75. Kirkpatrick, C; Maurer, LM; Oyelakin, NE; Yoncheva, YN; Maurer, R; Slonczewski, JL. Acetate and formate stress, opposite responses in the proteome of Escherichia coli. J. Bacteriol 2001, 183, 6466–6477. [Google Scholar]
  76. Cotter, PD; Hill, C. Surviving the acid test, responses of Gram-positive bacteria to low pH. Microbiol. Mol. Biol. Rev 2003, 67, 429–453. [Google Scholar]
  77. Cotter, PD; Gahan, CG; Hill, C. A glutamate decarboxylase system protects Listeria monocytogenes in gastric fluid. Mol Microbiol 2001, 40, 465–475. [Google Scholar]
  78. Azcarate-Peril, MA; Altermann, E; Hoover-Fitzula, RL; Cano, RJ; Klaenhammer, TR. Identification and inactivation of genetic loci involved with Lactobacillus acidophilus acid tolerance. Appl. Environ. Microbiol 2004, 70, 5315–5322. [Google Scholar]
  79. Axe, DD; Bailey, JE. Transport of lactate and acetate through the energized cytoplasmic membrane of Escherichia coli. Biotechnol. Bioeng 1995, 47, 8–19. [Google Scholar]
  80. Shelef, LA. Antimicrobial effects of lactates, a review. J. Food Prot 1994, 57, 445–450. [Google Scholar]
  81. Diez-Gonzalez, F; Russell, JB. The effects of carbonylcyanide-m-chlorophenylhydrazone (CCCP) and acetate on Escherichia coli O157, H7 and K-12, uncoupling versus anion accumulation. FEMS Microbiol. Lett 1997, 151, 71–76. [Google Scholar]
  82. Russell, JB; Diez-Gonzalez, F. The effects of fermentation acids on bacterial growth. Adv. Microbial. Physiol 1998, 39, 205–234. [Google Scholar]
  83. Lemos, JA; Chen, YY; Burne, RA. Genetic and physiologic analysis of the groE operon and role of the HrcA repressor in stress gene regulation and acid tolerance in Streptococcus mutans. J. Bacteriol 183, 6074–6084.
  84. Mcdonnell, G; Denver, RA. Antiseptics and Disinfectants: Activity, Action, and Resistance. Clin. Microbiol. Rev 1999, 12, 147–179. [Google Scholar]
  85. Hallsworth, JE. Ethanol-Induced Water Stress in Yeast. J. Ferment. Bioeng 1998, 85, 125–137. [Google Scholar]
  86. Plotto, A; Baib, J; Narciso, JA; Brecht, JK; Baldwin, EA. Ethanol vapor prior to processing extends fresh-cut mango storage by decreasing spoilage, but does not always delay ripening. Postharvest Biol. Technol 2006, 39, 134–145. [Google Scholar]
  87. Bai, J; Baldwin, EA; Soliva Fortuny, RC; Mattheis, JP; Stanley, R; Perera, C; Brecht, JK. Effect of pretreatment of intact ‘Gala’ apple with ethanol vapor, heat, or 1 methylcyclopropene on quality and shelf life of fresh-cut slices. J. Am. Soc. Hort. Sci 2004, 129, 583–593. [Google Scholar]
  88. Xie, F. Disinfection byproducts in drinking water: form, analysis and control. In Technology & Industrial Arts; CRC Press: Boca Raton, FL, USA, 2003; p. 176. [Google Scholar]
  89. Khadre, MA; Yousef, AE. Sporicidal action of ozone and hydrogen peroxide: a comparative study. Int. J. Food Microbiol 2001, 71, 131–138. [Google Scholar]
  90. Ongeng, D; Devlieghere, F; Devevere, J; Coosemans, J; Ryckeboer, J. The efficacy of electrolysed oxidising water for inactivating spoilage microorganisms in process water on minimally processed vegetables. Int. J. Food Microbiol 2006, 109, 187–197. [Google Scholar]
  91. Guzel-Seydim, Z; Bever, PI; Greene, A. Efficacy of ozone to reduce bacterial populations in the presence of food components. Food Microbiol 2004, 21, 475–479. [Google Scholar]
  92. Parish, M; Beuchat, L; Suslow, T; Harris, L; Garret, E; Farber, J; Busta, F. Methods to reduce or eliminate pathogens from fresh and fresh-cut produce. Comp. Rev. Food Sci. Food Safety 2003, 2, 161–173. [Google Scholar]
  93. Storz, G; Imlay, JA. Oxidative stress. Curr. Opin. Microbiol 1999, 2, 188–194. [Google Scholar]
  94. Thybert, D; Avner, S; Lucchetti-Miganeh, C; Chéron, A; Barloy-Hubler, F. OxyGene: an innovative platform for investigating oxidative-response genes in whole prokaryotic genomes. BMC Genomics 2008, 9, 637. [Google Scholar]
  95. Cabiscol, E; Tamarit, J; Ros, J. Oxidative stress in bacteria and protein damage by reactive oxygen species. Int. Microbiol 2000, 3, 3–8. [Google Scholar]
  96. Poole, LB; Karplus, PA; Claiborne, A. Protein sulfenic acids in redox signaling. Annu. Rev. Pharmacol. Toxicol 2004, 44, 325–347. [Google Scholar]
  97. Aydogan, B; Marshall, DT; Swarts, SG; Turner, JE; Boone, AJ; Richards, NG; Bolch, WE. Site-specific OH attack to the sugar moiety of DNA: a comparison of experimental data and computational simulation. Radiat. Res 2002, 157, 38–44. [Google Scholar]
  98. Rivett, AJ. Regulation of intracellular protein turnover: covalent modification as a mechanism of marking proteins for degradation. Curr. Top. Cell. Regul 1986, 28, 291–337. [Google Scholar]
  99. Ritz, D; Patel, H; Doan, B; Zheng, M; Aslund, F; Storz, G; Beckwith, J. Thioredoxin 2 is involved in the oxidative stress response in Escherichia coli. J. Biol. Chem 2000, 275, 2505–2512. [Google Scholar]
  100. Smulevich, G; Jakopitsch, C; Droghetti, E; Obinger, C. Probing the structure and bifunctionality of catalase-peroxidase (KatG). J. Inorg. Biochem 2006, 100, 568–585. [Google Scholar]
  101. Jonsson, TJ; Lowther, WT. The peroxiredoxin repair proteins. Subcell. Biochem 2007, 44, 115–141. [Google Scholar]
  102. Putz, S; Gelius-Dietrich, G; Piotrowski, M; Henze, K. Rubrerythrin and peroxiredoxin: two novel putative peroxidases in the hydrogenosomes of the microaerophilic protozoon Trichomonas vaginalis. Mol. Biochem. Parasitol 2005, 142, 212–223. [Google Scholar]
  103. Smith, J; Shrift, A. Phylogenetic distribution of glutathione peroxidase. Comp. Biochem. Physiol. B 1979, 63, 39–44. [Google Scholar]
  104. van Pee, KH. Bacterial haloperoxidases and their role in secondary metabolism. Biotechnol. Adv 1990, 8, 185–205. [Google Scholar]
  105. Wuerges, J; Lee, JW; Yim, YI; Yim, HS; Kang, SO; Djinovic Carugo, K. Crystal structure of nickel-containing superoxide dismutase reveals another type of active site. Proc. Natl. Acad. Sci. USA 2004, 101, 8569–8574. [Google Scholar]
  106. Lombard, M; Touati, D; Fontecave, M; Niviere, V. Superoxide reductase as a unique defense system against superoxide stress in the microaerophile Treponema pallidum. J. Biol. Chem 2000, 275, 27021–27026. [Google Scholar]
  107. Heylen, K; Vanparys, B; Gevers, D; Wittebolle, L; Boon, N; De Vos, P. Nitric oxide reductase (norB) gene sequence analysis reveals discrepancies with nitrite reductase (nir) gene phylogeny in cultivated denitrifiers. Environ. Microbiol 2007, 9, 1072–1077. [Google Scholar]
  108. Abee, T; Wouters, JA. Microbial stress response in minimal processing. Int. J. Food Microbiol 1999, 50, 65–91. [Google Scholar]
  109. Wood, JM; Bremer, E; Csonka, LN; Kraemer, R; Poolman, B; van der Heide, T; Smith, LT. Osmosensing and osmoregulatory compatible solute accumulation by bacteria. Comp. Biochem. Physiol. A Mol. Integr. Physiol 2001, 130, 437–460. [Google Scholar]
  110. Csonka, LN. Physiological and Genetic Responses of Bacteria to Osmotic-Stress. Microbiol. Rev 1989, 53, 121–147. [Google Scholar]
  111. Corry, JEL. The Water Relations and Heat Resistance of Micro-organisms. Prog. Ind. Microbiol 1972, 12, 73–108. [Google Scholar]
  112. Poolman, B; Spitzer, JJ; Wood, JM. Bacterial osmosensing: roles of membrane structure and electrostatics in lipid-protein and protein-protein interactions. Biochim Biophys Acta 2004, 1666, 88–104. [Google Scholar]
  113. Poolman, B; Blount, P; Folgering, JHA; Friesen, RHE; Moe, PC; van der Heide, T. How do membrane proteins sense water stress. Mol. Microbiol 2002, 44, 889–902. [Google Scholar]
  114. McLaggan, D; Naprstek, J; Buurman, E; Epstein, W. Interdependence of K+ and glutamate accumulation during osmotic adaptation of Escherichia coli. J. Biol. Chem 1994, 269, 1911–1917. [Google Scholar]
  115. Samelis, J; Sofos, JN. Chapter 9: Strategies to Control Stress-Adapted Pathogens. In Microbial Stress Adaptation and Food Safety; Yousef, AE, Juneja, VK, Eds.; CRC Press: Boca Raton, FL, USA, 2002. [Google Scholar]
  116. Canovas, D; Fletcher, SA; Hayashi, M; Csonka, LN. Role of Trehalose in Growth at High Temperature of Salmonella enterica Serovar Typhimurium. J. Bacteriol 2001, 183, 3365–3371. [Google Scholar]
  117. González-Aguilar, GA; Ruiz-Cruz, S; Cruz-Valenzuela, R; Ayala-Zavala, JF; de la Rosa, LA; Alvarez-Parrilla, E; Gutiérrez-López, GF. New Technologies to Preserve Quality of Fresh-Cut Produce. In Food Engineering: Integrated Approaches; Barbosa-Cánovas, GV, Welti-Chanes, J, Parada-Arias, E, Eds.; Springer: Berlin, Germany, 2008. [Google Scholar]
  118. Lado, B; Yousef, A. Alternative food-preservation technologies: efficacy and mechanisms. Microbes Infect 2002, 4, 433–440. [Google Scholar]
  119. Bintsis, T; Litopoulou-Tzanetaki, E; Robinson, R. Existing and potential applications of ultraviolet light in the food industry—a critical review. J. Sci. Food Agric 2000, 80, 637–645. [Google Scholar]
  120. Gómez-López, V; Devlieghere, F; Bonduellea, V; Debevere, J. Intense light pulses decontamination of minimally processed vegetables and their shelf-life. Int. J. Food Microbiol 2005, 103, 79–89. [Google Scholar]
  121. Takeshita, K; Shibato, J; Sameshima, T; Fukunaga, S; Isobe, S; Arihara, K; Itoh, M. Damage of yeast cells induced by pulsed light irradiation. Int. J. Food Microbiol 2003, 85, 151–158. [Google Scholar]
  122. Kiyosawa, K; Tanaka, M; Matsunaga, T; Nikaido, O; Yamamoto, K. Amplified UvrA protein can ameliorate the ultraviolet sensitivity of an Escherichia coli recA mutant. Mutat. Rese 2001, 487, 149–156. [Google Scholar]
  123. Lopez-Malo, A; Palou, E; Barboza-Canovas, GV; Swanson, BG; Welit-Chanes, J. Minimally processed foods with high hydrostatic pressure. In Trends in Food Engineering; Lozano, JE, Ed.; Technomic Pub Co: Lancaster, PA, USA, 2000; pp. 267–286. [Google Scholar]
  124. Ikeda, F; Baba, T; Como, G; Ohtsubo, T; Lizada, MCC. Effect of hydrostatic pressure on postharvest physiology in fruit. Acta. Hortic 2000, 518, 101–106. [Google Scholar]
  125. Dede, S; Alps, H; Bayindirli, A. High hydrostatic pressure treatment and storage of carrot and tomato juices: antioxidant activity and microbial safety. J. Sci. Food Agric 2007, 87, 773–782. [Google Scholar]
  126. Yanga, DS; Balandrán-Quintanab, RR; Ruizc, CF; Toledoc, RT; Kaysa, SJ. Effect of hyperbaric, controlled atmosphere, and UV treatments on peach volatiles. Postharvest Biol. Technol 2009, 51, 334–341. [Google Scholar]
  127. Pradillon, F; Gaill, F. Pressure and life: some biological strategies. Rev. Environ. Sci. Biotechnol 2007, 6, 181–195. [Google Scholar]
  128. Balny, C; Masson, P; Heremans, K. High pressure effects on biological macromolecules: from structural changes to alteration of cellular processes. Biochim. Biophys. Acta 2002, 1595, 3–10. [Google Scholar]
  129. Balny, C; Mozhaev, VV; Lange, R. Hydrostatic pressure and proteins: Basic concepts and new data. Comp. Biochem. Physiol 1997, 116, 299–304. [Google Scholar]
  130. Malone, AS; Chung, Y-K; Yousef, AE. Genes of Escherichia coli O157:H7 that are involved in high-pressure resistance. Appl. Environ. Microbiol 2006, 72, 2661–2671. [Google Scholar]
  131. Rico, D; Martin-Diana, AB; Barat, JM; Barry-Ryan, C. Extending and measuring the quality of fresh-cut fruit and vegetables: a review. Trends Food Sci. Technol 2007, 18, 373–386. [Google Scholar]
  132. Novak, JS; Sapers, GM; Juneja, VK. Microbial Safety of Minimally Processed Foods; CRC Press: Boca Raton, FL, USA, 2003. [Google Scholar]
  133. Matthews, KR. Microbiology of Fresh Produce; ASM Press: Washington, D.C., USA, 2006. [Google Scholar]
  134. Dixon, NM; Kell, DB. The inhibition by CO2 of the growth and metabolism of micro-organisms: a review. J. Appl. Bacteriol 1989, 67, 109–136. [Google Scholar]
  135. Hudson, JA; Mott, SJ; Penney, N. Growth of Listeria monocytogenes, Aeromonas hydrophila and Yersinia enterocolitica on vacuum and saturated carbon dioxide controlled atmosphere packaged sliced roast beef. J. Food Prot 1994, 57, 204–208. [Google Scholar]
  136. Marino, M; Bersani, C; Comi, G. Impedance measurement to study antimicrobial activity of essential oils from Lamiaceae and Compositae. Int. J. Food Microbiol 2001, 67, 187–195. [Google Scholar]
  137. Burt, S. Essential oils: their antibacterial properties and potential applications in foods - a review. Int. J. Food Microbiol 2004, 94, 223–253. [Google Scholar]
  138. Lanciotti, R; Gianotti, A; Patrignani, F; Belletti, N; Guerzoni, ME; Gardini, F. Use of natural aroma compounds to improve shelflife and safety of minimally processed fruits. Trends Food Sci. Technol 2004, 15, 201–208. [Google Scholar]
  139. Gutierrez, J; Rodriguez, G; Barry-Ryan, C; Bourke, P. Efficacy of plant essential oils against food-borne pathogens and spoilage bacteria associated with ready to eat vegetables: antimicrobial and sensory screening. J. Food Prot 2008, 71, 1846–1854. [Google Scholar]
  140. Gutierrez, J; Barry-Ryan, C; Bourke, P. Antimicrobial activity of plant essential oils using food model media: Efficacy, synergistic potential and interactions with food components. Food Microbiol 2009, 26, 142–150. [Google Scholar]
  141. Campaniello, D; Bevilacqua, A; Sinigaglia, M; Corbo, MR. Chitosan: Antimicrobial activity and potential applications for preservino minimally processed strawberries. Food Microbiol 2009, 25, 992–1000. [Google Scholar]
  142. Chien, PJ; Sheu, F; Lin, HR. Coating citrus (Murcott tangor) fruit with low molecular weight chitosan increases postharvest quality and shelf life. Food Chem 2007, 100, 1160–1164. [Google Scholar]
  143. Martin-Diana, AB; Rico, D; Frias, J; Mulcahy, J; Henehan, GTM; Barry-Ryan, C. Whey permeate as a bio-preservative for shelf life maintenance of fresh-cut vegetables. Inn. Food Sci. Emerg. Tech 2006, 7, 112–123. [Google Scholar]
  144. Cai, Y; Ng, LK; Farber, JM. Isolation and characterization of nisin-producing Lactococcus lactis subsp. lactis from bean-sprouts. J. Appl. Microbiol 1997, 83, 499–507. [Google Scholar]
  145. Vescovo, M; Torriani, S; Orsi, C; Macchiarolo, F; Scolari, G. Application of antimicrobial-producing lactic acid bacteria to control pathogens in ready-to-use vegetables. J. Appl. Bacteriol 1996, 81, 113–119. [Google Scholar]
  146. Liao, C-H; Sapers, GM. Influence of soft rot bacteria on growth of Listeria monocytogenes on potato tuber slices. J. Food Prot 1999, 62, 343–348. [Google Scholar]
  147. Allende, A; Martinez, B; Selma, V; Gil, MI; Suarez, JE; Rodriguez, A. Growth and bacteriocin production by lactic acid bacteria in vegetable broth and their effectiveness at reducing Listeria monocytogenes in vitro and in fresh-cut lettuce. Food Microbiol 2007, 24, 759–766. [Google Scholar]
  148. Carlin, F; Nguyen-the, C; Morris, CE. The influence of the background microflora on the fate of Listeria monocytogenes on minimally processed fresh broad leaved endive (Cichorium endivia var. latifolia). J. Food Prot 1996, 59, 698–703. [Google Scholar]
  149. Francis, GA; O’Beirne, D. Effects of storage atmosphere on Listeria monocytogenes and competing microflora using a surface model system. Int. J. Food Sci. Technol 1998, 33, 465–476. [Google Scholar]
  150. Duffy, G; Whiting, RC; Sheridan, JJ. The effects of a competitive microflora, pH and temperature on the growth kenetics of Escherichia coli O157:H7. Food Microbiol 1999, 16, 299–307. [Google Scholar]
  151. Liao, C-H; Cooke, PH. Response to trisodium phosphate treatment of Salmonella Chester attached to fresh-cut green pepper slices. Can. J. Microbiol 2001, 47, 25–32. [Google Scholar]
  152. Galvez, A; Lopez, RL; Abriouel, H; Valdivia, E; Omar, NB. Application of bacteriocins in the control of foodborne pathogenic and spoilage bacteria. Crit. Rev. Biotechnol 2008, 28, 125–152. [Google Scholar]
  153. Bennik, MHJ; Vorstman, W; Smid, EJ; Gorris, LGM. The influence of oxygen and carbon dioxide on the growth of prevalent Enterobacteriaceae and Pseudomonas species isolated from fresh and controlled-atmosphere-stored vegetables. Food Microbiol 1998, 15, 459–469. [Google Scholar]
  154. Barry-Ryan, C; Pacussi, JM; O’Beirne, D. Quality of shredded carrots as affected by packaging film and storage temperature. J. Food Sci 2000, 65, 726–730. [Google Scholar]
  155. Beuchat, LR. Surface decontamination of fruits and vegetables eaten raw: a review; World Health Organization, Food Safety Unit: Geneva, Switzerland, 1998. WHO/FSF/FOS/98.2, Available online: http://www.who.int/foodsafety/publications/fs_management/en/surface_decon.pdf/ (accessed May 25, 2009).
  156. Sun, Z; Zhong, J; Liang, X; Liu, J; Chen, X; Huan, L. Novel Mechanism for Nisin Resistance via Proteolytic Degradation of Nisin by the Nisin Resistance Protein NSR. Antimicrob. Agents Chemother 2009, 53, 1964–1973. [Google Scholar]
  157. Leverentz, B; Conway, WS; Alavidze, Z; Janisiewicz, WJ; Fuchs, Y; Camp, MJ; Chighladze, E; Sulakvelidze, A. Examination of bacteriophage as a biocontrol method for Salmonella on fresh-cut fruit—a model study. J. Food Prot 2001, 64, 1116–1121. [Google Scholar]
  158. Barry-Ryan, C; O’Beirne, D. Effects of slicing method on the quality and storage-life of modified atmosphere packaged carrot discs. J. Food Sci 1998, 63, 851–856. [Google Scholar]
  159. Leverentz, B; Conway, WS; Camp, MJ; Janisiewicz, WJ; Abuladze, T; Yang, M; Saftner, R; Sulakvelidze, A. Biocontrol of Listeria monocytogenes on Fresh-Cut Produce by Treatment with Lytic Bacteriophages and a Bacteriocin. Appl. Environ. Microbiol 2003, 69, 4519–4526. [Google Scholar]
  160. Coffey, A; Ross, RP. Bacteriophage-resistance systems in dairy starter strains: molecular analysis to application. Antonie Leeuwenhoek 2002, 82, 303–321. [Google Scholar]
  161. Yang, JM; DeUrraza, PJ; Matvienko, N; O’Sullivan, DJ. Involvement of the LlaKR2I Methylase in Expression of the AbiR Bacteriophage Defense System in Lactococcus lactis subsp. lactis biovar diacetylactis KR2. J. Bacteriol 2006, 188, 1920–1928. [Google Scholar]
  162. Chopin, M-C; Chopin, A; Bidnenko, E. Phage abortive infection in lactococci: variations on a theme. Curr. Opin. Microbiol 2005, 8, 473–479. [Google Scholar]
  163. Hazan, R; Engelberg-Kulka, H. Escherichia coli mazEF-mediated cell death as a defense mechanism that inhibits the spread of phage P1. Mol. Gen. Genomics 2004, 272, 227–234. [Google Scholar]
  164. Johnson, EA. Chapter 4: Microbial Adaptation and Survival in Foods. In Microbial Stress Adaptation and Food Safety; Yousef, AE, Juneja, VK, Eds.; CRC Press: Boca Raton, FL, USA, 2002. [Google Scholar]
  165. Rowe, MT; Kirk, RB. Cross-protection phenomenon in Escherichia coli strains harbouring cytotoxic necrotizing factors and cytolethal distending toxins.”. Lett. Appl. Microbiol 2001, 32, 67–70. [Google Scholar]
  166. Lou, YQ; Yousef, AE. Adaptation to sublethal environmental stresses protects Listeria monocytogenes against lethal preservation factors. Appl. Environ. Microbiol 1997, 63, 1252–1255. [Google Scholar]
  167. Periago, PM; Schaik, WV; Abee, T; Wouters, JA. Identification of Proteins Involved in the Heat Stress Response of Bacillus cereus ATCC 14579. Appl. Environ. Microbiol 2002, 68, 3486–3495. [Google Scholar]
  168. Leyer, GJ; Johnson, EA. Acid Adaptation Induces Cross-Protection against Environmental Stresses in Salmonella Typhimurium. Appl. Environ. Microbiol 1993, 59, 1842–1847. [Google Scholar]
  169. Mattick, KL; Jorgensen, F; Legan, JD; Lappin-Scott, HM; Humphrey, TJ. Habituation of Salmonella spp. at Reduced Water Activity and Its Effect on Heat Tolerance. Appl. Environ. Microbiol 2000, 66, 4921–4925. [Google Scholar]
  170. Phadtare, S; Inouye, M. Role of CspC and CspE in regulation of expression of RpoS and UspA, the stress response proteins in Escherichia coli. J. Bacteriol 2001, 183, 1205–1214. [Google Scholar]
  171. Derzelle, S; Hallet, B; Ferain, T; Delcour, J; Hols, P. Improved adaptation to cold-shock, stationary phase, and freezing stresses in Lactobacillus plantarum overproducing cold-shock proteins. Appl. Environ. Microbiol 2003, 69, 4285–4290. [Google Scholar]
  172. Katzif, S; Danavall, D; Bowers, S; Balthazar, JT; Shafer, WM. The major cold shock gene, cspA, is involved in the susceptibility of Staphylococcus aureus to an antimicrobial peptide of human cathepsin. G Infect. Immun 2003, 71, 4304–4312. [Google Scholar]
  173. Mangoli, S; Sanzgiri, VR; Mahajan, SK. A common regulator of cold and radiation response in Escherichia coli. J. Environ. Pathol. Toxicol. Oncol 2001, 20, 23–26. [Google Scholar]
  174. Yamanaka, K; Inouye, M. Growth-phasedependent expression of cspD, encoding a member of the CspA family in Escherichia coli. J. Bacteriol 1997, 179, 5126–5130. [Google Scholar]
  175. Schmid, B; Klumpp, J; Raimann, E; Loessner, MJ; Stephan, R; Tasara, T. Role of cold shock proteins in growth of Listeria monocytogenes under cold and osmotic stress conditions. Appl. Environ. Microbiol 2009, 75, 1621–1627. [Google Scholar]
  176. Ravishankar, S; Juneja, VK. Adaptation or Resistance Responses of Microorganisms to Stresses in the Food Processing Environment. In Microbial Stress Adaptation and Food Safety; Yousef, AE, Juneja, VK, Eds.; CRC Press: Boca Raton, FL, USA, 2002; Chapter 5. [Google Scholar]
  177. Ferrer, M; Chernikova, TN; Yakimov, MM; Golyshin, PN; Timmis, KN. Chaperonins govern growth of Escherichia coli at low temperatures. Nat. Biotechnol 2003, 21, 1266–1267. [Google Scholar]
  178. Juneja, VK; Klein, PG; Marmer, BS. Heat shock and thermotolerance of Escherichia coli O157:H7 in a model beef gravy system and ground beef. J. Appl. Microbiol 1998, 84, 677–684. [Google Scholar]
  179. Foster, JW. Low pH adaptation and the acid tolerance response of Salmonella typhimurium. Crit. Rev. Microbiol 1995, 21, 215–237. [Google Scholar]
  180. Masters, CI; Miles, CA; Mackey, BM. Survival and biological activity of heat damaged DNA. Lett. Appl. Microbiol 1998, 27, 279–282. [Google Scholar]
  181. El-Sharoud, WM. Ribosome inactivation for preservation: concepts and reservations. Sci. Progr 2004, 87, 137–152. [Google Scholar]
  182. Kim, YJ; Moon, MH; Song, JY; Smith, CP; Hong, S-K; Chang, YK. Acidic pH shock induces the expressions of a wide range of stress-response genes. BMC Genomics 2008, 9, 604. [Google Scholar] [Green Version]
  183. Abram, F; Starr, E; Karatzas, KAG; Matlawska-Wasowska, K; Boyd, A; Wiedmann, M; Boor, KJ; Connally, D; O’Byrne, CP. Identification of Components of the Sigma B Regulon in Listeria monocytogenes That Contribute to Acid and Salt Tolerance. Appl. Environ. Microbiol 2008, 74, 6848–6858. [Google Scholar]
  184. Streit, F; Delettre, J; Corrieu, G; Béal, C. Acid adaptation of Lactobacillus delbrueckii subsp. bulgaricus induces physiological responses at membrane and cytosolic levels that improves cryotolerance. J. Appl. Microbiol 2008, 105, 1071–1080. [Google Scholar]
  185. Bang, I-S; Frye, JG; McClelland, M; Velayudhan, J; Fang, FC. Alternative sigma factor interactions in Salmonella : σE and σH promote antioxidant defences by enhancing σS levels. Mol. Microbiol 2005, 56, 811–823. [Google Scholar]
  186. Gebhard, S; Hümpel, A; McLellan, AD; Cook, GM. The alternative sigma factor SigF of Mycobacterium smegmatis is required for survival of heat shock, acidic pH and oxidative stress. Microbiol 2008, 154, 2786–2795. [Google Scholar]
  187. Smirnova, GV; Zakirova, ON; Oktyabrskii, ON. The Role of Antioxidant Systems in the Cold Stress Response of Escherichia coli. Microbiology 2001, 70, 45–50. [Google Scholar]
  188. Gottesman, S. Proteolysis in bacterial regulatory circuits. Annu. Rev. Cell Dev. Biol 2003, 19, 565–587. [Google Scholar]
  189. Sauer, RT; Bolon, DN; Burton, BM; Burton, RE; Flynn, JM; Grant, RA. Sculpting the proteome with AAA(+) proteases and disassembly machines. Cell 2004, 119, 9–18. [Google Scholar]
  190. Hanson, PI; Whiteheart, SW. AAA+ proteins: have engine, will work. Nat. Rev. Mol. Cell Biol 2005, 6, 519–529. [Google Scholar]
  191. Neher, SB; Villen, J; Oakes, EC; Bakalarski, CE; Sauer, RT; Gygi, SP; Baker, TA. Proteomic profiling of ClpXP substrates after DNA damage reveals extensive instability within SOS regulon. Mol. Cell 2006, 22, 193–204. [Google Scholar]
  192. Pruteanu, M; Baker, TA. Controlled degradation by ClpXP protease tunes the levels of the excision repair protein UvrA to the extent of DNA damage. Mol. Microbiol 2009, 71, 912–924. [Google Scholar]
  193. Spano, G; Capozzi, V; Vernile, A; Massa, S. Cloning, molecular characterization and expression analysis of two small heat shock genes isolated from wine Lactobacillus plantarum. J. Appl. Microbiol 2004, 97, 774–782. [Google Scholar]
  194. Spano, G; Beneduce, L; Perrotta, C; Massa, S. Cloning and characterization of the hsp 18.55 gene, a new member of the small heat shock genes family isolated from wine Lactobacillus plantarum. Res. Microbiol 2005, 156, 219–224. [Google Scholar]
  195. Fiocco, D; Capozzi, V; Goffin, P; Hols, P; Spano, G. Improved adaptation to heat, cold, and solvent tolerance in Lactobacillus plantarum. Appl. Microbiol. Biotechnol 2007, 77, 909–915. [Google Scholar]
  196. Welch, TJ; Farewell, A; Neidhardt, FC; Bartlett, DH. Stress response of Escherichia coli to elevated hydrostatic pressure. J. Bacteriol 1993, 175, 7170–7177. [Google Scholar]
  197. Chen, H; Neetoo, H; Ye, M; Joerger, RD. Differences in pressure tolerance of Listeria monocytogenes strains are not correlated with other stress tolerances and are not based on differences in CtsR. Food Microbiol 2009, 26, 404–408. [Google Scholar]
  198. McDougald, D; Gong, L; Srinivasan, S; Hild, E; Thompson, L; Takayama, K; Rice, SA; Kjelleberg, S. Defences against oxidative stress during starvation in bacteria. Antonie van Leeuwenhoek 2002, 81, 3–13. [Google Scholar]
  199. Lucchetti-Miganeh, C; Burrowes, E; Baysse, C; Ermel, G. The post-transcriptional regulator CsrA plays a central role in the adaptation of bacterial pathogens to different stages of infection in animal hosts. Microbiology 2008, 154, 16–29. [Google Scholar]
  200. Giotis, ES; Julotok, M; Wilkinson, BJ; Blair, IS; McDowell, DA. Role of sigma B factor in the alkaline tolerance response of Listeria monocytogenes 10403S and cross-protection against subsequent ethanol and osmotic stress. J. Food Prot 2008, 71, 1481–1485. [Google Scholar]
  201. Shahjee, HM; Banerjee, K; Ahmad, F. Comparative analysis of naturally occurring L-amino acid osmolytes and their D-isomers on protection of Escherichia coli against environmental stresses. J. Biosci 2002, 27, 515–520. [Google Scholar]
  202. Chattopadhyay, MK. The cryoprotective effects of glycine betaine on bacteria. Trends Microbiol 2002, 10, 311. [Google Scholar]
  203. Chattopadhyay, MK; Kern, R; Mistou, MY; Dandekar, AM; Uratsu, SL; Richarme, G. The chemical chaperone proline relieves the thermosensitivity of a dnaK deletion mutant at 42°C. J. Bacteriol 2004, 186, 8149–8152. [Google Scholar]
  204. VanBogelen, RA; Neidhardt, FC. Ribosomes as sensors of heat and cold shock in Escherichia coli. Proc. Natl. Acad. Sci. USA 1990, 87, 5589–5593. [Google Scholar]
  205. Brissette, JL; Russel, M; Weiner, L; Model, P. Phage shock protein, a stress protein of Escherichia coli. Proc. Natl. Acad. Sci. USA 1990, 87, 862–866. [Google Scholar]
  206. Lloyd, LJ; Jones, SE; Jovanovic, G. Identification of a new member of the phage shock protein response in Escherichia coli, the phage shock protein G (PspG). J. Biol. Chem 2004, 279, 55707–55714. [Google Scholar]
  207. Hassani, AS; Malekzadeh, F; Amirmozafari, N; Hamdi, K; Ordouzadeh, N; Ghaemi, A. Phage Shock Protein G. A Novel Ethanol-Induced Stress Protein in Salmonella typhimurium. Curr. Microbiol 2009, 58, 239–244. [Google Scholar]
  208. Hussain, I; Mahmood, MS; Akhtar, M; Khan, A. Prevalence of Campylobacter species in meat, milk and other food commodities in Pakistan. Food Microbiol 2007, 24, 219–222. [Google Scholar]
  209. Bhagwat, AA. Microbiological safety of fresh-cut produce: where are we now? In Microbiology of Fresh Produce; Matthews, KR, Ed.; American Society for Microbiology: Washington, D.C., USA, 2006; pp. 121–165. [Google Scholar]
  210. Skandamisa, PN; Yoona, Y; Stopfortha, JD; Kendallb, PA; Sofos, JN. Heat and acid tolerance of Listeria monocytogenes after exposure to single and multiple sublethal stresses. Food Microbiol 2008, 25, 294–303. [Google Scholar]
  211. Chua, DK; Goh, R; Saftner, A; Bhagwat, AA. Fresh-Cut Lettuce in Modified Atmosphere Packages Stored at Improper Temperatures Supports Enterohemorrhagic E. coli Isolates to Survive Gastric Acid Challenge. J. Food Sci 2008, 73, M148–153. [Google Scholar]
  212. Artés, F; Gómez, P; Aguayo, E; Escalona, V; Artés-Hernández, F. Sustainable sanitation techniques for keeping quality and safety of fresh-cut plant commodities. Postharvest Biol. Technol 2009, 51, 287–296. [Google Scholar]
Figure 1. Example of fresh-cut lettuce operation.
Figure 1. Example of fresh-cut lettuce operation.
Ijms 10 03076f1
Figure 2. Exemplificative interdependent factors that influence pathogens survival and growth on fresh-cut minimal processing [mild heat pre-treatments (MHPT); hot water immersion treatment (HWT); hot water rinsing and brushing (HWRB); oscillating magnetic fields (ohmic heating, dielectric heating, microwaves)]. Combined effects of several antimicrobial strategies is known as “hurdle technology”.
Figure 2. Exemplificative interdependent factors that influence pathogens survival and growth on fresh-cut minimal processing [mild heat pre-treatments (MHPT); hot water immersion treatment (HWT); hot water rinsing and brushing (HWRB); oscillating magnetic fields (ohmic heating, dielectric heating, microwaves)]. Combined effects of several antimicrobial strategies is known as “hurdle technology”.
Ijms 10 03076f2

Share and Cite

MDPI and ACS Style

Capozzi, V.; Fiocco, D.; Amodio, M.L.; Gallone, A.; Spano, G. Bacterial Stressors in Minimally Processed Food. Int. J. Mol. Sci. 2009, 10, 3076-3105. https://doi.org/10.3390/ijms10073076

AMA Style

Capozzi V, Fiocco D, Amodio ML, Gallone A, Spano G. Bacterial Stressors in Minimally Processed Food. International Journal of Molecular Sciences. 2009; 10(7):3076-3105. https://doi.org/10.3390/ijms10073076

Chicago/Turabian Style

Capozzi, Vittorio, Daniela Fiocco, Maria Luisa Amodio, Anna Gallone, and Giuseppe Spano. 2009. "Bacterial Stressors in Minimally Processed Food" International Journal of Molecular Sciences 10, no. 7: 3076-3105. https://doi.org/10.3390/ijms10073076

Article Metrics

Back to TopTop