Skip to main content

REVIEW article

Front. Microbiol., 26 April 2022
Sec. Food Microbiology

Listeria monocytogenes – How This Pathogen Survives in Food-Production Environments?

\r\nJacek Osek
Jacek Osek*Beata LachtaraBeata LachtaraKinga Wieczorek\r\nKinga Wieczorek
  • Department of Hygiene of Food of Animal Origin, National Veterinary Research Institute, Puławy, Poland

The foodborne pathogen Listeria monocytogenes is the causative agent of human listeriosis, a severe disease, especially dangerous for the elderly, pregnant women, and newborns. Although this infection is comparatively rare, it is often associated with a significant mortality rate of 20–30% worldwide. Therefore, this microorganism has an important impact on food safety. L. monocytogenes can adapt, survive and even grow over a wide range of food production environmental stress conditions such as temperatures, low and high pH, high salt concentration, ultraviolet lights, presence of biocides and heavy metals. Furthermore, this bacterium is also able to form biofilm structures on a variety of surfaces in food production environments which makes it difficult to remove and allows it to persist for a long time. This increases the risk of contamination of food production facilities and finally foods. The present review focuses on the key issues related to the molecular mechanisms of the pathogen survival and adaptation to adverse environmental conditions. Knowledge and understanding of the L. monocytogenes adaptation approaches to environmental stress factors will have a significant influence on the development of new, efficient, and cost-effective methods of the pathogen control in the food industry, which is critical to ensure food production safety.

Introduction

Listeria monocytogenes is a Gram-positive facultatively anaerobic microorganism, motile at the temperature range of 22–28°C but non-motile above 30°C, characterized by the growth at temperature range from −0.4°C to 45°C, with an optimum temperature of 37°C (Allerberger, 2003). It is able to survive at a relatively low water activity (aW < 0.90) and a broad pH range between 4.6 and 9.5 as well as to tolerate salt conditions up to 20% (Buchanan et al., 1989; Bucur et al., 2018). These growth conditions made these bacteria able to survive and multiply in adverse environmental conditions which are often present at food production facilities (Gray et al., 2006; Ranasinghe et al., 2021). L. monocytogenes is therefore an important foodborne pathogen responsible the disease called listeriosis, which can appear as sporadic infections or disease outbreaks with significant mortality rate of 20–30% worldwide (Buchanan et al., 2017). Human infection can occur in three forms, depending on the age of the infected person, its immune status, the amount of ingested bacterial cells, and the virulence properties of the strain: severe or mild invasive listeriosis and non-invasive febrile gastroenteritis (Buchanan et al., 2017). Depending on the severity of the illness, symptoms may last from days to several weeks. Mild symptoms may include a fever, muscle aches, nausea, vomiting, and diarrhea. If the more severe form of listeriosis develops, symptoms may include headache, stiff neck, confusion, loss of balance, and convulsions (Buchanan et al., 2017). The disease is especially dangerous for the elderly, pregnant women, unborn babies and immunocompromised people (de Noordhout et al., 2014). According to the recent European Food Safety Authority (EFSA) and European Center for Disease Prevention and Control (ECDC) common report for year 2020, a total of 1,876 confirmed cases of invasive listeriosis in humans were noted in the European Union member states, with the notification rate of 0.42 cases per 100,000 population and 97.1% hospitalizations (EFSA and ECDC, 2021). In the United States, the Centers for Disease Control and Prevention (CDC) estimate that each year about 1,600 persons are infected with L. monocytogenes, with the hospitalization rate of ca. 94% (Scallan et al., 2011).

Listeria Monocytogenes in Food and Food Production Environments

Listeria monocytogenes is a ubiquitous bacterium and has been isolated from soil, water, and feed (Dhama et al., 2015). It has been demonstrated that the bacteria can survive in the environment for at least 8 weeks (Watkins and Sleath, 1981; Rodríguez-Campos et al., 2019). Several investigations have shown that L. monocytogenes is widely distributed in food processing environments where it is able to persist for a long time due to ineffective cleaning and sanitation (Carpentier and Cerf, 2011; Ferreira et al., 2014; Buchanan et al., 2017). Many strains survive in different food processing conditions which are often characterized with a low humidity or oxygen content, and thus becoming a main source of post-processing contamination (Hoelzer et al., 2012; Ferreira et al., 2014; Malley et al., 2015; Bucur et al., 2018). Persistence of such strains may be contributed by several external factors as poor hygiene practice or ineffective sanitizers but also by the presence of the specific genes in some L. monocytogenes strains that are responsible for biofilm production (Nilsson et al., 2011; Harter et al., 2017; Lee et al., 2019; Rodríguez-Campos et al., 2019).

Adverse Environmental Conditions and Survival of L. Monocytogenes

Low Temperatures

As above-mentioned, L. monocytogenes has the ability to grow in a broad range of temperatures (from −0.4°C to 45°C) but also under freezing conditions no significant changes in the live bacteria population was observed (Walker et al., 1990; Gougouli et al., 2008; Nowak et al., 2015). Tolerance to low temperatures resulted in the frequent detection of these bacteria in food products stored under refrigeration conditions (Tasara and Stephan, 2006). The mechanisms of this phenomenon are complex and involve a decrease in the metabolism of the bacterial cells, changes in cell membrane composition, expression of cold shock proteins, and uptake of cryoprotective compounds from the environment (Phadtare et al., 1999; Neunlist et al., 2005; Bucur et al., 2018).

The proper physical condition of the bacterial cell membrane lipids is essential to optimal structural and functional integrity of these membranes (Suutari and Laakso, 1994; Sohlenkamp and Geiger, 2016). Low temperatures affecting the cell lead to reduced membrane lipid fluidity. In response to this stress factor L. monocytogenes changes the membrane lipid composition toward an increase in the concentration of unsaturated fatty acids, which prevents formation of a gel-like state that may result in leakage of cytoplasmic content. It also creates the optimal membrane fluidity for enzyme activity and transport across the membrane (Suutari and Laakso, 1994; Gandhi and Chikindas, 2007; NicAogáin and O’Byrne, 2016; Bucur et al., 2018; Santos et al., 2019). Furthermore, the rate of intracellular enzyme activity decreases to the necessary minimum, and the cell fitness is improved (Mastronicolis et al., 2006; NicAogáin and O’Byrne, 2016). For all these purposes the bacteria modulate their genes expression, especially those involved in cell membrane function and synthesis of lipids, carbohydrates and amino acids as well as those involved in biogenesis and motility (Chan et al., 2007b; Cordero et al., 2016).

During exposure to the cold temperature stress conditions, L. monocytogenes responds in different ways. The cells increase the accumulation of glycine betaine and carnitine from the environment by a chill-activated transport system (Angelidis and Smith, 2003). Both these organic osmolytes are found in high amounts in various foods which may help to promote the survival and growth of L. monocytogenes at lower temperatures (Zeisel et al., 2003). The glycine betaine transporter (Gbu) is an ATP-binding cassette (ABC) transporter that is encoded by the gbu operon (Ko and Smith, 1999), whereas transport of carnitine in response to cold shock is depended on the OpuC ABC transporter, the product of the opuC operon (Fraser et al., 2000). It has been described that both glycine betaine and carnitine were accumulated much faster by L. monocytogenes at 7°C than at 30°C, and their levels increased several times within the cells when grown at 8°C compared to 37°C (Ko et al., 1994; Chan et al., 2007b; Singh et al., 2011).

The role of sigma factor protein σB (SigB) in adaptation to cold stress in L. monocytogenes has been studied (Chan et al., 2007a; Bucur et al., 2018). The protein is stimulated in response to temperature downshift and the sigB-deleted mutant was unable to accumulate solutes such as betaine and carnitine (Chan et al., 2007a; O’Byrne and Karatzas, 2008). It has also been shown that the sigma factor contributed to adaptation in a growth phase-dependent manner, since the absence of SigB protein impaired adaptation of stationary-phase cells to grow at the lower temperature and it was also necessary for efficient accumulation of betaine and carnitine by L. monocytogenes as cryoprotectants (Becker et al., 2000). Furthermore, it has been also demonstrated that only some cold-induced genes were under sigB control (e.g., opuCA gene encoding OpuCA protein with the ATPase-coupled transmembrane transporter activity), whereas other genes responsible for cold shock may be only partially sigB factor-dependent (Chan et al., 2007a; Miladi et al., 2017). On the other hand, Utratna et al. (2014) showed that σB does not play a key role in survival of L. monocytogenes under low temperature stress conditions. These and other authors also demonstrated that sigB is activated at 4°C in a manner that was independent of the levels of RsbV and RsbW proteins encoded by the respective sigB operon genes (Zhang et al., 2013; Utratna et al., 2014). Interestingly, it has been shown that at 4°C there is a significant correlation between the prfA virulence gene regulon responsible for the expression of listeriolysin regulatory protein PrfA and the σB regulon, which moderate the activity of PrfA during the host infection (Ollinger et al., 2009; Heras de las et al., 2011).

Other mechanisms of adaptation of L. monocytogenes to low temperatures have also been described. One such mechanism is the expression of cold shock-domain family proteins (Csps), which are produced mainly at the temperature range from 4°C to 10°C (Bayles et al., 1996; Phadtare et al., 1999; Hébraud and Guzzo, 2000). Csps are structurally related small proteins (65–70 amino acids long), widely distributed among prokaryotes, with a highly conserved structure, that bind to nucleic acids and regulate the expression of various genes including those involved in stress resistance and virulence, cellular aggregation, and motility in L. monocytogenes (Eshwar et al., 2017). Csps stabilize the nucleic acid conformation and act as molecular chaperone that facilitates replication, transcription, and translation at low temperatures (Bucur et al., 2018). Among many Csp proteins identified, CspA contributes to resistance of L. monocytogenes to low temperatures (Schmid et al., 2009; Muchaamba et al., 2021).

Another protein belonging to the cold shock-domain family is a low molecular weight (ca. 18 kDa) ferritin-like protein (Flp), which was detected in much higher levels in L. monocytogenes cultures kept at −20°C and grown at 4°C compared to the bacteria cultured at 37°C (Hébraud and Guzzo, 2000; Miladi et al., 2012). It has been suggested that regulation of Flp synthesis may occur at the transcriptional level since the increase of flp mRNA was detected upon heat and cold shock (Hébraud and Guzzo, 2000).

Pöntinen et al. (2015) described the two-component-system histidine kinases that have been involved in growth of L. monocytogenes at low temperatures and play a role in adaptation to cold stress. Two genes, yycGF and lisRK, responsible for the bacterial adaptation to cold stress conditions, were identified (Pöntinen et al., 2015). The authors suggested that YycF protein encoded by the yycF gene was more involved in the early stage of cells survival, whereas the lisRK product was rather responsible for a longer cold acclimation (Pöntinen et al., 2015).

High Temperatures

Thermal treatment is one of the methods that have been applied in food production and preservation to prevent or limit the growth of pathogenic microorganisms, including L. monocytogenes (Bucur et al., 2018; Ricci et al., 2021). However, the efficacy of high temperature used for inactivation of Listeria during food processing may be limited due to their natural resistance to elevated thermal conditions above 45°C (Arioli et al., 2019). It has been described that the total heat inactivation of L. monocytogenes requires the temperature range from 55°C for 10 min to 65°C for 12 s, respectively (Smelt and Brul, 2014). There are several external factors that have an influence on resistance of L. monocytogenes to heat such as bacterial cells’ age, previous growth and stress conditions, composition of food, strain serotype, etc. (Sörqvist, 1994; Doyle et al., 2001). Cells in the stationary growth phase are generally more resistant to thermal stress than those in log-phase (Doyle et al., 2001). It has been shown that some components present in foods or growth media protect the bacterial cells from heat damage either by stimulation of cellular membrane production or expression of stress-related proteins (Casadei et al., 1998; Juneja et al., 1998). Shen et al. (2014) demonstrated that L. monocytogenes isolates of serotype 1/2a showed a relatively low resistance to elevated temperatures, whereas strains classified to serotypes 1/2b and 4b were more heat-resistant although differences among strains of the same serotypes were also noted.

In response to elevated temperature the bacterial cells show increased production of heat shock proteins (HSPs) (Bucur et al., 2018; Wiktorczyk-Kapischke et al., 2021). In L. monocytogenes three classes of heat shock-associated genes have been identified (Van der Veen et al., 2007; Wiktorczyk-Kapischke et al., 2021). Several genes (grpE, dnaK, dnaJ, groEL, and groES) encode the class I HSPs that act as intracellular chaperones and their expression increases when heat-induced denatured proteins accumulate in the bacterial cytoplasm (Bucur et al., 2018). The class I HSP genes are controlled by the HrcA repressor which negatively regulates expression of this kind of stress response genes (Nair et al., 2000; Hu et al., 2007; Roncarati and Scarlato, 2017). The main role of this class of heat shock proteins is to stabilize and repair partially denatured proteins and to prevent their intracellular aggregation under heat stress conditions (Hendrick and Hartl, 1993; Hartl and Hayer-Hartl, 2002). The class II HSP genes encode general stress proteins whose transcription is dependent on the alternative sigma factor SigB in different growth-inhibiting conditions (Hecker et al., 1996; Nair et al., 2000). Class III heat shock genes, including clpP, clpE, and clpC operons, are negatively regulated by the class III HSP gene regulator CtsR which contains domains that are highly conserved among low GC Gram-positive bacteria, including L. monocytogenes (Karatzas et al., 2003). The ClpC and ClpE proteins possess ATPase activity and are classified to the heat shock protein Clp family of highly conserved molecular chaperones, whereas the serine protease ClpP is a protein possessing proteolytic properties (Schirmer et al., 1996). At increased temperature, McsB kinase, the product of the mcsB gene of the clpC operon, modifies CtsR conformation preventing its binding with gene promoters. As a result, RNA-s32 polymerase binds with promoters leading to gene expression, and CtsR is degraded (Wiktorczyk-Kapischke et al., 2021). It has been shown that ClpC expression is negatively controlled at the transcription level by the cAMP protein receptor PrfA (Heras de las et al., 2011).

Low pH

A low pH environment may be present in food that has undergone acidification, one of the methods of food preservation widely applied to dairy products, meat and vegetables, and is primarily the results of fermentation by bacteria either present in the raw food or added as starter cultures (Hill et al., 2017). Furthermore, L. monocytogenes meets acid conditions in the gastrointestinal tract of the host (NicAogáin and O’Byrne, 2016). The bacteria are able to survive in a low pH of the environment which is generated by artificially induced acidification during acid sanitation (Cotter and Hill, 2003). Low pH increases the concentration of hydrogen protons, which results in the inhibition of microbial growth (Ryan et al., 2008). Furthermore, it has been shown that low pH not only allows the bacteria to survive but also increases its virulence and provides L. monocytogenes higher protection against other adverse environmental conditions (Rodríguez-López et al., 2018).

The bacteria are able to adapt to this low pH environment by means of different mechanisms. Pre-exposure of L. monocytogenes to mild acidic pH of 5.5 for 2 h induces the acid tolerance response (ATR), the process in which the bacteria showed increased resistance to lethal acidic, temperature (52°C), salinity (25–30% NaCl) and alcoholic (15%) shocks (Phan-Thanh et al., 2000). These effects were even more evident when the bacteria adapted to acid gradually (Koutsoumanis et al., 2003). O’Driscoll et al. (1996) also demonstrated that the bacteria exhibited a significant adaptive acid tolerance response following a 1-h exposure to pH 5.5, which is capable of protecting cells from severe acid stress (pH 3.5). It has been also suggested that low pH conditions may have the influence on the selection of L. monocytogenes mutants with increased virulence properties (O’Driscoll et al., 1996).

Listeria monocytogenes uses a variety of metabolic and homeostatic mechanisms to maintain the intracellular pH within a range that is optimal for its growth and survival (Arcari et al., 2020). It is able to increase cytoplasmic buffer capacity through the glutamate decarboxylase (GAD) system or with the action of an internal proton pump (Cotter et al., 2001; Ryan et al., 2008). The GAD mechanism is considered as one of the major mechanisms responsible for the maintenance of the intracellular homeostasis (Cotter et al., 2001). GAD in most L. monocytogenes strains is encoded by five genes, of which three genes (gadD1, gadD2, and gadD3) encode decarboxylases, whereas two other genes (gadT1 and gadT2) are responsible for production of antiporters (Melo et al., 2015). All these five genes are organized in three separate genetic loci: gadD1T1, gadT2D2, and gadD3 (Cotter et al., 2005). The glutamate decarboxylase enzyme promotes the irreversible conversion of cytosolic glutamate to a neutral compound, the γ-aminobutyrate (GABA) (Cotter and Hill, 2003). During the GABA synthesis the intracellular proton level decreases resulted with the subsequent alkalization of the environment and increase of the pH inside of the L. monocytogenes cell. Furthermore, the extracellular GABA excretion leads to the slight neutralization of the pH outside the cell due to the exchange of extracellular glutamate for the more alkaline GABA, and finally the restart of the metabolic pathway (Cotter and Hill, 2003).

Another cell system that protect Gram positive bacteria from low pH is based on the arginine deiminase (ADI) pathway (Soares and Knuckley, 2016). ADI, with the participation of two other enzymes, carbamoyltransferase and carbamate kinase, all encoded by the arcABC operon, converts external arginine to ornithine which is then extracellularly transported in an energy-independent manner by a membrane-bound antiporter encoded by the arcD gene (Ryan et al., 2009). The level of ammonia produced as a byproduct of the system combines with intracellular protons to yield NH4+ maintaining the intracellular level of the cytoplasmic pH, thereby protecting the L. monocytogenes cell from adverse acidic extracellular environments (Cotter and Hill, 2003; Ryan et al., 2009; Matereke and Okoh, 2020). It has been shown that the transcription of arcA and argR genes is both sigma factor protein σB (SigB) and protein receptor PrfA-dependent (Ryan et al., 2009).

In addition to GAD and ADI systems, other proton pumps such as F0F1-ATPase have also been suggested as active mechanisms to maintain L. monocytogenes homeostasis in low (mild) pH environments (Cotter et al., 2000). ATP produced during arginine conversion under ADI mechanisms is used by F0F1-ATPase to generate a proton gradient, enabling H+ expulsion and homeostasis restoration (Smith et al., 2013). The enzyme consists of two distinct domains: the membrane domain F0, which is a channel for proton translocation, and the cytoplasmic domain F1, responsible for catalyzing ATP synthesis and hydrolysis (Smith et al., 2013).

The above-described low pH adaptation systems, i.e., ATR, GAD, and proton extrusion (F1F0-ATPase), act at the same time and ensure survival and adaptation to acid stress conditions of L. monocytogenes (Wiktorczyk-Kapischke et al., 2021).

Listeria monocytogenes also possesses two-component signal transduction system that plays a role in response to environmental stress conditions, including low pH (Cotter and Hill, 2003). This system typically contains two genes, lisR and lisK, encoding cytoplasmic response regulator and a membrane-associated histidine kinase sensor, respectively. The LisRK signal transduction system is able to sense pH changes in the environment by histidine kinase, and the response regulator enables the cell to respond by altering the relevant gene expression (Cotter et al., 1999). It has been shown that a LisRK transposon mutant of L. monocytogenes was more sensitive to low pH than the wild type strain during the logarithmic phase of growth but more acid resistant during stationary phase (Cotter et al., 1999).

High pH

In food production environment there are several alkaline stress factors, which are sublethal for L. monocytogenes and are mainly associated with the use of detergents and disinfectants (Beales, 2004). This bacterium has developed many strategies to withstand adverse high pH-related conditions and as a result they become more cross-resistant to subsequent other, usually more severe, stress factors, such as thermal, alkali, and ethanol stresses or cleaning procedures (Taormina and Beuchat, 2001). Alkali adaptation mechanisms of L. monocytogenes may be significant in the persistence of these pathogenic bacteria in food industry equipment and premises in the presence of alkaline-based detergents used in food processing environments (Giotis et al., 2007; Shen et al., 2016).

Alkali conditions, which may be present in the environment due to the use of detergents and disinfectants, can induce the solubilization of bacterial surface proteins, resulting in exposure of hydrophobic sites of lipids to the extracellular factors (Jacobsohn et al., 1992; Giotis et al., 2009). Such conditions may also directly change the structure of the cell membrane by saponification of membrane lipids or alteration of the membrane fatty acids ratio (Giotis et al., 2007). These changes induce damages that significantly disrupt cell metabolism and structure, preventing effective interactions between bacterial cells and their environment (Almakhlafi et al., 1995).

Generally, to resist alkali damage and maintain cytoplasmic pH at optimal values, L. monocytogenes responds in different ways. One of them is increased metabolic production of intracellular acids through deamination of amino acids and fermentation of sugars (Padan et al., 2005; Giotis et al., 2010). The bacteria are also able to induce transporters and enzymes directly responsible for proton retention and cell surface modifications (Soni et al., 2011). It has been proved that monovalent cation-proton antiporters are essential to maintain a neutral cytoplasmic pH and, therefore, to allow the bacterial growth under alkaline conditions (Gardan et al., 2003a; Giotis et al., 2010). In addition, the acidic cell wall polymers such as teichuronic acid and teichuronopeptides contribute to pH homeostasis, and provide a passive barrier to ion flux and elevation of the cytoplasmic buffering capacity (Krulwich et al., 1997; Gardan et al., 2003a).

A scanning electron microscopy study of L. monocytogenes exposed to sublethal alkaline stress performed by Giotis et al. (2009) showed that the bacterial cells significantly changed their length, radius and volume that may be associated with increased survival of Listeria in such adverse environments. Furthermore, in alkaline conditions, L. monocytogenes cells develop higher proportions of branched-chain fatty acids, including more anteiso forms that are important in adaptation to high pH (Giotis et al., 2007).

Osmotic Shock

Listeria monocytogenes can survive in elevated osmolarity and has the ability to grow even in media supplemented with 12% NaCl and can tolerate adverse salt conditions up to 20% (Bucur et al., 2018). High NaCl concentrations suppress bacterial growth by decreasing water activity in surrounding environment, enhancing plasmolysis and consequently resulting in decreased intracellular turgor pressure and finally, inhibiting the bacterial amplification (Amezaga et al., 1995). In addition to increasing osmotic pressure, NaCl decreases electrochemical potential across the cell membrane, thus, disturbing ATP production by oxidative phosphorylation (Shabala et al., 2006). The response of L. monocytogenes to osmotic stress is called osmoadaptation, a biphasic process consisting of primary and secondary response mechanisms (Hill et al., 2002). The primary phase of adaptation of the bacteria to elevated osmolarity covers physiological changes of the cells, which maintain their turgor by increasing the uptake of potassium ions (K+) and its counterion, glutamate, into the cell, and then replacing part of the accumulated K+ with low-molecular-weight molecules known as compatible solutes or osmolytes in the second stage of osmoadaptation (Sleator et al., 2003). L. monocytogenes possesses two K+ transporters, which play a main role in adaptation to high salt concentration: a high affinity KdpABC transporter system, and a low affinity system encoded by the lmo0993 gene (Kallipolitis and Ingmer, 2001; Brøndsted et al., 2003; Ballal et al., 2007).

Uptake of compatible solutes by L. monocytogenes as a response to elevated osmolarity helps the bacteria to restore turgor pressure and cell volume and stabilize cell protein structure and functions (Kallipolitis and Ingmer, 2001; Brøndsted et al., 2003; Sleator et al., 2003). Several compatible solutes which promote both salt and low temperature tolerances in L. monocytogenes have been identified, including betaine, carnitine, proline, proline betaine, acetylcarnitine, gamma-butyrobetaine, and 3-dimethylsulfoniopropionate (Bayles and Wilkinson, 2000). Among them, betaine has the strongest effect on reduced growth under high osmotic conditions (Beumer et al., 1994). The presence of the osmolytes resulted in an up to 2.6-fold increase in growth rate of salt-stressed L. monocytogenes cells compared to stressed cells without any osmoprotectants (Bayles and Wilkinson, 2000).

It has been described that the main carnitine transport system is encoded by the opuCABCD operon, the glycine betaine by gbuABC, whereas the glycine betaine uptake system depends on the betL gene (Chan et al., 2007b). The expression of genes encoding betaine, carnitine and proline transporters are transcriptionally regulated by general stress sigma factor σB (Sleator et al., 2003; Bae et al., 2012). In more detail, the product of the opuCABCD operon (OpuCA) releases energy during ATP hydrolysis which is needed for transport of the substrate by a complex consisting of the two transmembrane proteins OpuCB and OpuCD as well as a solute binding protein OpuCC (Fraser et al., 2000). The products of the remaining genes (betL and gbuABC) are involved in the primary response to the elevated level of NaCl and toleration of L. monocytogenes to a long-term osmolarity, respectively (Sleator et al., 2003). The strains possessing mutations in these genes showed reduced growth under high osmotic condition (Okada et al., 2008). It has been suggested that compatible solutes may play a dual role in osmoregulation process of the bacterial cells: firstly, they are involved in restoring of cell volume and, secondly, they stabilize protein structure and function under these adverse environmental conditions (Cayley et al., 1992). It has also been shown that the general stress protein Ctc of L. monocytogenes is involved in osmotolerance in the absence of any compatible solutes in the environment (Gardan et al., 2003b).

Several studies indicated that most osmotolerance-associated genes present in L. monocytogenes have also been activated during other stress environmental conditions such as low temperature and low pH generated during artificial food acidification as well as play a role in virulence in a mouse model (Okada et al., 2008). On the other hand, in response to osmotic stress, L. monocytogenes expresses genes other than those associated with osmolyte accumulation, such as csp encoding cold shock-domain family proteins (Csps) (Schmid et al., 2009). These proteins, mainly CspA and CspD, have chaperon activity and facilitate the repair of DNA lesions made by high concentrations of NaCl (Dmitrieva et al., 2004). Apart from stress survival functions, Csp proteins are also involved in cell regulatory networks, playing a crucial role in the regulation of virulence functions of L. monocytogenes, especially those connected with invasion and listeriolysin (LLO) secretion (Loepfe et al., 2010; Schärer et al., 2013).

Bae et al. (2012) described that the presence of osmotic conditions in L. monocytogenes environment decreased expression of genes associated with phosphoenolpyruvate (PEP)-dependent phosphotransferase carbohydrate systems (PTS), including those related to uptake of β-glucoside, galactitol, fructose, and cellobiose. This has an influence on a significantly lower growth rate of the bacteria and reduced uptake of carbohydrates under osmotic stress conditions (Stoll and Goebel, 2010).

High Hydrostatic Pressure

High hydrostatic pressure (HPP) is a food preservation technology used as an alternative to thermal processing. It is widely applied in the meat industry for microbial inactivation of both food spoilage microorganisms and foodborne pathogens, and it is conducted at room temperature, which enhances the safety and shelf life of food (Huang et al., 2014). The pressures applied for sterilization depends on the kind of food and potentially present microorganisms and usually ranges between 250 and 700 MPa (mainly 400 and 600 MPa) for a few seconds to 10 min (Bucur et al., 2018). Primary effects of HPP on bacterial cells are an increase in the permeability of the cell membrane, the disruption of the protein structure and function, and finally, inhibition of the physiological activities of the treated microorganisms (Huang et al., 2014).

It has been reported that HPP causes morphological, structural, physiological, and genetic changes or damages to L. monocytogenes cells (Ferreira et al., 2016). However, several factors influence the resistance of these bacteria to high hydrostatic pressure. Cells in the stationary phase of growth are much more resistant to pressures above 200 MPa than those in the exponential phase (Huang et al., 2014). Furthermore, resistance of L. monocytogenes to HPP depends on the strain and the type, composition and matrix of food products (Bruschi et al., 2017). It has also been shown that higher salt concentrations in food may induce uptake of compatible solutes, which in turn stabilizes cells during HPP (Abe, 2007). It has been shown that pressure-induced damage of the cell membrane has an influence on the Mg2+ leakage from the cell and therefore destabilization of ribosome structure, whereas Ca2+ strengthens the outer membrane of bacterial cells and make the bacteria more resistant to HPP (Niven et al., 1999; Gänzle and Liu, 2015). However, it has been reported that L. monocytogenes treated with HPP up to 550 MPa, resulting in still viable cells, was able to recover and grow during the storage under refrigeration conditions (Bozoglu et al., 2004; Valdramidis et al., 2015).

Several genes and mechanisms that may play a role in recovery from HPP damage of L. monocytogenes have been identified. High pressure processing induces expression of genes associated with DNA repair, transcription and translation protein complexes, cell division, general protein secretion system, flagella assemblage and motility, chemotaxis, and membrane and cell wall biosynthesis pathways (Bowman et al., 2008). On the other hand, HPP suppresses a wide range of energy production and conversion, carbohydrate metabolism and virulence-associated genes (Bucur et al., 2018). An important aspect of the stress-induced survival is induction of the general stress response mediated by the above-mentioned sigma factor protein σB (SigB) which can activate several protective genes under stressful conditions (Wemekamp-Kamphuis et al., 2004; Guerreiro et al., 2020). However, it has also been suggested that HPP may reduce expression of the sigma factor SigB and part of the sigB regulon (Bucur et al., 2018).

It has been shown that HPP also affected genes controlled by the transcription factor CodY, a known global regulator of metabolic genes, including the prfA gene regulon responsible for the expression of listeriolysin regulatory protein PrfA (Lobel et al., 2015). Another gene induced by HPP is cspL encoding a cold-shock protein, which supports the earlier observations that HPP also induces cross-resistance to other stress factors and conditions like heat, acid, and oxidative stress (Karatzas and Bennik, 2002; Bucur et al., 2018). Furthermore, mutations in CtsR, a class III stress genes repressor, have also been connected to spontaneous resistance of L. monocytogenes to HPP (Karatzas et al., 2003). Such mutants, characterized by a stable resistance to HPP, showed point mutations, insertions or deletions in the ctsR gene that down-regulated its activity. This feature was connected with increased expression of the clpB, clpC, clpE, and clpP genes, encoding a large protein complex Clp (caseinolytic protein) with both proteolytic and chaperone activities (Karatzas et al., 2003; Van Boeijen et al., 2010). Clp proteases are able to degradate damaged or denatured proteins forming in the bacterial cells during HPP treatment that are potentially harmful for L. monocytogenes; thus, they increase tolerance of the bacteria to high pressure (Tomoyasu et al., 2001). However, there is also evidence that isolates which did not have such ctsR and other genetic changes still showed resistance to HPP treatment (Karatzas et al., 2005; Chen et al., 2009).

It has been shown that after HPP treatment L. monocytogenes up-regulates the major PEP-PTS, especially fructose-, mannose-, galactitol-, cellobiose-, and ascorbate-specific, involved in the sugars transport (Stoll and Goebel, 2010; Duru et al., 2021). Furthermore, the cell-division-related genes (divIC, dicIVA, ftsE, and ftsX) were also down-regulated, whereas the peptidoglycan-synthesis genes responsible for cell-wall repair (murG, murC, and pbp2A) were upregulated (Duru et al., 2021). Thus, it seems that the bacterial tolerance response to HPP is complex and needs further investigations.

Ultraviolet Light

Pulsed ultraviolet light (PUV) is a non-thermal approach that has a high potential for decontamination of food, water, and air in food production environments (Gómez-López et al., 2007). For this purpose, ultra-short duration pulses of an intense broadband emission spectrum that is rich in UV-C light (200–280 nm band) of the highest germicidal efficacy is used. This UV spectrum mediates bacterial inactivation through several mechanisms, including damage to the bacterial cells in the form of pyrimidine dimers and the loss of cytoplasmic contents post-light absorption (Gómez-López et al., 2007; Rastogi et al., 2010). The main bactericidal effect of UV-C light is caused by DNA damage as a consequence of the formation of photoproducts such as cyclobutane-pyrimidine dimers (CPDs), pyrimidine 6-4 pyrimidone photoproducts (6-4PPs), and their Dewar isomers (Rastogi et al., 2010). Furthermore, UV light also has photophysical and photothermal effects on bacterial cells due to the absorption of the high energy light pulses resulting in leakage of cellular content (Gómez-López et al., 2007). However, during UV treatment, CDPs are the most common lesions in the bacterial genome, which have the principal effect on the cell development and survival, especially that connected with transcription, DNA replication and cell cycle progression (Beauchamp and Lacroix, 2012).

The efficacy of UV treatment in decontamination of food surfaces depends on many factors, such as the kind of food, distance of the product to the light source, energy level given by number and frequency of the light pulses, time applied for the UV treatment, level of contamination, and others (Gómez-López et al., 2007). It has been shown that L. monocytogenes is more resistant to UV-C light than other bacterial pathogens, such as E. coli (Beauchamp and Lacroix, 2012). The presence of organic materials such as food debris on stainless steel surfaces and NaCl content in ultraviolet treated foods have an influence on the efficacy of UV-C radiation on L. monocytogenes due to low ability of this light to penetrate organic substances (Bernbom et al., 2011). Other studies have also shown that the presence of salt in brine increases the amount of UV-C necessary for inactivation of L. monocytogenes in fluid (McKinney et al., 2009b). Similar increased resistance to ultraviolet light was observed for the bacterial cells grown in vitro in a low pH, presence of antibiotics, or disinfectants (McKinney et al., 2009a; Naitali et al., 2009).

There is very little information related to detailed molecular mechanisms of UV resistance in L. monocytogenes. It seems that UV-C resistance is not general stress sigma factor protein σB-dependent (Gayán et al., 2015). Another study of Uesugi et al. (2016) revealed that overall changes in gene expression resulting from ultraviolet treatment of L. monocytogenes 10403S strain were low. However, a number of genes encoding stress proteins, motility and transcriptional regulators were up-regulated by the UV exposure, although no induction of the lmo0588 gene, responsible for deoxyribodipyridimine photolyse activity induced by light was observed (Uesugi et al., 2016). Photolyase, the product of the lmo0588 gene, plays an important role in photoreactivation, i.e., the recovery of bacteria sublethally injured by UV light due to subsequent exposure of visible light (Gómez-López et al., 2007). It has been shown that, during photoreactivation, photolyase binds and repairs the pyrimidine DNA lesions using light energy absorbed by its chromophores (Sinha and Häder, 2002).

It has also been described that, among the putative stress response genes located on plasmids of L. monocytogenes, the uvrX gene being a part of the Y-family DNA polymerase, plays a role in response of the cells exposed to UV stress (Naditz et al., 2019; Cortes et al., 2020). A similar finding was described for the chromosomally encoded gene uvrA, which was shown to be necessary for the bacterial UV stress survival (Kim et al., 2006). Recently, Anast and Schmitz-Esser (2021) identified several L. monocytogenes plasmids that confer increased UV stress tolerance, although their precise role in this phenomenon needs further investigations.

Heavy Metals

Certain heavy metals such as copper, zinc, and iron in trace amounts are essential for bacterial survival and play a role of co-factors for a broad range of L. monocytogenes cellular proteins but the same metals at higher concentrations often become toxic (Jesse et al., 2014; Parsons et al., 2019). Other heavy metals (e.g., arsenic and cadmium) are probably not necessary for cellular functions and seem to be toxic in any concentration (Jesse et al., 2014; Parsons et al., 2019). However, L. monocytogenes possesses various mechanisms to maintain its cellular heavy metal homeostasis, avoid poisoning, and thus, survive in diverse environmental niches (McLaughlin et al., 2011; Jesse et al., 2014). In this section, resistance to the main toxic heavy metals, i.e., cadmium and arsenic, is discussed.

Cadmium Resistance

It has been reported that approximately 50% or more of tested L. monocytogenes isolates from foods and food processing plants were resistant to cadmium (Ratani et al., 2012). On the other hand, there is also information that most strains were susceptible to this metal at concentration of 64 μg/ml cadmium sulfate (Margolles et al., 2001). Resistance to cadmium is encoded by different genetic determinants often located on mobile genetic elements (mainly plasmids) (Parsons et al., 2019). At least five molecular mechanisms have been identified that contribute to cadmium resistance, although resistant strains lacking these genetic determinants have also been identified suggesting that there are yet unidentified means of metal tolerance (Lee et al., 2013). The first cadmium molecular resistance gene described in L. monocytogenes was the cadA1 located on plasmid-borne Tn5422 transposon, which is responsible for the activity of efflux P-type ATPase pumps (Lebrun et al., 1994). Few strains harbor Tn5422-associated cadA1 chromosomally (Hingston et al., 2019). A second putative cadmium resistance sequence, cadA2, was initially detected on the large plasmid pLI100 of L. innocua and then discovered on the approximately 80 kb plasmid pLM80 of L. monocytogenes H7858, a strain implicated in a large listeriosis outbreak in the United States in 1998–1999 (Nelson et al., 2004). It was also found that pLM80-associated cadAC (cadA2) is part of a putative composite transposon that also harbors genes for resistance to benzalkonium chloride (Elhanafi et al., 2010). In addition, pLM80 determines resistance to toxic triphenylmethane dyes such as crystal violet and malachite green via the tmr gene (Dutta et al., 2014). The third cadmium resistance gene, cadA3, is carried on the integrative and conjugative element (ICE) of L. monocytogenes EGD-e (Elhanafi et al., 2010). Finally, the cadA4 and cadA5 are on the large chromosomally located Listeria Genomic Island 2 (LGI2) and LGI2-1 islands, respectively (Parsons et al., 2020). The cadA3, cadA4, and cadA5 genes have so far been identified only on chromosome (Kuenne et al., 2010; Lee et al., 2017; Parsons et al., 2019). It has also been found that LGI2 harbors arsenic resistance cassette comprising of arsD1A1R1D2R2A2B1B2 genes (Lee et al., 2013). Comparison of the cadA1-cadA4 sequences revealed that the first three conferred a high level of resistance to cadmium (MIC > 140 μg/ml), whereas cadA4 was responsible for relatively lower resistance levels (MIC < 70 μg/ml) (Lee et al., 2013). Interestingly, LGI2 genetic element has been mainly connected with L. monocytogenes classified into hypervirulent clones of clonal complexes CC1 and CC2 as well as CC4 (Lee et al., 2017). Furthermore, strains with multiple cadA variant determinants have been identified (Lee et al., 2013). On the other hand, several L. monocytogenes isolates lacking any of the four cadmium resistance determinants were detected, which suggests the presence of one or more unidentified cadmium resistance genes (Ratani et al., 2012; Lee et al., 2013).

It has been described that the prevalence of the known cadmium resistance molecular determinants was serotype-related: cadA1 was more common in L. monocytogenes isolates of serotypes 1/2a and 1/2b than 4b from food and food-processing environment, while cadA2 was mainly associated with strains of serotype 4b (Ratani et al., 2012; Lachtara et al., 2021). However, several cadmium-resistant isolates lacking the known cadA determinants were classified to serotype 4b, which suggests that such strains may possess other than cadA-encoded resistance mechanisms (Ratani et al., 2012; Chmielowska et al., 2021). Furthermore, another study showed that cadmium resistance was more common among persistent L. monocytogenes strains, i.e., those repeatedly isolated from foods than among those recovered sporadically (Harvey and Gilmour, 2001).

Arsenic Resistance

One of the most important mechanisms responsible for arsenic resistance in L. monocytogenes appears to be encoded by genes carried on the above-mentioned LGI2, primarily associated with strains of serotype 4b, especially of the hypervirulent clones CC1, CC2, and CC4 (Lee et al., 2017). Molecular analysis of arsenic-resistant isolates harboring LGI2 revealed that this island was inserted in at least eight different locations, primarily within open reading frames (Lee et al., 2017). Strains classified to serotypes 1/2a, 1/2b, and 1/2c rarely carried the LGI2 insert, thus, are usually arsenic-sensitive (Hingston et al., 2019). It has been shown that the arsenic resistance cassettes contain three (arsRBC) to five (arsRDABC) genes (Rosen, 1999; Kuenne et al., 2010). The products of the arsA (encoding ATPase) and arsB (responsible for a membrane transporter) genes are an ATP-dependent anion pump that exports arsenite out of the bacterial cells (Tisa and Rosen, 1990). Furthermore, the arsA gene product is also able to act independently as a passive transporter of arsenite (Rosen, 1999). The arsC gene encodes a reductase that is responsible for the conversion of arsenate to arsenite, which is then extruded by ArsA or the ArsA/ArsB complex (Tisa and Rosen, 1990). The arsA, arsB, and arsC genes are regulated by two regulatory proteins encoded by arsR and arsD (Wu and Rosen, 1993). Interestingly, the LGI2 insert present in L. monocytogenes harbors several additional genes, including the putative cadmium resistance determinant cadA4 mentioned above (Parsons et al., 2019).

Biocides

Resistance to Quaternary Ammonium Compounds

Increased tolerance of L. monocytogenes to disinfectants (biocides) has been deeply studied in relation to quaternary ammonium compounds (QACs), such as benzalkonium chloride (BC), widely applied in food production and health care environments as well as in households due to their effectiveness, low toxicity, and non-corrosive properties (Gerba, 2015; Kode et al., 2021). QACs are usually used in concentrations ranging from 200 ppm to 400 ppm on food-contact surfaces; however, some formulations also have 1,000 ppm concentrations (Aryal and Muriana, 2019). The antimicrobial effectiveness of QACs and other biocides may be affected by the surface structure of the food production equipment to which bacteria are attached, an uneven distribution of disinfectants on the decontaminated surfaces, too high dilution of disinfectants due to presence of water on the equipment surfaces, or the presence of organic pollutants resulting from insufficient cleaning before disinfection (Duze et al., 2021). Biofilms, often produced by L. monocytogenes, also have an impact on resistance of these bacteria against QAC disinfectants (Duze et al., 2021). Most QACs are aerobically biodegradable and their concentrations in the environment varied resulting in the formation of concentration gradients (Tezel and Pavlostathis, 2015). This may lead to the generation of selective pressure for adaptation or acquisition of resistance genes of initially susceptible L. monocytogenes strains (Tezel and Pavlostathis, 2015; Conficoni et al., 2016).

Quaternary ammonium compounds are active agents that interact with the cytoplasmic membrane of bacteria, including L. monocytogenes, and also able to interact with intracellular targets as well as to bind to DNA (Zinchenko et al., 2004). It has been shown that at low concentrations (0.5 to 5 mg/L) they are characterized by bacteriostatic properties whereas, at concentrations of 10–50 mg/L, QACs are bacteriocidal for the same bacteria, depending upon the formulation (Gerba, 2015). However, there are studies showing that several L. monocytogenes isolates were resistant to BC at a concentration of 1,000 mg/L after 24 h of exposure time (Mohamed et al., 2018).

The action of these biocides involves absorption of QACs by the bacterial cell and penetration of the cell wall, reaction with the cytoplasmic membrane and its disruption, leakage of intracellular components, degradation of proteins and DNA, and finally lysis of bacterial cell wall by autolytic enzymes (McDonnell, 2017). It has been shown that the effectiveness of QACs seems to be not different between persistent and non-persistent L. monocytogenes strains isolated from food production environments (Ruckerl et al., 2014; Magalhães et al., 2016). However, there are also reports showing that persistent isolates from food processing plants and ecosystems exhibited higher resistance to QACs than non-persistent ones, especially strains classified to serotype 1/2c (Ortiz et al., 2016; Meier et al., 2017). Studies of Haubert et al. (2019) demonstrated that all 50 L. monocytogenes isolates from food tested were resistant to benzalkonium chloride, and more than 50% of these tolerant strains displayed cross-resistance to cadmium. Such correlation between BC and cadmium resistances has been previously reported by other authors, especially among L. monocytogenes of serotypes 1/2a and 1/2b (Mullapudi et al., 2008; Ratani et al., 2012).

It has been observed that exposure of L. monocytogenes to increasing concentrations of BC resulted in adaptation to higher levels of this and other biocides (Yu et al., 2018; Noll et al., 2020). These studies demonstrated that repeated exposure to subinhibitory concentrations of QACs and prolonged environmental persistence of tolerant strains may facilitate the development of bacterial resistance to these biocides over time. Thus, selection of BC-tolerant bacteria increases the ability of L. monocytogenes to survive under treatment with higher concentrations of the same biocide (Ortiz et al., 2014; Tezel and Pavlostathis, 2015). Consequently, this process further contributes to the persistence of L. monocytogenes in the food processing environments (Møretrø et al., 2017).

The major molecular mechanisms of quaternary ammonium compounds resistance in L. monocytogenes involve several efflux pump systems, including the three-gene cassette bcrABC associated with BC tolerance (Noll et al., 2020). These cover two endogenous multidrug efflux pump genes (multidrug resistant Listeria, mdrL, and Listeria drug efflux, lde) of the major facilitator superfamily and efflux pump genes (bcrABC cassette, qacH, emrE, and emrC) located on mobile genetic elements (Duze et al., 2021). It has been shown that the two major efflux pump genes, mdrL and lde, have been identified in almost all L. monocytogenes serotypes, and enhanced expression of these two endogenous efflux pumps, especially MdrL, resulted in BC resistance (Martínez-Suárez et al., 2016; Yu et al., 2018; Haubert et al., 2019; Jiang et al., 2019). However, the main role of these gene products is detoxification of macrolides, cefotaxime, heavy metals, and ethidium bromide (mdrL), and fluoroquinolones, ethidium bromide, and acridine orange (lde), respectively (Mata et al., 2000).

Among the four efflux pump genes located on mobile genetic elements, the bcrABC cassette was firstly identified in L. monocytogenes responsible for the multistate listeriosis outbreaks in 1998–1999 in the United States (Elhanafi et al., 2010). In most isolates, this cassette is located on the pLM80 plasmid but has also been identified on chromosome (Dutta et al., 2013). These authors have also demonstrated that, in BAC-tolerant L. monocytogenes from various sources, the bcrABC cassette was present in 98.6% of isolates (Dutta et al., 2013).

The qacH efflux pump gene is located on a chromosomally integrated Tn6188 transposon of 5,117 bp in size and consists of three transposase genes (tnpABC) as well as genes encoding a putative transcriptional regulator and QacH, a small multidrug resistance protein family (SMR) transporter associated with export of BC in bacteria (Müller et al., 2014). The significant expression of qacH-encoded efflux pumps has been shown in the presence of benzalkonium chloride and the qacH deletion mutants had lower tolerance to BC than wild type strains (Müller et al., 2014). It has also been described that QacH protein confers higher tolerance to other QACs and ethidium bromide (Müller et al., 2014). A study of Meier et al. (2017) demonstrated that the majority of Swiss and Finnish L. monocytogenes 1/2c clinical and food isolates resistant to BC were qacH-positive, although a subset of BC-resistant strains lacked genes for efflux pumps currently known to confer BC resistance. Similar observations were described by other authors (Ortiz et al., 2014; Ebner et al., 2015; Møretrø et al., 2017).

Other efflux pump genes responsible for the increased tolerance of L. monocytogenes to QAC are emrE and emrC sequences, located on the LGI1 genomic mobile island and pLMST6 plasmid, respectively (Kovacevic et al., 2016; Kremer et al., 2017). The emrE gene was first described in a study on L. monocytogenes isolates responsible for the deadliest listeriosis outbreak in Canada in 2008 (Kovacevic et al., 2016). During this investigation it was found that strains possessing the LGI1 island with the emrE sequence was characterized by a significantly improved bacterial growth in the presence of QACs compared to the adaptation and growth of genetically similar strains but lacking LGI1. Furthermore, the expression of emrE and several other genes on the LGI1 genomic island was induced in the presence of BC, whereas deletion of the emrE gene resulted in reduced bacterial growth and survival in the presence of QACs (Kovacevic et al., 2016).

The emrC QAC resistance gene, carried by plasmid pLMST6, was identified in L. monocytogenes of sequence type ST6 strains, isolated from adults suffering from listeriosis with meningitis (Kremer et al., 2017). Interestingly, Kropac et al. (2019) demonstrated that the plasmid pLMST6 was not associated with increased tolerance to benzalkonium chloride, but rather increased tolerance to other types of QAC-based biocides. Furthermore, pLMST6 plasmid had no impact on the sensitivity of L. monocytogenes to non-QAC disinfectants or on resistance of isolates to ampicillin, tetracycline and gentamicin (Kropac et al., 2019).

Resistance to Other Biocides

Chlorine-based disinfectants such as sodium hypochlorite, chlorine dioxide gas or aqueous chlorine dioxide are used in food industry to control L. monocytogenes contamination (Vaid et al., 2010). These chemicals possess fast and strong oxidizing properties and interact with bacterial cell wall membranes, mainly phospholipids, or penetrate directly into the cell wall where they form N-chloro groups that react with the bacterial metabolism due to the interference with key enzymes (Wei et al., 1985; Denyer and Stewart, 1988). The efficacy of chlorine-based disinfectants seems to be bacterial cell age-dependent since younger cultures (24 h) are more resistant than older ones (48 h) (El-Kest and Marth, 1988). Furthermore, in L. monocytogenes biofilms the efficacy of chlorine solutions depends on the material on which the biofilm is formed, e.g., bacteria are more easily destroyed when grown on stainless steel surfaces compared to those grown on polyvinyl chloride or Teflon surfaces (Bremer et al., 2002; Pan et al., 2006).

The effect of chlorine-based disinfectants against L. monocytogenes also depends on the chemical compounds used. It has been shown that chlorine dioxide is less toxic, more effective at low concentrations and needs a shorter reaction time than chlorine alone (Chang et al., 2000). One of the main disadvantages of chlorine-based biocides is the formation of toxic disinfection-by-products, especially when they are dissolved in water containing organic matter which is often the case in food production environments (Vaid et al., 2010; Duze et al., 2021). Such products, including trihalomethanes and haloacetic acids, are potential carcinogens and have been associated with various health problems (Chang et al., 2000; Rand et al., 2007).

Acid compounds, like chlorine-based disinfectants, are strong oxidizers and have an effective antibacterial properties (Chang et al., 2000; Vaid et al., 2010). They interfere with cellular phospholipids and cytosolic intracellular material causing irreversible damage (e.g., disruption of proton motive force) and subsequent cells death (Denyer and Stewart, 1988).

Biofilms

Listeria monocytogenes is able to attach to a variety of surfaces in food production environments, including stainless steel, polystyrene or glass, and then to form biofilms (Rodríguez-López et al., 2018). This is a serious concern for food safety because biofilm-contaminated food environments may serve as source of pathogenic bacteria for food products and finally for consumers (Colagiorgi et al., 2017). Bacterial cells in biofilms are embedded in a self-produced matrix of extracellular material, composed of extracellular DNA, proteins, polysaccharides, and other inorganic molecules, called extracellular component matrix (ECM) (Colagiorgi et al., 2017). In the L. monocytogenes biofilm matrix, various extracellular polymeric substances (EPSs) have been identified, with different polysaccharides (mainly teichoic acid), proteins, and extracellular DNA (Colagiorgi et al., 2016). It has been shown that L. monocytogenes biofilms are strongly influenced by temperature, bacterial strain, incubation time, medium, and the nature of the adhesion surface (Borucki et al., 2003; Midelet et al., 2006; Harvey et al., 2007; Tresse et al., 2007; Di Bonaventura et al., 2008; Mazaheri et al., 2021). Di Bonaventura et al. (2008) and Tomičić et al. (2016) observed that L. monocytogenes was able to form biofilms at 4 and 12°C with higher levels on glass compared to the more hydrophobic stainless steel and polystyrene. Furthermore, in both cases, the production of biofilms was significantly higher at 37°C than at 4°C. These authors suggested that these results were not due to a different cellular physiology but rather to a reduced growth of bacteria (Di Bonaventura et al., 2008; Tomičić et al., 2016). On the other hand, Bonsaglia et al. (2014) observed biofilm formation at 4°C on different surfaces, with higher levels of biofilm on stainless steel and glass compared to polystyrene. Similar data were presented earlier by Norwood and Gilmour (2001) who showed that some L. monocytogenes strains were able to adhere in the same way at 4°C and 30°C. Other studies suggested that cold-adapted L. monocytogenes, stored at −20°C for 6 and 24 months, was characterized by increased adhesion and biofilm formation on various abiotic surfaces (Slama et al., 2012).

It has been shown that flagella-mediated motility plays a key role in both initial surface attachment and subsequent biofilm formation by L. monocytogenes and the flaA mutants displayed reduced colonization ability (Lemon et al., 2007; Todhanakasem and Young, 2008; Gorski et al., 2009; Doghri et al., 2021; Mazaheri et al., 2021). Since temperature regulates flagella expression in L. monocytogenes, it is clear that this factor has a strong influence on biofilm formation (Todhanakasem and Young, 2008). It has been demonstrated that this pathogen is flagellated and motile at temperatures < 30°C, and not at all or much less flagellated and motile at temperatures above 30°C (Gründling et al., 2004). However, L. monocytogenes is also able to attach to abiotic surfaces through a flagella-independent binding process, which is not related to a temperature (Tresse et al., 2009).

A correlation between L. monocytogenes serotypes or clones and biofilm formation has been investigated but no clear dependence was detected (Borucki et al., 2003; Harvey et al., 2007). Doijad et al. (2015) tested the ability of 98 clinical and food isolates classified to serotypes 1/2a, 1/2b, and 4b to form a biofilm. Most of the strains (63.3%) were classified as weak biofilm producers, whereas the remaining isolates were defined as moderate and strong (9.2% of each) biofilm formers (Doijad et al., 2015). Interestingly, none of the strains of 4b serotype exhibited strong biofilm formation. It has been also shown that strong biofilm-forming isolates developed biofilm structures within 24 h on surfaces important in food industries such as stainless steel, ceramic tiles, high-density polyethylene plastics, polyvinyl chloride pipes, and glass (Doijad et al., 2015). Using whole-genome sequencing data from 166 environmental and food-related L. monocytogenes biofilm-forming isolates, it has been suggested that serotype-specific differences in biofilm development can be linked to the presence of stress survival islet 1 (SSI-1) (Keeney et al., 2018). In this study, strains of serotype 1/2b, the majority of which contained SSI-1, formed the strongest biofilms, while isolates classified to serotype 4b, which only some of them were SSI-1-positive, were the weakest biofilms producers.

Investigations performed by Takahashi et al. (2009) on 71 L. monocytogenes of food origin revealed a significant correlation between isolates of lineage I (serotypes l/2b and 4b) but not strains of lineage II (serotypes 1/2a and l/2c) and biofilm formation. However, it was also found that isolates classified to the same clonal lineage produced different levels of biofilms, which may suggest that environmental factors are involved in this process (Carpentier and Chassaing, 2004). On the other hand, Borucki et al. (2003) found a higher biofilm-forming ability for L. monocytogenes isolates of lineage II. A reason for these different results obtained by several authors may be caused by the differences in methods applied and the various strains used for the study.

Significant differences in gene expression between biofilm-forming and planktonic L. monocytogenes bacterial cells, especially those involved in expression of internalins (InlA and InlC) and listeriolysin O (LLO), have been observed (Lourenço et al., 2013; Mata et al., 2015; Gilmartin et al., 2016). Isolates with mutations in the inlA gene, which resulted in the reduced length of InlA protein, demonstrated enhanced biofilm forming abilities but a lower virulence potential compared to the strains possessing full-length InlA (Franciosa et al., 2009). Such mutations occur more commonly among food isolates than in strains responsible for human infections (Nightingale et al., 2005).

Lemon et al. (2010) found that prfA, the transcriptional activator of virulence genes, promotes biofilm formation in L. monocytogenes. Although in prfA negative mutants the flagellar motility remains intact and the cells are able to attach to abiotic surfaces, they are defective in next stages of biofilm formation (Lemon et al., 2010).

Schwab et al. (2005) investigated the role of the alternative stress sigma factor σB in biofilm formation and revealed that initial attachment of both wild type and mutant L. monocytogenes to the stainless steel surface was the same, but the number of sigB-deficient strain on the surface was significantly lower than the wild type after 48 h or 72 h of incubation.

Recently, Fan et al. (2020) studied a role of the two-component chemotactic system encoded by the cheA/cheY genes, located immediately downstream of the flaA flagellin gene. The cheY knockout mutant showed decreased biofilm formation ability along with reduced cell-surface hydrophobicity compared to wild type strain. Similar results were also obtained by Li et al. (2021) who showed that cheA and cheY are key genes in the formation of L. monocytogenes aggregates in vitro.

The agrBDCA operon present in L. monocytogenes consists of genes that code for AgrD, an auto-inducing peptide, AgrB, a protein involved in processing the peptide, AgrC, a two-component histidine kinase, and AgrA, a response regulator (Miller and Bassler, 2001). It has been shown that agr system, that is involved in quorum sensing, has a strong influence on biofilm formation as mutations in agrA and agrD display reduction in their ability to form biofilms compared to the wild type strains under both static and dynamic conditions (Rieu et al., 2007; Riedel et al., 2009; Zetzmann et al., 2016). Pieta et al. (2014) studied the presence and expression of the agrA gene in L. monocytogenes of serotypes 1/2a and 4b, grown at 7°C and 37°C. The authors found that the gene was not detected in strains of serotype 4b, and its transcription level in strains of serotype 1/2a was lower at 7°C compared to 37°C.

Alonso et al. (2014) identified 38 genetic loci possibly involved in L. monocytogenes biofilm formation when grown at 35°C. Among them, the D-alanylation pathway genes dltABCD and the phosphate-sensing two component system phoPR were important in this process since the deletion mutants showed decreased ability to produce biofilms. It may suggest that D-alanylation of lipoteichoic acids mediated by the gene products of the dltABCD operon and the phosphate-sensing phoPR system play a significant role for L. monocytogenes to form biofilms.

A role of autoinducer (AI-2) molecules and the luxS gene in quorum sensing and biofilm production by L. monocytogenes was tested by Garmyn et al. (2009). The authors revealed that S-ribosylhomocysteinase encoded by luxS catalyzes the hydrolysis of S-ribosylhomocysteine to homocysteine and 4,5-dihydroxy-2,3-pentadione, precursor molecules of AI-2, and thus are involved in biofilm formation (Bonsaglia et al., 2014). Mutation in luxS led to production of a denser biofilm and better attachment by a luxS-deficient mutant to a glass surface compared to the wild type strain (Belval et al., 2006; Sela et al., 2006). Furthermore, the culture supernatants of luxS mutants were shown to accumulate S-adenosyl homocysteine and S-ribosyl homocysteine, the AI-2 precursor molecules (Belval et al., 2006).

Resistance and Persistence

Persistent L. monocytogenes strains have been defined as isolates (clones) that are repeatedly cultured from the same source or ecological niche over time (Palaiodimou et al., 2021; Unrath et al., 2021). Such isolates have indistinguishable molecular background as tested by genome-based approaches, e.g., pulsed-field gel electrophoresis (PFGE) or recently, next generation sequencing (NGS) (Fox et al., 2011; Brown et al., 2021; Unrath et al., 2021). Persistence of L. monocytogenes is due to different characteristics of such isolates, e.g., tolerance to sanitizers, ability to grow at low temperatures, resistance to heavy metals, or ability to develop biofilm (Kathariou, 2002; Gandhi and Chikindas, 2007; Carpentier and Cerf, 2011; Palaiodimou et al., 2021). Persistent isolates present an important challenge to food producers, as they are associated with cross-contamination of food products because they are hardly or not at all eliminated from food production environments (Palaiodimou et al., 2021).

Several genetic determinants have been suggested to play a role in persistence of L. monocytogenes; however, the nature of the role of these mechanisms to the persistence phenomenon remains still poorly understood (Palaiodimou et al., 2021). This may be due to difficulties in creating appropriate conditions under in vitro studies that accurately reflect the natural environment in food production facilities. Overall, there is a lack of conclusive evidence on whether persistent strains are more resistant to particular stress conditions compared to sporadic strains from similar sources (Taylor and Stasiewicz, 2019).

Persistent L. monocytogenes strains have been isolated from food processing environments after cleaning and disinfection (Lundén et al., 2003; Soumet et al., 2005). The relationship between resistance to various biocides and persistence of certain subtypes of L. monocytogenes in different food processing environments has been studied but no clear correlation was identified (Heir et al., 2004; Kastbjerg and Gram, 2009; Ferreira et al., 2014). On the other hand, there are investigations that showed that there is a link between resistance to benzalkonium chloride and persistence of some strains, especially those positive for the bcrABC gene cassette (Elhanafi et al., 2010; Martínez-Suárez et al., 2016; Ortiz et al., 2016; Cherifi et al., 2018; Cooper et al., 2021).

A correlation between biofilm formation and persistence of L. monocytogenes in the food production environments has been investigated by several authors (Lundén et al., 2000; Borucki et al., 2003; Colagiorgi et al., 2017; Rodríguez-López et al., 2018; Lee et al., 2019; Lianou et al., 2020; Fagerlund et al., 2021; Mazaheri et al., 2021; Unrath et al., 2021). Generally, persistent strains have shown increased biofilm formation in relation to non-persistent strains (Borucki et al., 2003). It has been also observed that biofilms produced on stainless steel surfaces by persistent strains are thicker than those formed by strains found only sporadically (Lundén et al., 2000). Norwood and Gilmour (2001) tested the adherence capability to stainless steel surface of two L. monocytogenes strains with and without persistent ability. It was shown that mean counts of adherent cells over a 24-h period at 25°C were significantly higher for persistent strains (Norwood and Gilmour, 2001). Similar observations were noted by Lundén et al. (2000) and Borucki et al. (2003). Further studies confirmed that persistent L. monocytogenes genotypes were often associated with higher survival and biofilm formation capacity in the presence of sublethal concentrations of benzalkonium chloride (Maury et al., 2019). On the other hand, other authors have shown that there were no clear associations between biofilm formation efficiency and persistent or prevalent genotypes (Djordjevic et al., 2002; Harvey et al., 2007; Jensen et al., 2007; Lee et al., 2019).

In a study of Wen et al. (2011) persistent L. monocytogenes of serotype 4b were shown to be extremely resistant to high temperatures and pressure stresses. Resistance to cadmium has been more often noted among persistent clones compared with their sporadically contaminating counterparts (Harvey and Gilmour, 2001; Parsons et al., 2020). On the other hand, Palaiodimou et al. (2021) have shown that high frequencies of known cadmium resistance cassettes were almost equally present among both persistent (86%) and presumed non-persistent (83%) L. monocytogenes populations. However, their results suggest that the cadA1 gene was more common in persisters, whereas the cadA4 sequence, which provides lower tolerance to cadmium than cadA1, was only carried in non-persistent isolates (Palaiodimou et al., 2021).

Two L. monocytogenes stress survival islets (SSIs) which provide benefits to growth and/or survival under suboptimal or stress conditions, such as low pH (SSI-1), alkaline pH (SSI-2) or oxidative stress conditions (both islets), are usually overexpressed among persistent populations identified in food production environments (Ryan et al., 2010; Harter et al., 2017; Palaiodimou et al., 2021). Interestingly, it has been suggested that persistent L. monocytogenes may possess a lower virulent potential due to the presence of truncated inlA gene, frequent lack of additional virulence factors such as LIPI-3 and LIPI-4 and the mutations in the prfA gene (Ortiz et al., 2016; Palaiodimou et al., 2021).

Taylor and Stasiewicz (2019) compared the influence of different stress conditions (10% of NaCl; different concentrations of benzalkonium chloride; and energy sources) on the growth of persistent and sporadic L. monocytogenes strains of food origin and confirmed observations of other authors that there was not a significant difference in growth rate or ability to grow for isolates of persistent strains compared to sporadic strains for any treatments at 37°C.

Novel L. Monocytogenes Control Strategies

Several strategies for L. monocytogenes elimination from food chain have been developed and applied in the food industry (Khan et al., 2016; Rothrock et al., 2019; Zhang et al., 2021). One of them is irradiation processing technology with gamma irradiation. The approach has been shown to be safe and is a proven method used worldwide for food product preservation (Lacroix and Ouattara, 2000; Maherani et al., 2016). Food irradiation involves exposing food to gamma radiation to induce the demise of L. monocytogenes and other bacteria that can cause food poisoning or food spoilage (Lacroix and Follett, 2015).

Another method of elimination of the pathogens from food and food production environments is application of ozone, the eco-focused method which is categorized as generally recognized as safe (GRAS) (Panebianco et al., 2021). It has been recently shown application of gaseous ozone at 50 ppm on planktonic cells and biofilm of reference and food-related L. monocytogenes strains resulted in over 3 log10 CFU/ml reduction of bacterial load after 10 min (Panebianco et al., 2021). Furthermore, a complete inactivation of planktonic cells after 6 h of treatment as well as a significant decrease of the biofilm biomass were observed. Thus, the use of gaseous ozone is a promising method of L. monocytogenes contamination control on both food contact surfaces and on the final products (Botta et al., 2020).

One novel alternative biological method of L. monocytogenes control along the food chain is the use of phages (Kawacka et al., 2020). Phages are considered an effective tool against bacterial pathogens as they only target their specific organism and do not interfere with other microorganisms. This is especially important in the production of fermented foods as they do not have a negative influence on the sensory properties of the final product (Sadekuzzaman et al., 2017; Moye et al., 2018). There are commercial phage-based products which are successfully applied in food industry to control L. monocytogenes. One of them is ListShield™ (Intralytics, Columbia, MD, United States), a cocktail of six different lytic bacteriophages that is specifically designed for treating foods that are high risk for L. monocytogenes contamination, like ready-to-eat meat (RTE) products (Perera et al., 2015). It has been shown that ListShield™ significantly reduced by 82–99% the number of L. monocytogenes in different kinds of RTE food (Perera et al., 2015). In the case of smoked salmon, the phages completely eliminated the pathogen in both the naturally contaminated and experimentally contaminated samples without affecting the organoleptic quality of the food. ListShield™ can also be used to eliminate or significantly reduce the levels of L. monocytogenes on non-food contact equipment, surfaces, etc., in food processing plants and other food establishments (Ishaq et al., 2020).

LISTEXTM P100 (Micreos, Hague, Netherlands) is also approved by the FDA and recommended by EFSA as a phage cocktail product for the reduction of L. monocytogenes on meat and poultry foods during processing or in the final product (EFSA, 2016). The effectiveness of the broad host range bacteriophage P100 present in this products was tested for the reduction of L. monocytogenes in inoculated samples at different temperatures and the maximum decrease of the number of the pathogen was achieved at the level of 4.44 log CFU/g in contaminated food samples compared with the control group (Miguéis et al., 2017). When LISTEX™ P100 was applied on biofilms formed on stainless steel, 3.5–5.4 log CFU/cm2 reductions were observed depending on which of the 21 L. monocytogenes strains were tested (Soni and Nannapaneni, 2010; Gray et al., 2018). According to the EFSA opinion, LISTEX™ P100 is completely harmless, effective and does not contribute to antibacterial resistance (EFSA, 2016).

Despite many positive results and recommendations, the routine using of bacteriophages in food production industry and final food products is allowed only in some countries and regulations relate only to individual bacteriophage products (Kawacka et al., 2020). Despite the wide use of LISTEXTM P100 (e.g., in the United States, Canada and Switzerland), its acceptance as a processing aid in multiple countries (including Australia, New Zealand, Israel, Switzerland, Canada even in one EU member, Netherlands) (Aprea et al., 2018; Połaska and Sokołowska, 2019), and the previously stated positive opinion of EFSA, the EU has not approved this product for application in food industry (EFSA, 2016). Thus, further studies on the safety and effectiveness of phage-based preparation against L. monocytogenes in foods as well as monitoring of the occurrence of phage-resistant strains in food processing plants are needed.

Conclusion

Listeria monocytogenes is an important foodborne pathogen responsible for severe sporadic infections or disease outbreaks with high case fatality rates worldwide (Scallan et al., 2011; de Noordhout et al., 2014; Buchanan et al., 2017). These ubiquitous bacteria have been isolated from soil, water, feed, and food production environments, where they can survive and persist for a long time. Resistance of such strains to different food processing conditions is contributed to by several external factors such as poor hygiene practice or ineffective sanitizations, but also by the presence of diverse genetic determinants that are responsible for resistance to extreme temperatures, pH, heavy metals, biocides, and the ability to form biofilms. Although there is much knowledge about the mechanisms of stress responses and resistance to adverse conditions of L. monocytogenes, these pathogenic bacteria are still present in the food production environments and pose a severe threat to consumers. Thus, knowledge and understanding of the mechanisms of L. monocytogenes adaptation to environmental stress factors will have a significant influence on the development of new, efficient, and cost-effective methods of the pathogen control in the food industry which is critical to ensure food production safety.

Author Contributions

JO, BL, and KW conceptualized the idea of the manuscript. BL and KW collected relevant literature. JO drafted the manuscript. All authors reviewed, edited the manuscript, and read and approved the final version of the manuscript.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abe, F. (2007). Exploration of the effects of high hydrostatic pressure on microbial growth, physiology and survival: perspectives from piezophysiology. Biosci. Biotechnol. Biochem. 71, 2347–2357. doi: 10.1271/bbb.70015

PubMed Abstract | CrossRef Full Text | Google Scholar

Allerberger, F. (2003). Listeria: Growth, phenotypic differentiation and molecular microbiology. FEMS Immunol. Med. Microbiol. 35, 183–189. doi: 10.1016/S0928-8244(02)00447-9

CrossRef Full Text | Google Scholar

Almakhlafi, H., Lakamraju, M., Podhipleux, N., Singla, B., and McGuire, J. (1995). Measuring surface hydrophobicity as compared to measuring a hydrophobic effect on adhesion events. J. Food Prot. 58, 1034–1037. doi: 10.4315/0362-028X-58.9.1034

PubMed Abstract | CrossRef Full Text | Google Scholar

Alonso, A. N., Perry, K. J., Regeimbal, J. M., Regan, P. M., and Higgins, D. E. (2014). Identification of Listeria monocytogenes determinants required for biofilm formation. PLoS One 9:e113696. doi: 10.1371/journal.pone.0113696

PubMed Abstract | CrossRef Full Text | Google Scholar

Amezaga, M. R., Davidson, I., McLaggan, D., Verheul, A., Abee, T., and Booth, I. R. (1995). The role of peptide metabolism in the growth of Listeria monocytogenes ATCC 23074 at high osmolarity. Microbiology 141, 41–49. doi: 10.1099/00221287-141-1-41

PubMed Abstract | CrossRef Full Text | Google Scholar

Anast, J. M., and Schmitz-Esser, S. (2021). Certain Listeria monocytogenes plasmids contribute to increased UVC ultraviolet light stress. FEMS Microbiol. Lett. 368:fnab123. doi: 10.1093/femsle/fnab123

PubMed Abstract | CrossRef Full Text | Google Scholar

Angelidis, A. S., and Smith, G. M. (2003). Role of the glycine betaine and carnitine transporters in adaptation of Listeria monocytogenes to chill stress in defined medium. Appl. Environ. Microbiol. 69, 7492–7498. doi: 10.1128/AEM.69.12.7492-7498.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Aprea, G., Zocchi, L., Di Fabio, M., De Santis, S., Prencipe, V. A., and Migliorati, G. (2018). The applications of bacteriophages and their lysins as biocontrol agents against the foodborne pathogens Listeria monocytogenes and Campylobacter spp.: an updated look. Vet. Ital. 54, 293–303. doi: 10.12834/VetIt.311.1215.2

PubMed Abstract | CrossRef Full Text | Google Scholar

Arcari, T., Feger, M. L., Guerreiro, D. N., Wu, J., and O’Byrne, C. P. (2020). Comparative review of the responses of Listeria monocytogenes and Escherichia coli to low pH stress. Genes 11:1330. doi: 10.3390/genes11111330

PubMed Abstract | CrossRef Full Text | Google Scholar

Arioli, S., Montanari, C., Magnani, M., Tabanelli, G., Patrignani, F., Lanciotti, R., et al. (2019). Modelling of Listeria monocytogenes Scott A after a mild heat treatment in the presence of thymol and carvacrol: effects on culturability and viability. J. Food Eng. 240, 73–82. doi: 10.1016/j.jfoodeng.2018.07.014

CrossRef Full Text | Google Scholar

Aryal, M., and Muriana, P. M. (2019). Efficacy of commercial sanitizers used in food processing facilities for inactivation of Listeria monocytogenes, E. coli O157:H7, and Salmonella biofilms. Foods 8:639. doi: 10.3390/foods8120639

PubMed Abstract | CrossRef Full Text | Google Scholar

Bae, D., Liu, C., Zhang, T., Jones, M., Peterson, S. N., and Wang, C. (2012). Global gene expression of Listeria monocytogenes to salt stress. J. Food Prot. 75, 906–912. doi: 10.4315/0362-028X.JFP-11-282

PubMed Abstract | CrossRef Full Text | Google Scholar

Ballal, A., Basu, B., and Apte, S. K. (2007). The Kdp-ATPase system and its regulation. J. Biosci. 32, 559–568. doi: 10.1007/s12038-007-0055-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Bayles, D. O., Annous, B. A., and Wilkinson, B. J. (1996). Cold stress proteins induced in Listeria monocytogenes in response to temperature downshock and growth at low temperatures. Appl. Environ. Microbiol. 62, 1116–1119. doi: 10.1128/aem.62.3.1116-1119.1996

PubMed Abstract | CrossRef Full Text | Google Scholar

Bayles, D. O., and Wilkinson, B. J. (2000). Osmoprotectants and cryoprotectants for Listeria monocytogenes. Lett. Appl. Microbiol. 30, 23–27. doi: 10.1046/j.1472-765x.2000.00646.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Beales, N. (2004). Adaptation of microorganisms to cold temperatures, weak acid preservatives, low pH, and osmotic stress: a review. Compr. Rev. Food Sci. Food Safety 3, 1–20. doi: 10.1111/j.1541-4337.2004.tb00057.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Beauchamp, S., and Lacroix, M. (2012). Resistance of the genome of Escherichia coli and Listeria monocytogenes to irradiation evaluated by the induction of cyclobutane pyrimidine dimers and 6-4 photoproducts using gamma and UVC radiations. Rad. Phys. Chem. 81, 1193–1197. doi: 10.1016/j.radphyschem.2011.11.007

CrossRef Full Text | Google Scholar

Becker, L. A., Evans, S. N., Hutkins, R. W., and Benson, A. K. (2000). Role of sigma(B) in adaptation of Listeria monocytogenes to growth at low temperature. J. Bacteriol. 182, 7083–7087. doi: 10.1128/JB.182.24.7083-7087.2000

PubMed Abstract | CrossRef Full Text | Google Scholar

Belval, S. C., Gal, L., Margiewes, S., Garmyn, D., Piveteau, P., and Guzzo, J. (2006). Assessment of the roles of LuxS, S-ribosylhomocysteine and autoinducer 2 in cell attachment during biofilm formaion by Listeria monocytogenes EGD-e. Appl. Environ. Microbiol. 72, 2644–2650. doi: 10.1128/AEM.72.4.2644-2650.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernbom, N., Vogel, B. F., and Gram, L. (2011). Listeria monocytogenes survival of UV-C radiation is enhanced by presence of sodium chloride, organic food material and by biofilm formation. Int. J. Food Microbiol. 147, 69–73. doi: 10.1016/j.ijfoodmicro.2011.03.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Beumer, R. R., Te Giffel, M. C., Cox, L. J., Rombouts, F. M., and Abee, T. (1994). Effect of exogenous proline, betaine, and carnitine on growth of Listeria monocytogenes in a minimal medium. Appl. Environ. Microbiol. 60, 1359–1363. doi: 10.1128/aem.60.4.1359-1363.1994

PubMed Abstract | CrossRef Full Text | Google Scholar

Bonsaglia, E. C. R., Silva, N. C. C., Fernades, J. A., Araújo Júnior, J. P., Tsunemi, M. H., and Rall, V. L. M. (2014). Production of biofilm by Listeria monocytogenes in different materials and temperatures. Food Control 35, 386–391. doi: 10.1016/j.foodcont.2013.07.023

CrossRef Full Text | Google Scholar

Borucki, M. K., Peppin, J. D., White, D., Loge, F., and Call, D. R. (2003). Variation in biofilm formation among strains of Listeria monocytogenes. Appl. Environ. Microbiol. 69, 7336–7342. doi: 10.1128/AEM.69.12.7336-7342.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Botta, C., Ferrocino, I., Pessione, A., Cocolin, L., and Rantsiou, K. (2020). Spatiotemporal distribution of the environmental microbiota in food processing plants as impacted by cleaning and sanitizing procedures: The case of slaughterhouses and gaseous ozone. Appl. Environ. Microbiol. 86:e01861-20. doi: 10.1128/AEM.01861-20

PubMed Abstract | CrossRef Full Text | Google Scholar

Bowman, J. P., Bittencourt, C. R., and Ross, T. (2008). Differential gene expression of Listeria monocytogenes during high hydrostatic pressure processing. Microbiology 154, 462–475. doi: 10.1099/mic.0.2007/010314-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Bozoglu, F., Alpas, H., and Kaletunç, G. (2004). Injury recovery of foodborne pathogens in high hydrostatic pressure treated milk during storage. FEMS Immunol. Med. Microbiol. 40, 243–247. doi: 10.1016/S0928-8244(04)00002-1

CrossRef Full Text | Google Scholar

Bremer, P. J., Monk, I., and Butler, R. (2002). Inactivation of Listeria monocytogenes/Flavobacterium spp. biofilms using chlorine: Impact of substrate, pH, time and concentration. Lett. Appl. Microbiol. 35, 321–325. doi: 10.1046/j.1472-765X.2002.01198.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Brøndsted, L., Kallipolitis, B. H., Ingmer, H., and Knöchel, S. (2003). kdpE and a putative RsbQ homologue contribute to growth of Listeria monocytogenes at high osmolarity and low temperature. FEMS Microbiol. Lett. 219, 233–239. doi: 10.1016/S0378-1097(03)00052-1

CrossRef Full Text | Google Scholar

Brown, P., Chen, Y., Siletzky, R., Parsons, C., Jaykus, L.-A., Eifert, J., et al. (2021). Harnessing whole genome sequence data for facility-specific signatures for Listeria monocytogenes: a case study with turkey processing plants in the United States. Front. Sustain Food Syst. 5:742353. doi: 10.3389/fsufs.2021.742353

CrossRef Full Text | Google Scholar

Bruschi, C., Komora, N., Castro, S. M., Saraiva, J., Ferreira, V. B., and Teixeira, P. (2017). High hydrostatic pressure effects on Listeria monocytogenes and L. innocua: evidence for variability in inactivation behaviour and in resistance to pediocin bacHA-6111-2. Food Microbiol. 64, 226–231. doi: 10.1016/j.fm.2017.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Buchanan, R. L., Gorris, L. G. M., Hayman, M. M., Jackson, T. C., and Whiting, R. C. (2017). A review of Listeria monocytogenes: an update on outbreaks, virulence, dose-response, ecology, and risk assessments. Food Control 75, 1–13. doi: 10.1016/j.foodcont.2016.12.016

CrossRef Full Text | Google Scholar

Buchanan, R. L., Stahl, H. G., and Whiting, R. C. (1989). Effects and interactions of temperature, pH, atmosphere, sodium chloride, and sodium nitrite on the growth of Listeria monocytogenes. J. Food Prot. 52, 844–851. doi: 10.4315/0362-028X-52.12.844

PubMed Abstract | CrossRef Full Text | Google Scholar

Bucur, F. I., Grigore-Gurgu, L., Crauwels, P., Riedel, C. U., and Nicolau, A. I. (2018). Resistance of Listeria monocytogenes to stress conditions encountered in food and food processing environments. Front. Microbiol. 9:2700. doi: 10.3389/fmicb.2018.02700

PubMed Abstract | CrossRef Full Text | Google Scholar

Carpentier, B., and Cerf, O. (2011). Review - persistence of Listeria monocytogenes in food industry equipment and premises. Intern. J. Food Microbiol. 145, 1–8. doi: 10.1016/j.ijfoodmicro.2011.01.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Carpentier, B., and Chassaing, D. (2004). Interactions in biofilms between Listeria monocytogenes and resident microorganisms from food industry premises. Int. J. Food Microbiol. 97, 111–122. doi: 10.1016/j.ijfoodmicro.2004.03.031

PubMed Abstract | CrossRef Full Text | Google Scholar

Casadei, M. A., Esteves de Matos, R., Harrison, S. T., and Gaze, J. E. (1998). Heat resistance of Listeria monocytogenes in dairy products as affected by the growth medium. J. Appl. Microbiol. 84, 234–239. doi: 10.1046/j.1365-2672.1998.00334.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Cayley, S., Lewis, B. A., and Record, M. T. Jr. (1992). Origins of the osmoprotective properties of betaine and proline in Escherichia coli K-12. J. Bacteriol. 174, 1586–1595. doi: 10.1128/jb.174.5.1586-1595.1992

PubMed Abstract | CrossRef Full Text | Google Scholar

Chan, Y. C., Boor, K. J., and Wiedmann, M. (2007a). sB-dependent and sB-independent mechanisms contribute to transcription of Listeria monocytogenes cold stress genes during cold shock and coldg rowth. Appl. Environ. Microbiol. 73, 6019–6029. doi: 10.1128/AEM.00714-07

PubMed Abstract | CrossRef Full Text | Google Scholar

Chan, Y. C., Raengpradub, S., Boor, K. J., and Wiedmann, M. (2007b). Microarray-based characterization of the Listeria monocytogenes cold regulon in log-and stationary-phase cells. Appl. Environ. Microbiol. 73, 6484–6498. doi: 10.1128/AEM.00897-07

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, C. Y., Hsieh, Y. H., Shih, I. C., Hsu, S. S., and Wang, K. H. (2000). The formation and control of disinfection by-products using chlorine dioxide. Chemosphere 41, 1181–1186. doi: 10.1016/s0045-6535(00)00010-2

CrossRef Full Text | Google Scholar

Chen, H., Neetoo, H., Ye, M., and Joerger, R. D. (2009). 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. 26, 404–408. doi: 10.1016/j.fm.2009.01.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Cherifi, T., Carrillo, C., Lambert, D., Miniaï, I., Quessy, S., Larivière-Gauthier, G., et al. (2018). Genomic characterization of Listeria monocytogenes isolates reveals that their persistence in a pig slaughterhouse is linked to the presence of benzalkonium chloride resistance genes. BMC Microbiol. 18:220. doi: 10.1186/s12866-018-1363-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Chmielowska, C., Korsak, D., Szuplewska, M., Grzelecka, M., Maćkiw, E., Stasiak, M., et al. (2021). Benzalkonium chloride and heavy metal resistance profiles of Listeria monocytogenes strains isolated from fish, fish products and food-producing factories in Poland. Food Microbiol. 98:103756. doi: 10.1016/j.fm.2021.103756

PubMed Abstract | CrossRef Full Text | Google Scholar

Colagiorgi, A., Bruini, I., Di Ciccio, P. A., Zanardi, E., Ghidini, S., and Ianieri, A. (2017). Listeria monocytogenes biofilms in the wonderland of food industry. Pathogens 6:41. doi: 10.3390/pathogens6030041

PubMed Abstract | CrossRef Full Text | Google Scholar

Colagiorgi, A., Di Ciccio, P., Zanardi, E., Ghidini, S., and Ianieri, A. (2016). A look inside the Listeria monocytogenes biofilms extracellular matrix. Microorganisms 4:22. doi: 10.3390/microorganisms4030022

PubMed Abstract | CrossRef Full Text | Google Scholar

Conficoni, D., Losasso, C., Cortini, E. Di Cesare, A., Cibin, V., Giaccone, V., et al. (2016). Resistance to biocides in Listeria monocytogenes collected in meat-processing environments. Front. Microbiol. 7:1627. doi: 10.3389/fmicb.2016.01627

PubMed Abstract | CrossRef Full Text | Google Scholar

Cooper, A. L., Carrillo, C. D., Deschênes, M., and Blais, B. W. (2021). Genomic markers for quaternary ammonium compound resistance as a persistence indicator for Listeria monocytogenes contamination in food manufacturing environments. J. Food Prot. 84, 389–398. doi: 10.4315/JFP-20-328

PubMed Abstract | CrossRef Full Text | Google Scholar

Cordero, N., Maza, F., Navea-Perez, H., Aravena, A., Marquez-Fontt, B., Navarrete, P., et al. (2016). Different transcriptional responses from slow and fast growth rate strains of Listeria monocytogenes adapted to low temperature. Front. Microbiol. 7:229. doi: 10.3389/fmicb.2016.00229

PubMed Abstract | CrossRef Full Text | Google Scholar

Cortes, B. W., Naditz, A. L., Anast, J. M., and Smitz-Esser, S. (2020). Transcriptome sequencing of Listeria monocytogenes reveals major gene expression changes in response to lactic acid stress exposure but a less pronounced response to oxidative stress. Front. Microbiol. 10:3110. doi: 10.3389/fmicb.2019.03110

PubMed Abstract | CrossRef Full Text | Google Scholar

Cotter, P. D., Emerson, N., Gahan, C. G., and Hill, C. (1999). Identification and disruption of lisRK, a genetic locus encoding a two-component signal transduction system involved in stress tolerance and virulence in Listeria monocytogenes. J. Bacteriol. 181, 6840–6843. doi: 10.1128/JB.181.21.6840-6843.1999

PubMed Abstract | CrossRef Full Text | Google Scholar

Cotter, P. D., Gahan, C. G. M., and Hill, C. (2000). Analysis of the role of the Listeria monocytogenes F0F1-ATPase operon in the acid tolerance response. Intern. J. Food Microbiol. 60, 137–146. doi: 10.1016/s0168-1605(00)00305-6

CrossRef Full Text | Google Scholar

Cotter, P. D., and Hill, C. (2003). Surviving the acid test: responses of gram-positive bacteria to low pH. Microbiol. Mol. Biol. Rev. 67, 429–453. doi: 10.1128/MMBR.67.3.429-453.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Cotter, P. D., O’Reilly, K., and Hill, C. (2001). Role of the glutamate decarboxylase acid resistance system in the survival of Listeria monocytogenes LO28 in low pH foods. J. Food Prot. 64, 1362–1368. doi: 10.4315/0362-028x-64.9.1362

PubMed Abstract | CrossRef Full Text | Google Scholar

Cotter, P. D., Ryan, S., Gahan, C. G. M., and Hill, C. (2005). Presence of GadD1 glutamate decarboxylase in selected Listeria monocytogenes strains is associated with an ability to grow at low pH. Appl. Environ. Microbiol. 71, 2832–2839.

Google Scholar

de Noordhout, C. M., Devleesschauwer, B., Angulo, F. J., Verbeke, G., Haagsma, J., Kirk, M., et al. (2014). The global burden of listeriosis: a systematic review and meta-analysis. Lancet Infect. Dis. 14, 1073–1082. doi: 10.1016/S1473-3099(14)70870-9

CrossRef Full Text | Google Scholar

Denyer, S. P., and Stewart, G. S. A. B. (1988). Mechanisms of action of disinfectants. Intern. Biodeter. Biodegr. 41, 261–268. doi: 10.1016/S0964-8305(98)00023-7

CrossRef Full Text | Google Scholar

Dhama, K., Karthik, K., Tiwari, R., Shabbir, M. Z., Barbuddhe, S., Malik, S. V. S., et al. (2015). Listeriosis in animals, its public health significance (food-borne zoonosis) and advances in diagnosis and control: a comprehensive review. Vet. Quart. 35, 211–235. doi: 10.1080/01652176.2015.1063023

PubMed Abstract | CrossRef Full Text | Google Scholar

Di Bonaventura, G., Piccolomini, R., Paludi, D., D’Orio, V., Vergara, A., Conter, M., et al. (2008). Influence of temperature on biofilm formation by Listeria monocytogenes on various food-contact surfaces: relationship with motility and cell surface hydrophobicity. J. Appl. Microbiol. 104, 1552–1561. doi: 10.1111/j.1365-2672.2007.03688.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Djordjevic, D., Wiedmann, M., and McLandsborough, L. A. (2002). Microtiter plate assay for assessment of Listeria monocytogenes biofilm formation. Appl. Environ. Microbiol. 68, 2950–2958. doi: 10.1128/AEM.68.6.2950-2958.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Dmitrieva, N. I., Cai, Q., and Burg, M. B. (2004). Cells adapted to high NaCl have many DNA breaks and impaired DNA repair both in cell culture and in vivo. Proc. Nat. Acad. Sci. U.S.A. 101, 2317–2322. doi: 10.1073/pnas.0308463100

PubMed Abstract | CrossRef Full Text | Google Scholar

Doghri, I., Cherifi, T., Goetz, C., Malouin, F., Jacques, M., and Fravalo, P. (2021). Counteracting bacterial motility: a promising strategy to narrow Listeria monocytogenes biofilm in food processing industry. Front. Microbiol. 12:673484. doi: 10.3389/fmicb.2021.673484

PubMed Abstract | CrossRef Full Text | Google Scholar

Doijad, S. P., Barbuddhe, S. B., Garg, S., Poharkar, K. V., Kalorey, D. R., Kurkure, N. V., et al. (2015). Biofilm-forming abilities of Listeria monocytogenes serotypes isolated from different sources. PLoS One 10:e0137046. doi: 10.1371/journal.pone.0137046

PubMed Abstract | CrossRef Full Text | Google Scholar

Doyle, M. E., Mazzotta, A. S., Wang, T., Wiseman, D. W., and Scott, V. N. (2001). Heat resistance of Listeria monocytogenes. J. Food Prot. 64, 410–429. doi: 10.4315/0362-028X-64.3.410

PubMed Abstract | CrossRef Full Text | Google Scholar

Duru, I. C., Bucur, F. I., Andreevskaya, M., Nikparvar, B., Ylinen, A., Grigore-Gurgu, L., et al. (2021). High-pressure processing-induced transcriptome response during recovery of Listeria monocytogenes. BMC Genomics 22:117. doi: 10.1186/s12864-021-07407-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Dutta, V., Elhanaf, D., and Kathariou, S. (2013). Conservation and distribution of the benzalkonium chloride resistance cassette bcrABC in Listeria monocytogenes. Appl. Environ. Microbiol. 79, 6067–6074. doi: 10.1128/AEM.01751-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Dutta, V., Elhanafi, D., Osborne, J., Martinez, M. R., and Kathariou, S. (2014). Genetic characterization of plasmid-associated triphenylmethane reductase in Listeria monocytogenes. Appl. Environ. Microbiol. 80, 5379–5385. doi: 10.1128/AEM.01398-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Duze, S. T., Marimani, M., and Patel, M. (2021). Tolerance of Listeria monocytogenes to biocides used in food processing environments. Food Microbiol. 97:103758. doi: 10.1016/j.fm.2021.103758

PubMed Abstract | CrossRef Full Text | Google Scholar

Ebner, R., Stephan, R., Althaus, D., Brisse, S., Maury, M., and Tasara, T. (2015). Phenotypic and genotypic characteristics of Listeria monocytogenes strains isolated during 2011-2014 from different food matrices in Switzerland. Food Control 57, 321–326.

Google Scholar

EFSA (2016). Scientific opinion on the evaluation of the safety and efficacy of ListexTMP100 for reduction of pathogens on different ready-to-eat (RTE) food products. EFSA J. 14:4565. doi: 10.2903/j.efsa.2016.4565

CrossRef Full Text | Google Scholar

EFSA and ECDC (2021). The European Union One Health 2020 zoonoses report. EFSA J. 19:6971. doi: 10.2903/j.efsa.2021.6971

CrossRef Full Text | Google Scholar

Elhanafi, D., Dutta, V., and Kathariou, S. (2010). Genetic characterization of plasmid-associated benzalkonium chloride resistance determinants in a Listeria monocytogenes strain from the 1998–1999 outbreak. Appl. Environ. Microbiol. 76, 8231–8238. doi: 10.1128/AEM.02056-10

PubMed Abstract | CrossRef Full Text | Google Scholar

El-Kest, S. E., and Marth, E. H. (1988). Inactivation of Listeria monocytogenes by chlorine. J. Food Prot. 51, 520–524. doi: 10.4315/0362-028X-51.7.520

PubMed Abstract | CrossRef Full Text | Google Scholar

Eshwar, A. K., Guldimann, C., Oevermann, A., and Tasara, T. (2017). Cold-shock domain family proteins (Csps) are involved in regulation of virulence, cellular aggregation, and flagella-based motility in Listeria monocytogenes. Front. Cell. Infect. Microbiol. 7:453. doi: 10.3389/fcimb.2017.00453

PubMed Abstract | CrossRef Full Text | Google Scholar

Fagerlund, A., Langsrud, S., and Møretrø, T. (2021). Microbial diversity and ecology of biofilms in food industry environments associated with Listeria monocytogenes persistence. Curr. Opin. Food Sci. 37, 171–178. doi: 10.1016/j.cofs.2020.10.015

CrossRef Full Text | Google Scholar

Fan, Y., Qiao, J., Lu, Z., Fen, Z., Tao, Y., Lv, F., et al. (2020). Influence of different factors on biofilm formation of Listeria monocytogenes and the regulation of cheY gene. Food Res. Inter. 137:109405. doi: 10.1016/j.foodres.2020.109405

PubMed Abstract | CrossRef Full Text | Google Scholar

Ferreira, M., Almeida, A., Delgadillo, I., Saraiva, J., and Cunha, Â (2016). Susceptibility of Listeria monocytogenes to high pressure processing: a review. Food Rev. Intern. 32, 377–399. doi: 10.1080/87559129.2015.1094816

CrossRef Full Text | Google Scholar

Ferreira, V., Wiedmann, M., Teixeira, P., and Stasiewicz, M. J. (2014). Listeria monocytogenes persistence in food-associated environments: epidemiology, strain characteristics, and implications for public health. J. Food Prot. 77, 150–170. doi: 10.4315/0362-028X.JFP-13-150

PubMed Abstract | CrossRef Full Text | Google Scholar

Fox, E. M., Leonard, N., and Jordan, K. (2011). Physiological and transcriptional characterization of persistent and non-persistent Listeria monocytogenes isolates. Appl. Environ. Microbiol. 77, 6559–6569. doi: 10.1128/AEM.05529-11

PubMed Abstract | CrossRef Full Text | Google Scholar

Franciosa, G., Maugliani, A., Scalfaro, C., Floridi, F., and Aureli, P. (2009). Expression of internalin A and biofilm formation among Listeria monocytogenes clinical isolates. Intern. J. Immunopath. Pharmacol. 22, 183–193. doi: 10.1177/039463200902200121

PubMed Abstract | CrossRef Full Text | Google Scholar

Fraser, K. R., Harvie, D., Coote, P. J., and O’Byrne, C. P. (2000). Identification and characterization of an ATP binding cassette L-carnitine transporter in Listeria monocytogenes. Appl. Environ. Microbiol. 66, 4696–4704. doi: 10.1128/AEM.66.11.4696-4704.2000

PubMed Abstract | CrossRef Full Text | Google Scholar

Gandhi, M., and Chikindas, M. L. (2007). Listeria: A foodborne pathogen that knows how to survive. Intern. J. Food Microbiol. 113, 1–15. doi: 10.1016/j.ijfoodmicro.2006.07.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Gänzle, M., and Liu, Y. (2015). Mechanisms of pressure-mediated cell death and injury in Escherichia coli: from fundamentals to food applications. Front. Microbiol. 5:599. doi: 10.3389/fmicb.2015.00599

PubMed Abstract | CrossRef Full Text | Google Scholar

Gardan, R., Cossart, P., and the European Listeria Genome Consortium, and Labadie, J. (2003a). Identification of Listeria monocytogenes genes involved in salt and alkaline-pH tolerance. Appl. Environ. Microbiol. 69, 3137–3143.

Google Scholar

Gardan, R., Duché, O., Leroy-Sétrin, S., and the European Listeria Genome Consortium, and Labadie, J. (2003b). Role of ctc from Listeria monocytogenes in osmotolerance. Appl. Environ. Microbiol. 69, 154–161.

Google Scholar

Garmyn, D., Gal, L., Lemaitre, J. P., Hartmann, A., and Piveteau, P. (2009). Communication and autoinduction in the species Listeria monocytogenes: a central role for the agr system. Commun. Integr. Biol. 2, 371–374. doi: 10.4161/cib.2.4.8610

PubMed Abstract | CrossRef Full Text | Google Scholar

Gayán, E., Serrano, M. J., Pagán, R., Álvarez, I., and Condón, S. (2015). Environmental and biological factors influencing the UV-C resistance of Listeria monocytogenes. Food Microbiol. 46, 246–253. doi: 10.1016/j.fm.2014.08.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Gerba, C. P. (2015). Quaternary ammonium biocides: efficacy in application. Appl. Environ. Microbiol. 81, 464–469. doi: 10.1128/AEM.02633-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Gilmartin, N., Gião, M. S., Keevil, C. W., and O’Kennedy, R. (2016). Differential internalin A levels in biofilms of Listeria monocytogenes grown on different surfaces and nutrient conditions. Intern. J. Food Microbiol. 219, 50–55. doi: 10.1016/j.ijfoodmicro.2015.12.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Giotis, E. S., Blair, I. S., and McDowell, D. A. (2009). Effects of short-term alkaline adaptation on surface properties of Listeria monocytogenes 10403S. Open Food Sci. J. 3, 62–65. doi: 10.2174/1874256400903010062

CrossRef Full Text | Google Scholar

Giotis, E. S., McDowell, D. A., Blair, I. S., and Wilkinson, B. J. (2007). Role of branched-chain fatty acids in pH stress tolerance in Listeria monocytogenes. Appl. Environ. Microbiol. 73, 997–1001.

Google Scholar

Giotis, E. S., Muthaiyan, A., Natesan, S., Wilkinson, B. J., Blair, I. S., and McDowell, D. A. (2010). Transcriptome analysis of alkali shock and alkali adaptation in Listeria monocytogenes 10403S. Foodborne Path. Dis. 7, 1147–1157.

Google Scholar

Gómez-López, V. M., Ragaert, P., Debevere, J., and Devlieghere, F. (2007). Pulsed light for food decontamination: a review. Trends Food Sci. Technol. 18, 464–473.

Google Scholar

Gorski, L., Duhé, J. M., and Flaherty, D. (2009). The use of flagella and motility for plant colonization and fitness by different strains of the foodborne pathogen Listeria monocytogenes. PLoS One 4:e5142. doi: 10.1371/journal.pone.0005142

PubMed Abstract | CrossRef Full Text | Google Scholar

Gougouli, M., Angelidis, A. S., and Koutsoumanis, K. (2008). A study on the kinetic behavior of Listeria monocytogenes in ice cream stored under static and dynamic chilling and freezing conditions. J. Dairy. Sci. 91, 523–530. doi: 10.3168/jds.2007-0255

PubMed Abstract | CrossRef Full Text | Google Scholar

Gray, J. A., Chandry, P. S., Kaur, M., Kocharunchitt, C., Bowman, J. P., and Fox, E. M. (2018). Novel biocontrol methods for Listeria monocytogenes biofilms in food production facilities. Front. Microbiol. 9:605. doi: 10.3389/fmicb.2018.00605.

PubMed Abstract | CrossRef Full Text | Google Scholar

Gray, M. J., Freitag, N. E., and Boor, K. J. (2006). How the bacterial pathogen Listeria monocytogenes mediates the switch from environmental Dr. Jekyll to pathogenic Mr. Hyde. Infect. Immun. 74, 2505–2512.

Google Scholar

Gründling, A., Burrack, L. S., Bouwer, H. G. A., and Higgins, D. E. (2004). Listeria monocytogenes regulates flagellar motility gene expression through MogR, a transcriptional repressor required for virulence. Proc. Nat. Acad. Sci. U.S.A. 101, 12318–12323. doi: 10.1073/pnas.0404924101

PubMed Abstract | CrossRef Full Text | Google Scholar

Guerreiro, D. N., Wu, J., Dessaux, C., Oliveira, A. H., Tiensuu, T., Gudynaite, D., et al. (2020). Mild stress conditions during laboratory culture promote the proliferation of mutations that negatively affect Sigma B activity in Listeria monocytogenes. J. Bacteriol. 202:e00751–19. doi: 10.1128/JB.00751-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Harter, E., Wagner, E. M., Zaiser, A., Halecker, S., Wagner, M., and Rychli, K. (2017). Stress survival islet 2, predominantly present in Listeria monocytogenes strains of sequence type 121, is involved in the alkaline and oxidative stress responses. Appl. Environ. Microbiol. 83:e00827-17. doi: 10.1128/AEM.00827-17

PubMed Abstract | CrossRef Full Text | Google Scholar

Hartl, F. U., and Hayer-Hartl, M. (2002). Molecular chaperone in the cytosol: from nascent chain to folded protein. Science 295, 1852–1858.

Google Scholar

Harvey, J., and Gilmour, A. (2001). Characterization of recurrent and sporadic Listeria monocytogenes isolates from raw milk and nondairy foods by pulsed-field gel electrophoresis, monocin typing, plasmid profiling, and cadmium and antibiotic resistance determination. Appl. Environ. Microbiol. 67, 840–847.

Google Scholar

Harvey, J., Keenan, K. P., and Gilmour, A. (2007). Assessing biofilm formation by Listeria monocytogenes strains. Food Microbiol. 24, 380–392.

Google Scholar

Haubert, L., Zehetmeyr, M. L., and da Silva, W. P. (2019). Resistance to benzalkonium chloride and cadmium chloride in Listeria monocytogenes isolates from food and food-processing environments in southern Brazil. Can. J. Microbiol. 65, 429–435.

Google Scholar

Hébraud, M., and Guzzo, J. (2000). The main cold shock protein of Listeria monocytogenes belongs to the family of ferritin-like proteins. FEMS Microbiol. Lett. 190, 29–34.

Google Scholar

Hecker, M., Schumann, W., and Volker, U. (1996). Heat-shock and general stress response in Bacillus subtilis. Mol. Microbiol. 19, 417–428.

Google Scholar

Heir, E., Lindstedt, B. A., Røtterud, O. J., Vardund, T., Kapperud, G., and Nesbakken, T. (2004). Molecular epidemiology and disinfectant susceptibility of Listeria monocytogenes from meat processing plants and human infections. Intern. J. Food Microbiol. 96, 85–96.

Google Scholar

Hendrick, J. P., and Hartl, F. U. (1993). Molecular chaperone functions of heat shock proteins. Ann. Rev. Biochem. 62, 349–384.

Google Scholar

Heras de las, A., Cain, R. J., Bielecka, M. K., and Vázquez-Boland, J. A. (2011). Regulation of Listeria virulence: PrfA master and commander. Curr. Opin. Microbiol. 14, 118–127. doi: 10.1016/j.mib.2011.01.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Hill, C., Cotter, P. D., Sleator, R. D., and Gahan, C. G. M. (2002). Bacterial stress response in Listeria monocytogenes: jumping the hurdles imposed by minimal processing. Intern. Dairy J. 12, 273–283.

Google Scholar

Hill, D., Sugrue, I., Arendt, E., Hill, C., and Stanton, C. (2017). Recent advances in microbial fermentation for dairy and health [version 1; referees:3 approved]. F1000Res. 6:751.

Google Scholar

Hingston, P., Brenner, T., Truelstrup Hansen, L., and Wang, S. (2019). Comparative analysis of Listeria monocytogenes plasmids and expression levels of plasmid-encoded genes during growth under salt and acid stress conditions. Toxins 11:426.

Google Scholar

Hoelzer, K., Pouillot, R., Gallagher, D., Silverman, M. B., Kause, J., and Dennis, S. (2012). Estimation of Listeria monocytogenes transfer coefficients and efficacy of bacterial removal through cleaning and sanitation. Intern. J. Food Microbiol. 157, 267–277.

Google Scholar

Hu, Y., Oliver, H. F., Raengpradub, S., Palmer, M. E., Orsi, R. H., Wiedmann, M., et al. (2007). Transcriptomic and phenotypic analyses suggest a network between the transcriptional regulators HrcA and σB in Listeria monocytogenes. Appl. Environ. Microbiol. 73, 7981–7991.

Google Scholar

Huang, H. W., Lung, H. M., Yang, B. B., and Wang, C. Y. (2014). Responses of microorganisms to high hydrostatic pressure processing. Food Control 40, 250–259.

Google Scholar

Ishaq, A., Ebner, P. D., Syed, Q. A., and ur Rahman, H. U. (2020). Employing list-shield bacteriophage as a bio-control intervention for Listeria monocytogenes from raw beef surface and maintain meat quality during refrigeration storage. LWT- Food Sci. Technol. 132:109784.

Google Scholar

Jacobsohn, M. K., Lehman, M. M., and Jacobsohn, G. M. (1992). Cell membranes and multilamellar vesicles - influence of pH on solvent induced damage. Lipids 27, 694–700.

Google Scholar

Jensen, A., Larsen, M. H., Ingmer, H., Vogel, B. F., and Gram, L. (2007). Sodium chloride enhances adherence and aggregation and strain variation influences invasiveness of Listeria monocytogenes strains. J. Food Prot. 70, 592–599.

Google Scholar

Jesse, H. E., Roberts, I. S., and Cavet, J. S. (2014). Metal ion homeostasis in Listeria monocytogenes and importance in host-pathogen interactions. Adv. Microbial Phys. 65, 83–123.

Google Scholar

Jiang, X., Yu, T., Xu, Y., Wang, H., Korkeala, H., and Shi, L. (2019). MdrL, a major facilitator superfamily efflux pump of Listeria monocytogenes involved in tolerance to benzalkonium chloride. Appl. Gen. Mol. Biotechnol. 103, 1339–1350.

Google Scholar

Juneja, V. K., Foglia, T. A., and Marmer, B. S. (1998). Heat resistance and fatty acid composition of Listeria monocytogenes: effect of pH, acidulant, and growth temperature. J. Food Prot. 61, 683–687.

Google Scholar

Kallipolitis, B. H., and Ingmer, H. (2001). Listeria monocytogenes response regulators important for stress tolerance and pathogenesis. FEMS Microbiol. Lett. 204, 111–115.

Google Scholar

Karatzas, K. A. G., and Bennik, M. H. J. (2002). Characterization of a Listeria monocytogenes Scott A isolate with high tolerance towards high hydrostatic pressure. Appl. Environ. Microbiol. 68, 3183–3189.

Google Scholar

Karatzas, K. A. G., Valdramidis, V. P., and Wells-Bennik, M. H. J. (2005). Contingency locus in ctsR of Listeria monocytogenes Scott A: a strategy for occurrence of abundant piezotolerant isolates within clonal populations. Appl. Environ. Microbiol. 71, 8390–8396.

Google Scholar

Karatzas, K. A. G., Wouters, J. A., Gahan, C. G. M., Hill, C., Abee, T., and Bennik, M. H. J. (2003). The CtsR regulator of Listeria monocytogenes contains a variant glycine repeat region that affects piezotolerance, stress resistance, motility and virulence. Mol. Microbiol. 49, 1227–1238.

Google Scholar

Kastbjerg, V. G., and Gram, L. (2009). Model systems allowing quantification of sensitivity to disinfectants and comparison of disinfectant susceptibility of persistent and presumed nonpersistent Listeria monocytogenes. J. Appl. Microbiol. 106, 1667–1681.

Google Scholar

Kathariou, S. (2002). Listeria monocytogenes virulence and pathogenicity, a food safety perspective. J. Food Prot. 65, 1811–1829.

Google Scholar

Kawacka, I., Olejnik-Schmidt, A., Schmidt, M., and Sip, A. (2020). Effectiveness of phage-based inhibition of Listeria monocytogenes in food products and food processing environments. Microorganisms 8:1764.

Google Scholar

Keeney, K., Trmcic, A., Zhu, Z., Delaquis, P., and Wang, S. (2018). Stress survival islet 1 contributes to serotype-specific differences in biofilm formation in Listeria monocytogenes. Lett. Appl. Microbiol. 67, 530–536.

Google Scholar

Khan, I., Khan, J., Miskeen, S., Tango, C. N., Park, J. S., and Oh, D. H. (2016). Prevalence and control of Listeria monocytogenes in the food industry – a review. Czech J. Food Sci. 34, 469–487.

Google Scholar

Kim, S. H., Gorski, L., Reynolds, J., Orozco, E., Fielding, S., Park, Y. H., et al. (2006). Role of uvrA in the growth and survival of Listeria monocytogenes under UV radiation and acid and bile stress. J. Food Prot. 69, 3031–3036. doi: 10.4315/0362-028x-69.12.3031.

PubMed Abstract | CrossRef Full Text | Google Scholar

Ko, R., and Smith, L. T. (1999). Identification of an ATP-driven, osmoregulated glycine betaine transport system in Listeria monocytogenes. Appl. Environ. Microbiol. 65, 4040–4048.

Google Scholar

Ko, R., Smith, L. T., and Smith, G. M. (1994). Glycine betaine confers enhanced osmotolerance and cryotolerance of L. monocytogenes. J. Bacteriol. 176, 426–431.

Google Scholar

Kode, D., Nannapaneni, R., and Chang, S. (2021). Low-level tolerance to antibiotic trimethoprim in QAC-adapted subpopulations of Listeria monocytogenes. Foods 10:1800.

Google Scholar

Koutsoumanis, K. P., Kendall, P. A., and Sofos, J. N. (2003). Effect of food processing-related stresses on acid tolerance of Listeria monocytogenes. Appl. Environ. Microbiol. 69, 7514–7516.

Google Scholar

Kovacevic, J., Ziegler, J., Wałecka-Zacharska, E., Reimer, A., Kitts, D. D., and Gilmour, M. W. (2016). Tolerance of Listeria monocytogenes to quaternary ammonium sanitizers is mediated by a novel efflux pump encoded by emrE. Appl. Environ. Microbiol. 82, 939–953.

Google Scholar

Kremer, P. H. C., Lees, J. A., Koopmans, M. M., Ferwerda, B., Arends, A. W. M., Feller, M., et al. (2017). Benzalkonium tolerance genes and outcome in Listeria monocytogenes meningitis. Clin. Microbiol. Infect. 23, 265.e1–265.e7.

Google Scholar

Kropac, A. C., Eshwar, A. K., Stephan, R., and Tasara, T. (2019). New insights on the role of the pLMST6 plasmid in Listeria monocytogenes biocide tolerance and virulence. Front. Microbiol. 10:1538. doi: 10.3389/fmicb.2019.01538

PubMed Abstract | CrossRef Full Text | Google Scholar

Krulwich, T. A., Ito, M., Gilmour, R., and Guffanti, A. A. (1997). Mechanisms of cytoplasmic pH regulation in alkaliphilic strains of Bacillus. Extremophiles 1, 163–169.

Google Scholar

Kuenne, C., Voget, S., Pischimarov, J., Oehm, S., Goesmann, A., Daniel, R., et al. (2010). Comparative analysis of plasmids in the genus Listeria. PLoS One 5:e12511. doi: 10.1371/journal.pone.0012511

PubMed Abstract | CrossRef Full Text | Google Scholar

Lachtara, B., Osek, J., and Wieczorek, K. (2021). Molecular typing of Listeria monocytogenes IVb serogroup isolated from food and food production environments in Poland. Pathogens 10:482. doi: 10.3390/pathogens10040482

PubMed Abstract | CrossRef Full Text | Google Scholar

Lacroix, M., and Follett, P. (2015). Combination irradiation treatments for food safety and phytosanitary uses. Stewart Postharvest Rev. 3:4. doi: 10.2212/spr.2015.3.4

PubMed Abstract | CrossRef Full Text | Google Scholar

Lacroix, M., and Ouattara, B. (2000). Combined industrial processes with irradiation to assure innocuity and preservation of food products – a review. Food Res. Intern. 33, 719–724. doi: 10.1016/S0963-9969(00)00085-5

CrossRef Full Text | Google Scholar

Lebrun, M., Audurier, A., and Cossart, P. (1994). Plasmid-borne cadmium resistance genes in Listeria monocytogenes are present on Tn5422, a novel transposon closely related to Tn917. J. Bacteriol. 176, 3049–3061. doi: 10.1128/jb.176.10.3049-3061.1994

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, B. H., Cole, S., Badel-Berchoux, S., Guillier, L., Felix, B., Krezdorn, N., et al. (2019). Biofilm formation of Listeria monocytogenes strains under food processing environments and pan-genome-wide association study. Front. Microbiol. 10:2698. doi: 10.3389/fmicb.2019.02698

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, S., Rakic-Martinez, M., Graves, L. M., Ward, T. J., Siletzky, R. M., and Kathariou, S. (2013). Genetic determinants for cadmium and arsenic resistance among Listeria monocytogenes serotype 4B isolates from sporadic human listeriosis patients. Appl. Environ. Microbiol. 79, 2471–2476. doi: 10.1128/AEM.03551-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, S., Ward, T. J., Jima, D. D., Parsons, C., and Kathariou, S. (2017). The arsenic resistance-associated Listeria genomic island LGI2 exhibits sequence and integration site diversity and a propensity for three Listeria monocytogenes clones with enhanced virulence. Appl. Environ. Microbiol. 83:e01189-17. doi: 10.1128/AEM.01189-17

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemon, K. P., Freitag, N. E., and Kolter, R. (2010). The virulence regulator PrfA promotes biofilm formation by Listeria monocytogenes. J. Bacteriol. 192, 3969–3976. doi: 10.1128/JB.00179-10

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemon, K. P., Higgins, D. E., and Kolter, R. (2007). Flagellar motility is critical for Listeria monocytogenes biofilm formation. J. Bacteriol. 189, 4418–4424. doi: 10.1128/JB.01967-06

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, Q., Guo, A., Ma, Y., Liu, L., Liu, W., Zhong, Y., et al. (2021). Gene analysis of Listeria monocytogenes suspended aggregates induced by Ralstonia insidiosa cell-free supernatants under nutrient-poor environments. Microorganisms 9:2591. doi: 10.3390/microorganisms9122591

PubMed Abstract | CrossRef Full Text | Google Scholar

Lianou, A., Nychas, G.-J. E., and Koutsoumanis, K. P. (2020). Strain variability in biofilm formation: a food safety and quality perspective. Food Res. Intern. 137:109424. doi: 10.1016/j.foodres.2020.109424

PubMed Abstract | CrossRef Full Text | Google Scholar

Lobel, L., Sigal, N., Borovok, I., Belitsky, B. R., Sonenshein, A. L., and Herskovits, A. A. (2015). The metabolic regulator CodY links Listeria monocytogenes metabolism to virulence by directly activating the virulence regulatory gene prfA. Mol. Microbiol. 95, 624–644. doi: 10.1111/mmi.12890

PubMed Abstract | CrossRef Full Text | Google Scholar

Loepfe, C., Raimann, E., Stephan, R., and Tasara, T. (2010). Reduced host cell invasiveness and oxidative stress tolerance in double and triple csp gene family deletion mutants of Listeria monocytogenes. Foodborne Path. Dis. 7, 775–783. doi: 10.1089/fpd.2009.0458

PubMed Abstract | CrossRef Full Text | Google Scholar

Lourenço, A., de Las Heras, A., Scortti, M., Vazquez-Boland, J., Frank, J. F., and Brito, L. (2013). Comparison of Listeria monocytogenes exoproteomes from biofilm and planktonic state: Lmo2504, a protein associated with biofilms. Appl. Environ. Microbiol. 79, 6075–6082. doi: 10.1128/AEM.01592-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Lundén, J. M., Autio, T. J., Markkula, A., Hellström, S., and Korkeala, H. (2003). Adaptive and cross-adaptive responses of persistent and nonpersistent Listeria monocytogenes strains to disinfectants. Intern. J. Food Microbiol. 82, 265–272. doi: 10.1016/s0168-1605(02)00312-4

CrossRef Full Text | Google Scholar

Lundén, J. M., Miettinen, M. K., Autio, T. J., and Korkeala, H. J. (2000). Persistent Listeria monocytogenes strains show enhanced adherence to food contact surface after short contact times. J. Food Prot. 63, 1204–1207. doi: 10.4315/0362-028x-63.9.1204

PubMed Abstract | CrossRef Full Text | Google Scholar

Magalhães, R., Ferreira, V., Brandão, T. R., Casquete Palencia, R. C., Almeida, G., and Teixeira, P. (2016). Persistent and non-persistent strains of Listeria monocytogenes: a focus on growth kinetics under different temperature, salt, and pH conditions and their sensitivity to sanitizers. Food Microbiol. 57, 103–108. doi: 10.1016/j.fm.2016.02.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Maherani, B., Hossain, F., Criado, P., Ben-Fadhel, Y., Salmieri, S., and Lacroix, M. (2016). World market development and consumer acceptance of irradiation technology. Foods 5:79. doi: 10.3390/foods5040079

PubMed Abstract | CrossRef Full Text | Google Scholar

Malley, T. J. V., Butts, J., and Wiedmann, M. (2015). Seek and destroy process: Listeria monocytogenes process controls in the ready-to-eat meat and poultry industry. J. Food Prot. 78, 436–445. doi: 10.4315/0362-028X.JFP-13-507

PubMed Abstract | CrossRef Full Text | Google Scholar

Margolles, A., Mayo, B., and de los Reyes-Gavilán, C. G. (2001). Susceptibility of Listeria monocytogenes and Listeria innocua strains isolated from short-ripened cheeses to some antibiotics and heavy metal salts. Food Microbiol. 18, 67–73. doi: 10.1006/fmic.2000.0377

CrossRef Full Text | Google Scholar

Martínez-Suárez, J. V., Ortiz, S., and López-Alonso, V. (2016). Potential impact of the resistance to quaternary ammonium disinfectants on the persistence of Listeria monocytogenes in food processing environments. Front. Microbiol. 7:638. doi: 10.3389/fmicb.2016.00638

PubMed Abstract | CrossRef Full Text | Google Scholar

Mastronicolis, S. K., Boura, A., Karaliota, A., Magiatis, P., Arvanitis, N., Litos, C., et al. (2006). Effect of cold temperature on the composition of different lipid classes of the foodborne pathogen Listeria monocytogenes: Focus on neutral lipids. Food Microbiol. 23, 184–194. doi: 10.1016/j.fm.2005.03.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Mata, M. M., da Silva, W. P., Wilson, R., Lowe, E., and Bowman, J. P. (2015). Attached and planktonic Listeria monocytogenes global proteomic responses and associated influence of strain genetics and temperature. J. Prot. Res. 14, 1161–1173. doi: 10.1021/pr501114e

PubMed Abstract | CrossRef Full Text | Google Scholar

Mata, M. T., Baquero, F., and Pérez-Díaz, J. C. (2000). A multidrug efflux transporter in Listeria monocytogenes. FEMS Microbiol. Lett. 15, 185–188. doi: 10.1016/S0378-1097(00)00199-3

CrossRef Full Text | Google Scholar

Matereke, L. T., and Okoh, A. I. (2020). Listeria monocytogenes virulence, antimicrobial resistance and environmental persistence: a review. Pathogens 9:528. doi: 10.3390/pathogens9070528

PubMed Abstract | CrossRef Full Text | Google Scholar

Maury, M. M., Bracq-Dieye, H., Huang, L., Vales, G., Lavina, M., Thouvenot, P., et al. (2019). Hypervirulent Listeria monocytogenes clones’ adapting to mammalian gut accounts for their association with dairy products. Nature Commun. 10:2488. doi: 10.1038/s41467-019-10380-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Mazaheri, T., Cervantes-Huamán, B. R. H., Bermúdez-Capdevila, M., Ripolles-Avila, C., and Rodríguez-Jerez, J. J. (2021). Listeria monocytogenes biofilms in the food industry: is the current hygiene program sufficient to combat the persistence of the pathogen? Microorganisms 9:181. doi: 10.3390/microorganisms9010181

PubMed Abstract | CrossRef Full Text | Google Scholar

McDonnell, G. E. (2017). “Antisepsis, disinfection, and sterilization,” in Types, Action, and Resistance, 2nd Edn, ed. G. E. McDonnel (Washington, DC: ASM Press), 432.

Google Scholar

McKinney, J. M., Williams, R. C., Boardman, G. D., Eifert, J. D., and Sumner, S. S. (2009a). Effect of acid stress, antibiotic resistance, and heat shock on the resistance of Listeria monocytogenes to UV light when suspended in distilled water and fresh brine. J. Food Prot. 72, 1634–1640. doi: 10.4315/0362-028x-72.8.1634

PubMed Abstract | CrossRef Full Text | Google Scholar

McKinney, J., Williams, R. C., Boardman, G. D., Eifert, J. D., and Sumner, S. S. (2009b). Dose of UV light required to inactivate Listeria monocytogenes in distilled water, fresh brine, and spent brine. J. Food Prot. 72, 2144–2150. doi: 10.4315/0362-028x-72.10.2144

PubMed Abstract | CrossRef Full Text | Google Scholar

McLaughlin, H. P., Hill, C., and Gahan, C. G. M. (2011). The impact of iron on Listeria monocytogenes; inside and outside the host. Curr. Opin. Biotechnol. 22, 194–199. doi: 10.1016/j.copbio.2010.10.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Meier, A. B., Guldimann, C., Markkula, A., Pöntinen, A., Korkeala, H., and Tasara, T. (2017). Comparative phenotypic and genotypic analysis of Swiss and Finnish Listeria monocytogenes isolates with respect to benzalkonium chloride resistance. Front. Microbiol. 8:397. doi: 10.3389/fmicb.2017.00397

PubMed Abstract | CrossRef Full Text | Google Scholar

Melo, J., Andrew, P. W., and Faleiro, M. L. (2015). Listeria monocytogenes in cheese and the dairy environment remains a food safety challenge: the role of stress responses. Food Res. Intern. 67, 75–90. doi: 10.1016/j.foodres.2014.10.031

CrossRef Full Text | Google Scholar

Midelet, G., Kobilinsky, A., and Carpentier, B. (2006). Construction and analysis of fractional multifactorial designs to study attachment strength and transfer of Listeria monocytogenes from pure or mixed biofilms after contact with a solid model food. Appl. Environ. Microbiol. 72, 2313–2321. doi: 10.1128/AEM.72.4.2313-2321.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Miguéis, S., Saraiva, C., and Esteves, A. (2017). Efficacy of LISTEX P100 at different concentrations for reduction of Listeria monocytogenes inoculated in sashimi. J. Food Prot. 80, 2094–2098. doi: 10.4315/0362-028X.JFP-17-098

PubMed Abstract | CrossRef Full Text | Google Scholar

Miladi, H., Elabed, H., Ben Slama, R., Rhim, A., and Bakhrouf, A. (2017). Molecular analysis of the role of osmolyte transporters opuCA and betL in Listeria monocytogenes after cold and freezing stress. Arch. Microbiol. 99, 259–265. doi: 10.1007/s00203-016-1300-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Miladi, H., Soukri, A., Bakhrouf, A., and Ammar, E. (2012). Expression of ferritin-like protein in Listeria monocytogenes after cold and freezing stress. Folia Microbiol. 57, 551–556. doi: 10.1007/s12223-012-0172-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, M. B., and Bassler, B. L. (2001). Quorum sensing in bacteria. Ann. Rev. Microbiol. 55, 165–199. doi: 10.1146/annurev.micro.55.1.165

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohamed, R. I., Abdelmonem, M. A., and Amin, H. M. (2018). Virulence and antimicrobial susceptibility profile of Listeria monocytogenes isolated from frozen vegetables available in the Egyptian market. Afr. J. Microbiol. Res. 12, 218–224. doi: 10.5897/AJMR2018.8794

CrossRef Full Text | Google Scholar

Møretrø, T., Schirmer, B. C. T., Heir, E., Fagerlund, A., Hjemli, P., and Langsrud, S. (2017). Tolerance to quaternary ammonium compound disinfectants may enhance growth of Listeria monocytogenes in the food industry. Intern. J. Food Microbiol. 241, 215–224. doi: 10.1016/j.ijfoodmicro.2016.10.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Moye, Z., Woolston, J., and Sulakvelidze, A. (2018). Bacteriophage applications for food production and processing. Viruses 10:205. doi: 10.3390/v10040205

PubMed Abstract | CrossRef Full Text | Google Scholar

Muchaamba, F., Stephan, R., and Tasara, T. (2021). Listeria monocytogenes cold shock proteins: Small proteins with a huge impact. Microorganisms 9:1061. doi: 10.3390/microorganisms9051061

PubMed Abstract | CrossRef Full Text | Google Scholar

Mullapudi, S., Siletzky, R. M., and Kathariou, S. (2008). Heavy-metal and benzalkonium chloride resistance of Listeria monocytogenes isolates from the environment of Turkey-processing plants. Appl. Environ. Microbiol. 74, 1464–1468. doi: 10.1128/AEM.02426-07

PubMed Abstract | CrossRef Full Text | Google Scholar

Müller, A., Rychli, K., Zaiser, A., Wieser, C., Wagner, M., and Schmitz-Esser, S. (2014). The Listeria monocytogenes transposon Tn6188 provides increased tolerance to various quaternary ammonium compounds and ethidium bromide. FEMS Microbiol. Lett. 361, 166–173. doi: 10.1111/1574-6968.12626

PubMed Abstract | CrossRef Full Text | Google Scholar

Naditz, A. L., Dzieciol, M., Wagner, M., and Schmitz-Esser, S. (2019). Plasmids contribute to food processing environment-associated stress survival in three Listeria monocytogenes ST121, ST8, and ST5 strains. Intern. J. Food Microbiol. 299, 39–46. doi: 10.1016/j.ijfoodmicro.2019.03.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Nair, S., Derré, I., Msadek, T., Gaillot, O., and Berche, P. (2000). CtsR controls class III heat shock gene expression in the human pathogen Listeria monocytogenes. Mol. Microbiol. 35, 800–811. doi: 10.1046/j.1365-2958.2000.01752.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Naitali, M., Dubois-Brissonnet, F., Cuvelier, G., and Bellon-Fontaine, M. N. (2009). Effects of pH and oil-in-water emulsions on growth and physicochemical cell surface properties of Listeria monocytogenes: Impact on tolerance to the bactericidal activity of disinfectants. Intern. J. Food Microbiol. 130, 101–107. doi: 10.1016/j.ijfoodmicro.2009.01.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Nelson, K. E., Fouts, D. E., Mongodin, E. F., Ravel, J., DeBoy, R. T., Kolonay, J. F., et al. (2004). Whole genome comparisons of serotype 4b and 1/2a strains of the food-borne pathogen Listeria monocytogenes reveal new insights into the core genome components of this species. Nucleic Acids Res. 32, 2386–2395. doi: 10.1093/nar/gkh562

PubMed Abstract | CrossRef Full Text | Google Scholar

Neunlist, M. R., Federighi, M., Laroche, M., Sohier, D., Delattre, G., Jacquet, C., et al. (2005). Cellular lipid fatty acid pattern heterogeneity between reference and recent food isolates of Listeria monocytogenes as a response to cold stress. Antonie van Leeuwenhoek 88, 199–206. doi: 10.1007/s10482-005-5412-7

PubMed Abstract | CrossRef Full Text | Google Scholar

NicAogáin, K., and O’Byrne, C. P. (2016). The role of stress and stress adaptations in determining the fate of the bacterial pathogen Listeria monocytogenes in the food chain. Front. Microbiol. 7:1865. doi: 10.3389/fmicb.2016.01865

PubMed Abstract | CrossRef Full Text | Google Scholar

Nightingale, K. K., Windham, K., Martin, K. E., Yeung, M., and Wiedmann, M. (2005). Select Listeria monocytogenes subtypes commonly found in foods carry distinct nonsense mutations in inlA, leading to expression of truncated and secreted internalin A, and are associated with a reduced invasion phenotype for human intestinal epithelial cells. Appl. Environ. Microbiol. 71, 8764–8772. doi: 10.1128/AEM.71.12.8764-8772.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Nilsson, R. E., Ross, T., and Bowman, J. P. (2011). Variability in biofilm production by Listeria monocytogenes correlated to strain origin and growth conditions. Intern. J. Food Microbiol. 150, 14–24. doi: 10.1016/j.ijfoodmicro.2011.07.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Niven, G. W., Miles, C. A., and Mackey, B. M. (1999). The effects of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145, 419–425. doi: 10.1099/13500872-145-2-419

PubMed Abstract | CrossRef Full Text | Google Scholar

Noll, M., Trunzer, K., Vondran, A., Vincze, S., Dieckmann, R., Al Dahouk, S., et al. (2020). Benzalkonium chloride induces a VBNC state in Listeria monocytogenes. Microorganisms 8:184. doi: 10.3390/microorganisms8020184

PubMed Abstract | CrossRef Full Text | Google Scholar

Norwood, D. E., and Gilmour, A. (2001). The differential adherence capabilities of two Listeria monocytogenes strains in monoculture and multispecies biofilms as a function of temperature. Lett. Appl. Microbiol. 33, 320–324. doi: 10.1046/j.1472-765X.2001.01004.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Nowak, J., Cruz, C. D., Palmer, J., Fletcher, G. C., and Flint, S. (2015). Biofilm formation of the L. monocytogenes strain 15G01 is influenced by changes in environmental conditions. J. Microbiol. Meth. 119, 189–195. doi: 10.1016/j.mimet.2015.10.022

PubMed Abstract | CrossRef Full Text | Google Scholar

O’Byrne, C. P., and Karatzas, K. A. (2008). The role of Sigma B (sB) in the stress adaptations of Listeria monocytogenes: overlaps between stress adaptation and virulence. Adv. Appl. Microbiol. 65, 115–140. doi: 10.1016/S0065-2164(08)00605-9

CrossRef Full Text | Google Scholar

O’Driscoll, B., Gahan, C. G., and Hill, C. (1996). Adaptive acid tolerance response in Listeria monocytogenes: isolation of an acid-tolerant mutant which demonstrates increased virulence. Appl. Environ. Microbiol. 62, 1693–1698. doi: 10.1128/aem.62.5.1693-1698.1996

PubMed Abstract | CrossRef Full Text | Google Scholar

Okada, Y., Makino, S., Okada, N., Asakura, H., Yamamoto, S., and Igimi, S. (2008). Identification and analysis of the osmotolerance associated genes in Listeria monocytogenes. Food Add. Contamin. 25, 1089–1094. doi: 10.1080/02652030802056634

PubMed Abstract | CrossRef Full Text | Google Scholar

Ollinger, J., Bowen, B., Wiedmann, M., Boor, K. J., and Bergholz, T. M. (2009). Listeria monocytogenes sigmaB modulates PrfA-mediated virulence factor expression. Infect. Immun. 77, 2113–2124. doi: 10.1128/IAI.01205-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Ortiz, S., López, V., and Martínez-Suárez, J. V. (2014). The influence of subminimal inhibitory concentrations of benzalkonium chloride on biofilm formation by Listeria monocytogenes. Intern. J. Food Microbiol. 189, 106–112.

Google Scholar

Ortiz, S., López-Alonso, V., Rodríguez, P., and Martínez-Suárez, J. V. (2016). The connection between persistent, disinfectant-resistant Listeria monocytogenes strains from two geographically separate Iberian pork processing plants: Evidence from comparative genome analysis. Appl. Environ. Microbiol. 82, 308–317. doi: 10.1128/AEM.02824-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Padan, E., Bibi, E., Ito, M., and Krulwich, T. A. (2005). Alkaline pH homeostasis in bacteria: new insights. Bioch. Bioph. Acta 1717, 67–88. doi: 10.1016/j.bbamem.2005.09.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Palaiodimou, L., Fanning, S., and Fox, E. M. (2021). Genomic insights into persistence of Listeria species in the food processing environment. J. Appl. Microbiol. 131, 2082–2094. doi: 10.1111/jam.15089

PubMed Abstract | CrossRef Full Text | Google Scholar

Pan, Y., Breidt, F., and Kathariou, S. (2006). Resistance of Listeria monocytogenes biofilms to sanitizing agents in a simulated food processing environment. Appl. Environ. Microbiol. 72, 7711–7717. doi: 10.1128/AEM.01065-06

PubMed Abstract | CrossRef Full Text | Google Scholar

Panebianco, F., Rubiola, S., Chiesa, F., Civera, T., and Di Ciccio, P. A. (2021). Effect of gaseous ozone on Listeria monocytogenes planktonic cells and biofilm: an in vitro study. Foods 10:1484. doi: 10.3390/foods10071484

PubMed Abstract | CrossRef Full Text | Google Scholar

Parsons, C., Lee, S., and Kathariou, S. (2019). Heavy metal resistance determinants of the foodborne pathogen Listeria monocytogenes. Genes 10:11. doi: 10.3390/genes10010011

PubMed Abstract | CrossRef Full Text | Google Scholar

Parsons, C., Lee, S., and Kathariou, S. (2020). Dissemination and conservation of cadmium and arsenic resistance determinants in Listeria and other Gram-positive bacteria. Mol. Microbiol. 113, 560–569. doi: 10.1111/mmi.14470

PubMed Abstract | CrossRef Full Text | Google Scholar

Perera, M. N., Abuladze, T., Li, M., Woolston, J., and Sulakvelidze, A. (2015). Bacteriophage cocktail significantly reduces or eliminates Listeria monocytogenes contamination on lettuce, apples, cheese, smoked salmon and frozen foods. Food Microbiol. 52, 42–48. doi: 10.1016/j.fm.2015.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Phadtare, S., Alsina, J., and Inouye, M. (1999). Cold-shock response and cold-shock proteins. Curr. Opin. Microbiol. 2, 175–180. doi: 10.1016/S1369-5274(99)80031-9

CrossRef Full Text | Google Scholar

Phan-Thanh, L., Mahouin, F., and Aligé, S. (2000). Acid responses of Listeria monocytogenes. Intern. J. Food Microbiol. 55, 121–126. doi: 10.1016/s0168-1605(00)00167-7

CrossRef Full Text | Google Scholar

Pieta, L. Brusch Garcia, F. Pelicioli Riboldi, G. Abruzzi de Oliveira, L. Guedes Frazzon, A. P., and Frazzon, J. (2014). Transcriptional analysis of genes related to biofilm formation, stress-response, and virulence in Listeria monocytogenes strains grown at different temperatures. Ann. Microbiol. 64, 1707–1714. doi: 10.1007/s13213-014-0814-2

CrossRef Full Text | Google Scholar

Połaska, M., and Sokołowska, B. (2019). Bacteriophages - A new hope or a huge problem in the food industry. AIMS Microbiol. 5, 324–346. doi: 10.3934/microbiol.2019.4.324

PubMed Abstract | CrossRef Full Text | Google Scholar

Pöntinen, A., Markkula, A., Lindstrom, M., and Korkeala, H. (2015). Two-component-system histidine kinases involved in growth of Listeria monocytogenes EGD-e at low temperatures. Appl. Environ. Microbiol. 81, 3994–4004. doi: 10.1128/AEM.00626-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Ranasinghe, R. A. S. S., Satharasinghe, D. A., Tang, J. Y. H., Rukayadi, Y., Radu, K. R., New, C. Y., et al. (2021). Persistence of Listeria monocytogenes in food commodities: foodborne pathogenesis, virulence factors, and implications for public health. Food Res. 5, 1–16. doi: 10.26656/fr.2017.5(1).199

CrossRef Full Text | Google Scholar

Rand, J. L., Hofmann, R., Alam, M. Z. B., Chauret, C., Cantwell, R., Andrews, R. C., et al. (2007). A field study evaluation for mitigating biofouling with chlorine dioxide or chlorine integrated with UV disinfection. Water Res. 41, 1939–1948. doi: 10.1016/j.watres.2007.02.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Rastogi, R. P., Richa, A., Kumar, A., Tyagi, M. B., and Singha, R. P. (2010). Molecular mechanism of ultraviolet radiation-induced DNA damage and repair. J. Nucleic Acids 2010:592980. doi: 10.4061/2010/592980

PubMed Abstract | CrossRef Full Text | Google Scholar

Ratani, S. S., Siletzky, R. M., Dutta, V., Yildirim, S., Osborne, J., Lin, W., et al. (2012). Heavy metal and disinfectant resistance of Listeria monocytogenes from foods and food processing plants. Appl. Environ. Microbiol. 78, 6938–6945. doi: 10.1128/AEM.01553-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Ricci, A., Alinovi, M., Martelli, F., Bernini, V., Garofalo, A., Perna, G., et al. (2021). Heat resistance of Listeria monocytogenes in dairy matrices involved in Mozzarella di Bufala Campana PDO cheese. Front. Microbiol. 11:581934. doi: 10.3389/fmicb.2020.581934

PubMed Abstract | CrossRef Full Text | Google Scholar

Riedel, C. U., Monk, I. R., Casey, P. G., Waidmann, M. S., Gahan, C. G. M., and Hill, C. (2009). AgrD-dependent quorum sensing affects biofilm formation, invasion, virulence and global gene expression profiles in Listeria monocytogenes. Mol. Microbiol. 71, 1177–1189.

Google Scholar

Rieu, A., Weidmann, S., Garmyn, D., Piveteau, P., and Guzzo, J. (2007). Agr system of Listeria monocytogenes EGD-e: role in adherence and differential expression pattern. Appl. Environ. Microbiol. 73, 6125–6133.

Google Scholar

Rodríguez-Campos, D., Rodríguez-Melcón, C., Alonso-Calleja, C., and Capita, R. (2019). Persistent Listeria monocytogenes isolates from a poultry-processing facility form more biofilm but do not have a greater resistance to disinfectants than sporadic strains. Pathogens 8:250. doi: 10.3390/pathogens8040250

PubMed Abstract | CrossRef Full Text | Google Scholar

Rodríguez-López, P., Rodríguez-Herrera, J. J., Vázquez-Sánchez, D., and López Cabo, M. (2018). Current knowledge on Listeria monocytogenes biofilms in food-related environments: incidence, resistance to biocides, ecology and biocontrol. Foods 7:85. doi: 10.3390/foods7060085

PubMed Abstract | CrossRef Full Text | Google Scholar

Roncarati, D., and Scarlato, V. (2017). Regulation of heat-shock genes in bacteria: from signal sensing to gene expression output. FEMS Microbiol. Rev. 41, 549–574. doi: 10.1093/femsre/fux015

PubMed Abstract | CrossRef Full Text | Google Scholar

Rosen, B. P. (1999). Families of arsenic transporters. Trends Microbiol. 7, 207–212. doi: 10.1016/s0966-842x(99)01494-8

CrossRef Full Text | Google Scholar

Rothrock, M. J. Jr., Micciche, A. C., Bodie, A. R., and Ricke, S. C. (2019). Listeria occurrence and potential control strategies in alternative and conventional poultry processing and retail. Front. Sustain. Food Syst. 3:33. doi: 10.3389/fsufs.2019.00033

CrossRef Full Text | Google Scholar

Ruckerl, I., Muhterem-Uyar, M., Muri-Klinger, S., Wagner, K. H., Wagner, M., and Stessl, B. (2014). L. monocytogenes in a cheese processing facility: learning from contamination scenarios over three years of sampling. Intern. J. Food Microbiol. 189, 98–105. doi: 10.1016/j.ijfoodmicro.2014.08.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, S., Begley, M., Gahan, C. G., and Hill, C. (2009). Molecular characterization of the arginine deiminase system in Listeria monocytogenes: regulation and role in acid tolerance. Environ. Microbiol. 11, 432–445. doi: 10.1111/j.1462-2920.2008.01782.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, S., Begley, M., Hill, C., and Gahan, C. G. (2010). A five-gene stress survival islet (SSI-1) that contributes to the growth of Listeria monocytogenes in suboptimal conditions. J. Appl. Microbiol. 109, 984–995. doi: 10.1111/j.1365-2672.2010.04726.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, S., Hill, C., and Gahan, C. G. (2008). Acid stress responses in Listeria monocytogenes. Adv. Appl. Microbiol. 65, 67–91. doi: 10.1016/S0065-2164(08)00603-5

CrossRef Full Text | Google Scholar

Sadekuzzaman, M., Yang, S., Mizan, F. R., Kim, H. S., and Ha, S. D. (2017). Effectiveness of a phage cocktail as a biocontrol agent against L. monocytogenes biofilms. Food Control 78, 256–263. doi: 10.1016/j.foodcont.2016.10.056

CrossRef Full Text | Google Scholar

Santos, T., Viala, D., Chambon, C., Esbelin, J., and Hébraud, M. (2019). Listeria monocytogenes biofilm adaptation to different temperatures seen through shotgun proteomics. Front. Nutr. 6:89. doi: 10.3389/fnut.2019.00089

PubMed Abstract | CrossRef Full Text | Google Scholar

Scallan, E., Hoekstra, B. M., Angulo, F. J., Tauxe, R. V., Widdowson, M. A., Roy, S. L., et al. (2011). Foodborne illness acquired in the United States – major pathogens. Emerg. Infect. Dis. 17, 7–15. doi: 10.3201/eid1701.P11101

PubMed Abstract | CrossRef Full Text | Google Scholar

Schärer, K., Stephan, R., and Tasara, T. (2013). Cold shock proteins contribute to the regulation of listeriolysin O production in Listeria monocytogenes. Foodborne Path. Dis. 10, 1023–1029. doi: 10.1089/fpd.2013.1562

PubMed Abstract | CrossRef Full Text | Google Scholar

Schirmer, E. C., Glover, J. R., Singer, M. A., and Lindquist, S. (1996). HSP100/Clp proteins: a common mechanism explains diverse functions. Trends Biochem. Sci. 21, 289–296. doi: 10.1016/S0968-0004(96)10038-4

CrossRef Full Text | Google Scholar

Schmid, B., Klumpp, J., Raimann, E., Loessner, M. J., Stephan, R., and Tasara, T. (2009). Role of cold shock proteins in growth of Listeria monocytogenes under cold and osmotic stress conditions. Appl. Environ. Microbiol. 75, 1621–1627. doi: 10.1128/AEM.02154-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwab, U., Hu, Y., Wiedmann, M., and Boor, K. J. (2005). Alternative sigma factor B is not essential for Listeria monocytogenes surface attachment. J. Food Prot. 68, 311–317. doi: 10.4315/0362-028x-68.2.311

PubMed Abstract | CrossRef Full Text | Google Scholar

Sela, S., Frank, S., Belausov, E., and Pinto, R. (2006). A mutation in the luxS gene influences Listeria monocytogenes biofilm formation. Appl. Environ. Microbiol. 72, 5653–5658. doi: 10.1128/AEM.00048-06

PubMed Abstract | CrossRef Full Text | Google Scholar

Shabala, L., Ross, T., McMeekin, T., and Shabala, S. (2006). Non-invasive microelectrode ion flux measurements to study adaptive responses of microorganisms to the environment. FEMS Microbiol. Rev. 30, 472–486. doi: 10.1111/j.1574-6976.2006.00019.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Shen, Q., Jangam, P. M., Soni, K. A., Nannapaneni, R., Schilling, W., and Silva, J. L. (2014). Low, medium, and high heat tolerant strains of Listeria monocytogenes and increased heat stress resistance after exposure to sublethal heat. J. Food Prot. 77, 1298–1307. doi: 10.4315/0362-028x.jfp-13-423

PubMed Abstract | CrossRef Full Text | Google Scholar

Shen, Q., Pandare, P., Soni, K. A., Nannapaneni, R., Mahmoud, B. S. M., and Sharma, C. S. (2016). Influence of temperature on alkali stress adaptation in Listeria monocytogenes. Food Control 62, 74–80. doi: 10.1016/j.foodcont.2015.10.005

CrossRef Full Text | Google Scholar

Singh, A. K., Ulanov, A. V., Li, Z., Jayaswal, R. K., and Wilkinson, B. J. (2011). Metabolomes of the psychrotolerant bacterium Listeria monocytogenes 10403S grown at 37°C and 8°C. Intern. J. Food Microbiol. 148, 107–114.

Google Scholar

Sinha, R. P., and Häder, D. P. (2002). UV-induced DNA damage and repair: a review. Photochem. Photobiol. Sci. 1, 225–236. doi: 10.1039/b201230h

PubMed Abstract | CrossRef Full Text | Google Scholar

Slama, R. B., Bekir, K., Miladi, H., Noumi, A., and Bakhrouf, A. (2012). Adhesive ability and biofilm metabolic activity of Listeria monocytogenes strains before and after cold stress. Afr. J. Biotechnol. 11, 12475–12482. doi: 10.5897/AJB11.3939

CrossRef Full Text | Google Scholar

Sleator, R. D., Gahan, C. G. M., and Hill, C. (2003). A postgenomic appraisal of osmotolerance in Listeria monocytogenes. Appl. Environ. Microbiol. 69, 1–9. doi: 10.1128/AEM.69.1.1-9.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Smelt, J. P. P. M., and Brul, S. (2014). Thermal inactivation of microorganisms. Crit. Rev. Food Sci. Nutr. 54, 1371–1385. doi: 10.1080/10408398.2011.637645

PubMed Abstract | CrossRef Full Text | Google Scholar

Smith, J. L., Liu, Y., and Paoli, G. C. (2013). How does Listeria monocytogenes combat acid conditions? Can. J. Microbiol. 59, 141–152. doi: 10.1139/cjm-2012-0392

PubMed Abstract | CrossRef Full Text | Google Scholar

Soares, C. A., and Knuckley, B. (2016). Mechanistic studies of the agmatine deiminase from Listeria monocytogenes. Biochem. J. 473, 1553–1561. doi: 10.1042/BCJ20160221

PubMed Abstract | CrossRef Full Text | Google Scholar

Sohlenkamp, C., and Geiger, O. (2016). Bacterial membrane lipids: diversity in structures and pathways. FEMS Microbiol. Rev. 40, 133–159. doi: 10.1093/femsre/fuv008

PubMed Abstract | CrossRef Full Text | Google Scholar

Soni, K. A., and Nannapaneni, R. (2010). Removal of Listeria monocytogenes biofilms with bacteriophage P100. J. Food Prot. 73, 1519–1524. doi: 10.4315/0362-028X-73.8.1519

PubMed Abstract | CrossRef Full Text | Google Scholar

Soni, K. A., Nannapaneni, R., and Tasara, T. (2011). The contribution of transcriptomic and proteomic analysis in elucidating stress adaptation responses of Listeria monocytogenes. Foodborne Path. Dis. 8, 842–852. doi: 10.1089/fpd.2010.0746

PubMed Abstract | CrossRef Full Text | Google Scholar

Sörqvist, S. (1994). Heat resistance of different serovars of Listeria monocytogenes. J. Appl. Bacteriol. 76, 383–388. doi: 10.1111/j.1365-2672.1994.tb01644.x

CrossRef Full Text | Google Scholar

Soumet, C., Ragimbeau, C., and Maris, P. (2005). Screening of benzalkonium chloride resistance in Listeria monocytogenes strains isolated during cold smoked fish production. Lett. Appl. Microbiol. 41, 291–296. doi: 10.1111/j.1472-765X.2005.01763.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Stoll, R., and Goebel, W. (2010). The major PEP-phosphotransferase systems (PTSs) for glucose, mannose and cellobiose of Listeria monocytogenes, and their significance for extra- and intracellular growth. Microbiology 156, 1069–1083. doi: 10.1099/mic.0.034934-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Suutari, M., and Laakso, S. (1994). Microbial fatty acids and thermal adaptation. Crit. Rev. Microbiol. 20, 285–328. doi: 10.3109/10408419409113560

PubMed Abstract | CrossRef Full Text | Google Scholar

Takahashi, H., Miya, S., Igarashi, K., Suda, T., Kuramoto, S., and Kimura, B. (2009). Biofilm formation ability of Listeria monocytogenes isolates from raw ready-to-eat seafood. J. Food Prot. 72, 1476–1480. doi: 10.4315/0362-028X-72.7.1476

PubMed Abstract | CrossRef Full Text | Google Scholar

Taormina, P. J., and Beuchat, L. R. (2001). Survival and heat resistance of Listeria monocytogenes after exposure to alkali and chlorine. Appl. Environ. Microbiol. 67, 2555–2563.

Google Scholar

Tasara, T., and Stephan, R. (2006). Cold stress tolerance of Listeria monocytogenes: a review of molecular adaptive mechanisms and food safety implications. J. Food Prot. 69, 1473–1484. doi: 10.4315/0362-028X-69.6.1437

CrossRef Full Text | Google Scholar

Taylor, A. J., and Stasiewicz, M. J. (2019). Persistent and sporadic Listeria monocytogenes strains do not differ when growing at 37°C, in planktonic state, under different food associated stresses or energy sources. BMC Microbiol. 19:257. doi: 10.1186/s12866-019-1631-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Tezel, U., and Pavlostathis, S. G. (2015). Quaternary ammonium disinfectants: Microbial adaptation, degradation and ecology. Curr. Opin. Biotechnol. 33, 296–304. doi: 10.1016/j.copbio.2015.03.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Tisa, L. S., and Rosen, B. P. (1990). Molecular characterization of an anion pump. The ArsB protein is the membrane anchor for the ArsA protein. J. Biol. Chem. 265, 190–194.

Google Scholar

Todhanakasem, T., and Young, G. M. (2008). Loss of flagellum-based motility by Listeria monocytogenes results in formation of hyperbiofilms. J. Bacteriol. 190, 6030–6034. doi: 10.1128/JB.00155-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Tomičić, R. M., Čabarkapa, I. S., Vukmirović, D. M., Lević, J. D., and Tomičić, Z. M. (2016). Influence of growth conditions on biofilm formation of Listeria monocytogenes. Food Feed Res. 43, 19–24. doi: 10.5937/FFR1601019T

CrossRef Full Text | Google Scholar

Tomoyasu, T., Mogk, A., Langen, H., Goloubinoff, P., and Bukau, B. (2001). Genetic dissection of the roles of chaperones and proteases in protein folding and degradation in the Escherichia coli cytosol. Mol. Microbiol. 40, 397–413. doi: 10.1046/j.1365-2958.2001.02383.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Tresse, O., Lebret, V., Garmyn, D., and Dussurget, O. (2009). The impact of growth history and flagellation on the adhesion of various Listeria monocytogenes strains to polystyrene. Can. J. Microbiol. 55, 189–196. doi: 10.1139/W08-114

PubMed Abstract | CrossRef Full Text | Google Scholar

Tresse, O., Shannon, K., Pinon, A., Malle, P., Vialette, M., and Midelet-Bourdin, G. (2007). Variable adhesion of Listeria monocytogenes isolates from food-processing facilities and clinical cases to inert surfaces. J. Food Prot. 70, 1569–1578. doi: 10.4315/0362-028x-70.7.1569

PubMed Abstract | CrossRef Full Text | Google Scholar

Uesugi, A. R., Hsu, L. C., Worobo, R. W., and Moraru, C. I. (2016). Gene expression analysis for Listeria monocytogenes following exposure to pulsed light and continuous ultraviolet light treatments. Food Sci. Technol. 68, 579–588. doi: 10.1016/j.lwt.2016.01.007

CrossRef Full Text | Google Scholar

Unrath, N., McCabe, E., Macori, G., and Fanning, S. (2021). Application of whole genome sequencing to aid in deciphering the persistence potential of Listeria monocytogenes in food production environments. Microorganisms 9:1856. doi: 10.3390/microorganisms9091856

PubMed Abstract | CrossRef Full Text | Google Scholar

Utratna, M., Cosgrave, E., Baustian, C., Ceredig, R. H., and O’Byrne, C. P. (2014). Effects of growth phase and temperature on σB activity within a Listeria monocytogenes population: evidence for RsbV-independent activation of σB at refrigeration temperatures. BioMed Res. Intern. 2014:641647. doi: 10.1155/2014/641647

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaid, R., Linton, R. H., and Morgan, M. T. (2010). Comparison of inactivation of Listeria monocytogenes within a biofilm matrix using chlorine dioxide gas, aqueous chlorine dioxide and sodium hypochlorite treatments. Food Microbiol. 27, 979–984. doi: 10.1016/j.fm.2010.05.024

PubMed Abstract | CrossRef Full Text | Google Scholar

Valdramidis, V. P., Patterson, M. F., and Linton, M. (2015). Modelling the recovery of Listeria monocytogenes in high pressure processed simulated cured meat. Food Control 47, 353–358. doi: 10.1016/j.foodcont.2014.07.022

CrossRef Full Text | Google Scholar

Van Boeijen, I. K. H., Chavaroche, A. A. E., Valderrama, W. B., Moezelaar, R., Zwietering, M. H., and Abee, T. (2010). Population diversity of Listeria monocytogenes LO28: phenotypic and genotypic characterization of variants resistant to high hydrostatic pressure. Appl. Environ. Microbiol. 76, 2225–2233. doi: 10.1128/AEM.02434-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Van der Veen, S., Hain, T., Wouters, J. A., Hossain, H., de Vos, W. M., Abee, T., et al. (2007). The heat-shock response of Listeria monocytogenes comprises genes involved in heat shock, cell division, cell wall synthesis, and the SOS response. Microbiology 153, 3593–3607. doi: 10.1099/mic.0.2007/006361-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Walker, S. J., Archer, P., and Banks, J. G. (1990). Growth of Listeria monocytogenes at refrigeration temperatures. J. Appl. Bacteriol. 68, 157–162. doi: 10.1111/j.1365-2672.1990.tb02561.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Watkins, J., and Sleath, K. P. (1981). Isolation and enumeration of Listeria monocytogenes from sewage, sewage sludge and river water. J. Appl. Bacteriol. 50, 1–9. doi: 10.1111/j.1365-2672.1981.tb00865.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, C. I., Cook, D. L., and Kirk, J. R. (1985). Use of chlorine compounds in the food industry. Food Technol. 39, 107–115.

Google Scholar

Wemekamp-Kamphuis, H. H., Wouters, J., Leeuw, P. P. L., De Hain, T., Chakraborty, T., and Abee, T. (2004). Identification of sigma factor B-controlled genes and their impact on acid stress, high hydrostatic pressure, and freeze survival in Listeria monocytogenes EGD-e. Appl. Environ. Microbiol. 70, 3457–3466. doi: 10.1128/AEM.70.6.3457-3466.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Wen, J., Deng, X., Li, Z., Dudley, E. G., Anantheswaran, R. C., Knabel, S. J., et al. (2011). Transcriptomic response of Listeria monocytogenes during the transition to the long-term-survival phase. Appl. Environ. Microbiol. 77, 5966–5972. doi: 10.1128/AEM.00596-11

PubMed Abstract | CrossRef Full Text | Google Scholar

Wiktorczyk-Kapischke, N., Skowron, K., Grudlewska-Buda, K., Wałecka-Zacharska, E., Korkus, J., and Gospodarek-Komkowska, E. (2021). Adaptive response of Listeria monocytogenes to the stress factors in the food processing environment. Front. Microbiol. 12:710085. doi: 10.3389/fmicb.2021.710085

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, J., and Rosen, B. P. (1993). The arsD gene encodes a second trans-acting regulatory protein of the plasmid-encoded arsenical resistance operon. Mol. Microbiol. 8, 615–623. doi: 10.1111/j.1365-2958.1993.tb01605.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, T., Jiang, X., Zhang, Y., Ji, S., Gao, W., and Shi, L. (2018). Effect of benzalkonium chloride adaptation on sensitivity to antimicrobial agents and tolerance to environmental stresses in Listeria monocytogenes. Front. Microbiol. 9:2906. doi: 10.3389/fmicb.2018.02906

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeisel, S. H., Mar, M. H., Howe, J. C., and Holden, J. M. (2003). Concentrations of choline-containing compounds and betaine in common foods. J. Nutr. 133, 1302–1307. doi: 10.1093/jn/133.5.1302

PubMed Abstract | CrossRef Full Text | Google Scholar

Zetzmann, M., Sánchez-Kopper, A., Waidmann, M. S., Blombach, B., and Riedel, C. U. (2016). Identification of the agr peptide of Listeria monocytogenes. Front. Microbiol. 7:989. doi: 10.3389/fmicb.2016.00989

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X., Wang, S., Chen, X., and Qu, C. (2021). Review controlling Listeria monocytogenes in ready-to-eat meat and poultry products: An overview of outbreaks, current legislations, challenges, and future prospects. Trends Food Sci. Technol. 116, 24–35. doi: 10.1016/j.tifs.2021.07.014

CrossRef Full Text | Google Scholar

Zhang, Z., Meng, Q., Qiao, J., Yang, L., Cai, X., Wang, G., et al. (2013). RsbV of Listeria monocytogenes contributes to regulation of environmental stress and virulence. Arch. Microbiol. 195, 113–120. doi: 10.1007/s00203-012-0855-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Zinchenko, A. A., Sergeyev, V. G., Yamabe, K., Murata, S., and Yoshikawa, K. (2004). DNA compaction by divalent cations: structural specificity revealed by the potentiality of designed quaternary diammonium salts. Chembiochem 5, 360–386. doi: 10.1002/cbic.200300797

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Listeria monocytogenes, stress response, food processing, bacterial persistence, molecular mechanisms, food safety

Citation: Osek J, Lachtara B and Wieczorek K (2022) Listeria monocytogenes – How This Pathogen Survives in Food-Production Environments? Front. Microbiol. 13:866462. doi: 10.3389/fmicb.2022.866462

Received: 31 January 2022; Accepted: 04 April 2022;
Published: 26 April 2022.

Edited by:

Min Suk Rhee, Korea University, South Korea

Reviewed by:

Roger Stephan, University of Zurich, Switzerland
Dara Leong, Teagasc, Ireland
Sangmi Lee, Chungbuk National University, South Korea

Copyright © 2022 Osek, Lachtara and Wieczorek. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jacek Osek, josek@piwet.pulawy.pl

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.