ArticlePDF AvailableLiterature Review

Listeria monocytogenes: review of pathogenesis and virulence determinants-targeted immunological assays

Authors:

Abstract and Figures

Listeria monocytogenes is one of the most invasive foodborne pathogens and is responsible for numerous outbreaks worldwide. Most of the methods to detect this bacterium in food require selective enrichment using traditional bacterial culture techniques that can be time-consuming and labour-intensive. Moreover, molecular methods are expensive and need specific technical knowledge. In contrast, immunological approaches are faster, simpler, and user-friendly alternatives and have been developed for the detection of L. monocytogenes in food, environmental, and clinical samples. These techniques are dependent on the constitutive expression of L. monocytogenes antigens and the specificity of the antibodies used. Here, updated knowledge on pathogenesis and the key immunogenic virulence determinants of L. monocytogenes that are used for the generation of monoclonal and polyclonal antibodies for the serological assay development are summarised. In addition, immunological approaches based on enzyme-linked immunosorbent assay, immunofluorescence, lateral flow immunochromatographic assays, and immunosensors with relevant improvements are highlighted. Though the sensitivity and specificity of the assays were improved significantly, methods still face many challenges that require further validation before use.
Content may be subject to copyright.
Full Terms & Conditions of access and use can be found at
https://www.tandfonline.com/action/journalInformation?journalCode=imby20
Critical Reviews in Microbiology
ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/imby20
Listeria monocytogenes: review of pathogenesis and
virulence determinants-targeted immunological
assays
Leonardo Lopes-Luz, Marcelo Mendonça, Matheus Bernardes Fogaça, André
Kipnis, Arun K. Bhunia & Samira Bührer-Sékula
To cite this article: Leonardo Lopes-Luz, Marcelo Mendonça, Matheus Bernardes Fogaça,
André Kipnis, Arun K. Bhunia & Samira Bührer-Sékula (2021): Listeria�monocytogenes: review
of pathogenesis and virulence determinants-targeted immunological assays, Critical Reviews in
Microbiology, DOI: 10.1080/1040841X.2021.1911930
To link to this article: https://doi.org/10.1080/1040841X.2021.1911930
Published online: 24 Apr 2021.
Submit your article to this journal
View related articles
View Crossmark data
REVIEW ARTICLE
Listeria monocytogenes: review of pathogenesis and virulence
determinants-targeted immunological assays
Leonardo Lopes-Luz
a
, Marcelo Mendonc¸a
b
, Matheus Bernardes Fogac¸a
a
, Andr
e Kipnis
a
, Arun K. Bhunia
c,d,e
and Samira B
uhrer-S
ekula
a
a
Instituto de Patologia Tropical e Sa
ude P
ublica, Universidade Federal de Goi
as, Goi^
ania, Brasil;
b
Curso de Medicina Veterin
aria,
Universidade Federal do Agreste de Pernambuco, Garanhuns, Brasil;
c
Department of Food Science, Purdue University, West Lafayette,
IN, USA;
d
Department of Comparative Pathobiology, Purdue University, West Lafayette, IN, USA;
e
Purdue Institute of Inflammation,
Immunology and Infectious Disease, Purdue University, West Lafayette, IN, USA
ABSTRACT
Listeria monocytogenes is one of the most invasive foodborne pathogens and is responsible for
numerous outbreaks worldwide. Most of the methods to detect this bacterium in food require
selective enrichment using traditional bacterial culture techniques that can be time-consuming
and labour-intensive. Moreover, molecular methods are expensive and need specific technical
knowledge. In contrast, immunological approaches are faster, simpler, and user-friendly alterna-
tives and have been developed for the detection of L. monocytogenes in food, environmental,
and clinical samples. These techniques are dependent on the constitutive expression of L. mono-
cytogenes antigens and the specificity of the antibodies used. Here, updated knowledge on
pathogenesis and the key immunogenic virulence determinants of L. monocytogenes that are
used for the generation of monoclonal and polyclonal antibodies for the serological assay devel-
opment are summarised. In addition, immunological approaches based on enzyme-linked
immunosorbent assay, immunofluorescence, lateral flow immunochromatographic assays, and
immunosensors with relevant improvements are highlighted. Though the sensitivity and specifi-
city of the assays were improved significantly, methods still face many challenges that require
further validation before use.
ARTICLE HISTORY
Received 27 August 2020
Revised 19 March 2021
Accepted 29 March 2021
Published online 20 April
2021
KEYWORDS
Listeria monocytogenes;
pathogenesis; listeriosis;
virulence; antibodies;
antigens; serological
methods; lateral flow
immunoassay; detection
1. Introduction
Listeria monocytogenes is a Gram-positive, invasive,
foodborne bacterial pathogen that causes listeriosis, a
systemic disease resulting from the ingestion of conta-
minated food. The systemic spread of the pathogen
from the gastrointestinal tract depends on its ability to
cross intestinal, blood-brain, and placental barriers
(Radoshevich and Cossart 2018; Drolia and Bhunia
2019). Localised infection in the gastrointestinal tract
may lead to the onset of a rare form of gastroenteritis
characterised by diarrhoea, abdominal cramp, and flu-
like symptoms in healthy adults, whereas the systemic
invasive disease is manifested by fever, headache, men-
ingitis, encephalitis, liver abscess, abortion, premature
birth, stillbirth, and neonatal infection with sepsis and
pneumonia in immunocompromised hosts (Drolia and
Bhunia 2019; Schlech 2019). The mortality rate in
immunocompromised hosts including the elderly, preg-
nant women, and their foetuses, neonates, cancer
patients, and chemotherapy recipients is 2030%
(Swaminathan and Gerner-Smidt 2007). In particular,
pregnant women and their foetuses are highly suscep-
tible to listeriosis and 1627% of all L. monocytogenes
infections are reported in pregnant women (Lamont
et al. 2011; Craig et al. 2019). A recent study suggested
that L. monocytogenes infection in the first trimester
presents the greater risk of fetal loss than in the third
trimester (Wolfe 2017). Contaminated ready-to-eat-
foods such as hotdogs, soft cheeses mainly made with
unpasteurised milk, smoked fish, ice cream, p^
at
e,
polony, cantaloupe, apple, and vegetables were impli-
cated to be the primary vehicle for transmission
(Buchanan et al. 2017; Smith et al. 2019). The infectious
dose varies widely (<100 cells to 10
11
cells) and
depends on the amount of contaminated food con-
sumed, the bacterial strain, and the immunological sta-
tus of the host (Pouillot et al. 2016; Buchanan
et al. 2017).
The genus Listeria has 20 species and is subdivided
into two major groups: (i) Listeria sensu stricto
CONTACT Arun K. Bhunia bhunia@purdue.edu Department of Food Science, Purdue University, West Lafayette, IN, USA
ß2021 Informa UK Limited, trading as Taylor & Francis Group
CRITICAL REVIEWS IN MICROBIOLOGY
https://doi.org/10.1080/1040841X.2021.1911930
comprising of the species L. monocytogenes, L. innocua,
L. seeligeri, L. ivanovii, L. welshimeri and L. marthii and
(ii) Listeria sensu lato including L. grayi, L. fleischmannii,
L. weihenstephanensis, L. rocourtiae, L. floridensis, L. cor-
nellensis, L. riparia. L. aquatica, L. grandensis, L. newyor-
kensis, L. goaensis, L. booriae, L. costaricensis and L.
thailandensis (N
u~
nez-Montero et al. 2018; Leclercq et al.
2019). Only L. monocytogenes and L. ivanovii are consid-
ered pathogenic infecting humans and animals,
respectively. L. monocytogenes species can be further
subdivided into four lineages (I-IV), with lineage sero-
types I and II (1/2a, 1/2b, and 4b) involved in most lis-
teriosis cases reported worldwide (Weller et al. 2015).
The dissemination of L. monocytogenes through
foods is facilitated by the fact that the pathogen is
resistant and adaptable to adverse environmental con-
ditions such as high salt concentration (Gandhi and
Chikindas 2007), broad temperature range (145 C),
low pH (Tasara and Stephan 2006), and oxygen-limiting
conditions (Lungu et al. 2009; Roberts et al. 2020),
which are also encountered in human gastrointestinal
tract thus exacerbating L. monocytogenes infection in
humans (Horn and Bhunia 2018). Through the forma-
tion of biofilms (Lee et al. 2019; Bai et al. 2021) and
osmoadaptation, L. monocytogenes also resist desicca-
tion and sanitising agents (Pang et al. 2019). Thus, L.
monocytogenes can easily multiply on several surfaces,
increasing the chance of listeriosis outbreaks through
contaminated foods (Ferreira et al. 2014).
Traditional techniques for the listeriosis diagnostic
and L. monocytogenes detection in food include pre-
enrichment, selective enrichment, and isolation steps
that are time-consuming and labour-intensive (Le
Monnier et al. 2011; Bhunia 2014; Malica et al. 2015).
Most of the detection methods need pre-enrichment
cultivation to increase the pathogen concentration to
detectable levels of 10
4
to 10
5
CFU/mL (Gasanov et al.
2005;Wu2019). As alternatives to these problems,
many methods based on bacterial nucleic acid detec-
tion have been developed, like polymerase chain reac-
tion (PCR) and DNA hybridisation. Though these
molecular methods are highly sensitive, they are expen-
sive and need specialised technical knowledge and
equipment (Wang and Salazar 2016). PCR methods are
also highly susceptible to inhibitors and may not differ-
entiate viable from dead cells unless RNA has been
used as a target or cells are treated with propidium
monoazide before assay (Chen et al. 2017; Zheng et al.
2018). Therefore, immunological techniques (Figure 1),
exemplified mainly by enzyme-linked immunosorbent
assay (ELISA), immunosensors, and lateral flow immuno-
chromatographic assays (LFIA), are alternative
approaches to L. monocytogenes detection because of
their low cost, high sensitivity, and specificity, and sim-
plicity in data interpretation (Banada and Bhunia 2008;
Liu, Du, et al. 2017; Lv et al. 2019; Xu et al. 2021).
Immunological methods are developed using poly-
clonal (PAb) and monoclonal (MAb) antibodies pro-
duced against various target antigens of L.
monocytogenes (Bhunia 1997; Banada and Bhunia 2008;
Lathrop et al. 2014). However, some methods have
reported low specificity due to high genetic diversity
among L. monocytogenes strains (Dussurget et al. 2004).
Moreover, Listeria antigen expression may severely be
affected by the enrichment broths or growth conditions
used for culturing the bacteria (Nannapaneni et al.
1998; Lathrop et al. 2008) or acid, salt, and tempera-
ture-induced stressors before immunological detection
(Geng et al. 2003; Hahm and Bhunia 2006). Vital target
proteins should be strongly associated with the cell sur-
face and should be expressed by all L. monocytogenes
serotypes (Paoli et al. 2007). Furthermore, some avail-
able methods recognise antigenic targets shared by
other Listeria species and/or other non-Listeria bacteria
(Meyer et al. 2011). Thus, studies of surface biomarkers
of L. monocytogenes have been carried out in the past
several years, aiming to develop new immunological
approaches to pathogen control, prevention and to
help eliminate selective pre-enrichment steps (Boivin
et al. 2016; Mendonc¸a et al. 2016; Zhang et al. 2016;
Phraephaisarn et al. 2017).
The L. monocytogenes antigens for the generation
and selection of monoclonal and polyclonal antibodies
were reviewed previously (Bhunia 1997; Banada and
Bhunia 2008; Bhunia 2018) and are summarised in
Table 1. We will highlight the application of highly sen-
sitive and specific antibodies combined with relevant
and recent improvements of immunological
approaches to L. monocytogenes detection thus ena-
bling the improvement in the assay sensitivity.
Furthermore, we will provide a brief updated informa-
tion on L. monocytogenes pathogenesis and the viru-
lence factors (Table 1) that play central roles in
pathogenesis. Immunological assays based on virulence
determinants provide a solid foundation for the devel-
opment of a pathogen-specific diagnostic or detection
tool (Figure 1).
2. Virulence factors of Listeria monocytogenes
and their antigenic potentials
2.1. Listeria monocytogenes pathogenesis
After ingestion of contaminated food by the host, L.
monocytogenes overcome an acidic environment in the
2 L. LOPES-LUZ ET AL.
Table 1. Listeria monocytogenes virulence factors.
Virulence factors Size (kDa) Host cell receptor Function References
Protein regulatory
factor (PrfA)
27 Regulation of many virulence protein expression (Dussurget et al. 2002; Johansson and
Freitag 2019)
Sigma B (r
B
) 30.5 Regulation of stress and virulence genes (Liu et al. 2019)
Adhesion Proteins
Listeria adhesion
protein (LAP)
104 Hsp60
(chaperone protein)
Adhesion to intestinal epithelial cells; Disruption of
intestinal epithelial barrier
(Drolia et al. 2018; Drolia and Bhunia 2019)
Listeria adhesion
protein B (LapB)
184 Unknown Adhesion and invasion into host cells (Boivin et al. 2016)
Autolysin
amidase (Ami)
99 Unknown Adhesion to hepatocytes (Asano et al. 2011)
Fibronectin binding
protein (FbpA)
55.3, 48.6, 46.7, 42.4,
and 26.8
Fibronectin Adhesion to cells and also serves as a chaperone to
stabilise and secretion of LLO, InlB
(Osanai et al. 2013; Gelb
ı
cov
a et al. 2016)
Internalin J (InlJ) 92 Mucus Adhesion to epithelial cells and binds to human
intestinal mucin-2 (MUC2)
(Lind
en et al. 2008)
Internalin F (InlF) 90 Vimetin protein The crossing of blood-brain barrier (Dramsi et al., 1997; Ghosh et al. 2018)
Autolysin IspC 86 Unknown Adhesion to non-phagocytic cells (Ronholm et al. 2013)
Lmo1656 12 sorting nexin
6BAR complex
Transcytosis in goblet cells (David et al. 2018)
Invasion
Internalin (InlA) 88 E-cadherin (tight
junction protein)
Promotes bacterial internalisation into enterocytes and
bacterial transcytosis across the intestinal barrier
(Ribet and Cossart 2015; Phelps et al. 2018;
Drolia and Bhunia 2019)
Internalin B (InlB) 65 Met (tyrosine kinase
receptor c-Met),
gC1q-R/p32
Acts in the invasion of enterocytes and passage
through M-cells of Peyers patches
(Bierne and Cossart 2002; Lathrop et al. 2008;
Auriemma et al. 2010)
Virulence invasion
protein (Vip)
43 Gp96
(chaperone protein)
Invasion of epithelial cells (Cabanes et al. 2005)
LAP 104 Hsp60 Induces junctional protein dysregulation and increases
epithelial permeability (translocation)
(Drolia et al. 2018; Drolia and Bhunia 2019)
Lysis of Vacuole
Listeriolysin (LLO) hlyA 5860 Cholesterol A haemolysin that disrupts the vacuolar membrane and
releases the bacteria into the cytoplasm
(Phelps et al. 2018)
Phospholipase (plcA
PI-PLC; plcB
PC-PLC)
2933 Lyses of vacuole membrane (Poussin and Goldfine 2005; Suryawanshi
et al. 2017)
Cell-to-cell Spread
Actin polymerisation
protein (ActA)
8892 Initiates host cell actin polymerisation for bacterial
movement inside the cytoplasm
(Lathrop et al. 2008; Radoshevich and
Cossart 2018)
PC-PLC 2932 Cholesterol Lyses of vacuole membrane (Grundling et al. 2003)
Metalloprotease (Mpl) 29 Helps synthesis of PLC (Bhunia 1997; Yeung et al. 2005)
Miscellaneous
Activities
P60 (cell
wall hydrolase)
60 Invasion of non-professional phagocytic cells (Luo and Cai 2012; Lingala and Ghany 2015)
Bile salt
hydrolase (BSH)
36 Survival in gut (Dussurget et al. 2002)
Fructose-1,6-
bisphophate
aldolase (FAB)
30 Moonlighting protein: i) adhesion to the hosts cells and
ii) role in the pathogenesis
(Mendonc¸a et al. 2016)
Internalin C (InlC) 33 Human intestinal
mucin-2 (MUC2)
Perturbs apical cell junctions (Rajabian et al. 2009)
(continued)
CRITICAL REVIEWS IN MICROBIOLOGY 3
stomach by regulating bacterial pH homeostasis. The
glutamate decarboxylase system (Karatzas et al. 2012;
Paudyal et al. 2018), arginine deiminase pathway
(Cheng et al. 2017), and an agmatine deiminase system
(Ryan et al. 2009) are important mechanisms for bacter-
ial resistance to acids. Bile salt hydrolase (BSH) and bile
salt exclusion (BSE) proteins defend against the anti-
microbial activity of bile (Gahan and Hill 2014). Being a
facultative anaerobe, L. monocytogenes can also prolif-
erate under an oxygen-limiting environment in the
intestine (Burkholder et al. 2009; Lungu et al. 2009;
Roberts et al. 2020). Resident gut microbiota provides
little resistance against L. monocytogenes colonisation
and spread (Quereda et al. 2016; Becattini et al. 2017).
The crossing of the intestinal epithelial barrier is crit-
ical for the systemic spread of the infection which can
occur through both the small intestine and large intes-
tine (Drolia and Bhunia 2019; Bai et al. 2021). Listeria
adhesion protein (LAP), also known as acetaldehyde
alcohol dehydrogenase (AdhE) promotes bacterial pas-
sage through the epithelial barrier by opening the tight
junction barrier following interaction with the host cell
receptor, heat shock protein 60 (Hsp60), and conse-
quent activation of nuclear factor-jB (NF-jB) and
myosin light chain kinase (MLCK) early in the infection
process (Burkholder and Bhunia 2010; Drolia et al. 2018,
2020). While Internalin A (InlA) facilitates bacterial pas-
sage by interacting with the host cell E-cadherin via
intracellular route by transcytosis (Nikitas et al. 2011;
Bou Ghanem et al. 2012). F-actin-mediated bacterial
adhesion independent of interaction with E-cadherin
also promotes L. monocytogenes invasion into epithelial
cells (Ortega et al. 2017). M cells overlying Peyers
patches also help bacterial invasion, but in the absence
of M cells, L. monocytogenes still can reach lamina prop-
ria using the aforementioned virulence factors (Pron
et al. 1998). Besides, the Internalin B (InlB), a non-cova-
lently anchored surface protein interacts with the tyro-
sine kinase receptor c-Met/HFGR (Hepatocyte Growth
Factor Receptor) and also with glycosaminoglycans on
epithelial cells (Shen et al. 2000; Bierne and Cossart
2002; Quereda et al. 2019). InlB is also required for pas-
sage through M-cells of Peyers patches and also helps
bacterial internalisation into enterocytes (Cabanes et al.
2002; Sobyanin et al. 2017; Phelps et al. 2018). To cross
the blood-brain barrier, L. monocytogenes employs
Internalin (InlF) as the primary invasion protein (Ghosh
et al. 2018), while LLO and ActA are reported to play an
important role during olfactory epithelium invasion for
neurolisteriosis in neonates via the intranasal route of
infection (P
agelow et al., 2018).
Table 1. Continued.
Virulence factors Size (kDa) Host cell receptor Function References
Internalin H (InlH) 58 Contributes to systemic listeriosis (Markkula et al. 2011)
Autolysin
amidase (Ami)
99 Unknown Bacteriolysin: enhances the host immune response (Asano et al. 2011)
LAP 104 Hsp60 Upregulates TNF-a and IL-6 expression in intestinal Cells (Drolia and Bhunia 2019)
Listeriolysin (LLO) hlyA 5860 Cholesterol Induces lymphocyte apoptosis and suppresses
proinflammatory cytokines
(Pattabiraman et al. 2017)
Listeria nuclear-
targeted protein
A (IntA)
23 BAHD1 Decreases the hosts immune response (Lebreton et al. 2011)
Listeriolysin S <5Hemolytic and cytotoxic; bacteriocin (bactericidal) (Cotter et al. 2008; Quereda et al. 2016)
4 L. LOPES-LUZ ET AL.
Internalised L. monocytogenes are trapped in the
vacuole or phagosome which is disrupted by listerioly-
sin O (LLO), phosphatidylcholine-specific phospholipase
(PC-PLC), and phosphatidylinositol-specific phospholip-
ase C (PI-PLC), releasing the bacteria in the cytoplasm
(Phelps et al. 2018). In some cases, bacteria can persist
in the vacuole thus prolonging the incubation period
(Kortebi et al. 2017; Bierne et al. 2018). In the absence
of ActA, vacuolar bacteria may maintain a viable but
nonculturable state (Bierne et al. 2018). Once in the
cytoplasm, the bacterium replicates rapidly and spreads
by inducing actin nucleation and polymerisation by
inducing ActA expression (Theriot et al. 1992;
Radoshevich and Cossart 2018; Johansson and Freitag
2019). At this time of infection, L. monocytogenes may
produce extracellular vesicles that will release virulence
factors and toxins to overcome the autophagy system
of the host cells (Cheng et al. 2018; Coelho et al. 2019).
Besides, it has been shown that LLO can induce
lymphocyte apoptosis, increasing pathogenesis by sup-
pressing proinflammatory cytokines (Interleukin 6 and
TNFa) and increasing IL-10 expression induced by
innate apoptotic immunity (Pattabiraman et al. 2017).
Target organ infection may depend on the spread
dynamics of the L. monocytogenes population. Within a
population, there are bacteria called rare pioneerthat
determine the boundaries of infection foci and may
promote the persistence of bacterial infection. The cell-
to-cell spread is not homogeneous and the rare pio-
neerbacteria can infect non-adjacent cells (Ortega
et al. 2019). Although gut microbiota act by preventing
the spread of the pathogen in the liver and spleen
(Becattini et al. 2017; Pickard et al. 2017), both organs
tend to have the same bacterial transport capacity
found in the small intestine and colon. The gallbladder,
however, is the organ that has shown higher bacterial
density (Zhang, Abel, et al. 2017) probably due to the
poor ability to recruit immune effector cells (Gonzalez-
Escobedo and Gunn 2013) and bacterial resistance to
biles (Gahan and Hill 2014).
2.2. Innate immune response
The innate immune system is the first line of defense
orchestrated by resident microbiota, mucus, epithelial
cells, and phagocytes which produce inflammatory
cytokines for recruitment and activation of immune
cells for controlling L. monocytogenes. Recent studies
have shown that IL-6 receptor depletion in inflamma-
tory monocytes decreased the expression of CD38
ectoenzyme that is important in phagocytosis and early
control of L. monocytogenes infection (Hoge et al. 2013).
Among phagocytes, neutrophils play an essential role
in the initial clearance of L. monocytogenes in the liver
(Shi et al. 2011). In addition, a recent study showed that
L. monocytogenes infected hepatocytes act on macro-
phage migration to the liver through TLR-2 signalling,
promoting bacterial clearance and resistance to infec-
tion (Wang et al. 2019). However, the bacterial clear-
ance by macrophages can be affected by the induction
Figure 1. Immunoassay formats include ELISA (Enzyme- Linked Immunosorbent Assay), Immunofluorescence Assay, and Lateral
Flow Immunochromatographic Assay.
CRITICAL REVIEWS IN MICROBIOLOGY 5
of actin polymerisation that protects the bacteria from
autophagic recognition (Cheng et al. 2018).
2.3. Virulence factors and antigenic potential
L. monocytogenes express an array of virulence factors
(Table 1) that are essential for its survival and persist-
ence in the gastrointestinal tract and crossing of the
intestinal, blood-brain, and placental barriers (Bhunia
2018). The specific steps in the infection process involve
bacterial adhesion and intestinal barrier crossing (Drolia
and Bhunia 2019), epithelial cell invasion (Camejo et al.
2011), cell-to-cell spread (Travier et al. 2013; Travier and
Lecuit 2014; Bhunia 2018) and systemic dissemination
(Coelho et al. 2019).
2.3.1. Listeria adhesion protein
Among the virulence factors, L. monocytogenes surface
protein LAP interacts with the host cell receptor, Hsp60
for adhesion and translocation across the intestinal epi-
thelial barrier (Jagadeesan et al. 2011; Drolia et al. 2018;
Bai et al. 2021). LAP is a 104 kDa bifunctional house-
keeping enzyme (acetaldehyde alcohol dehydrogenase,
AdhE) and is present in all Listeria species of sensu
stricto group (Jagadeesan et al. 2010; Kim et al. 2015).
LAP is a highly immunogenic protein and a monoclonal
antibody raised against it has been used to study the
expression of this protein in Listeria under various
growth conditions (Pandiripally et al. 1999; Santiago
et al. 1999; Jaradat and Bhunia 2002).
Alternative to antibodies, Hsp60 is useful as a cap-
ture molecule that was coupled to paramagnetic beads
(Koo et al. 2011) or on a microfluidic device (Koo et al.
2009) for L. monocytogenes detection. Hsp60 is a human
cell receptor that interacts with LAP and the affinity
constant for Hsp60-LAP is 1.68 10
8
M, equivalent to
antigen-antibody interaction (Kim et al. 2006;
Jagadeesan et al. 2011). Thus, LAP-Hsp60 are useful sur-
face proteins for the development of immunoassays for
the detection of L. monocytogenes.
2.3.2. Internalin proteins
Members of the internalin multigene family cover sur-
face proteins important in bacterial adhesion and inva-
sion (Cossart and Helenius 2014; Quereda et al. 2019),
being highly immunogenic molecules that generate
antibodies useful for applications in the diagnosis and
detection of L. monocytogenes in foods (Heo et al. 2007;
Paoli and Brewster 2007). The two most studied inter-
nalins, InlA, and InlB contain a leucine-rich repeat (LRR)
in the amino-terminal region followed by a conserved
inter-repeated (IR) region (Cabanes et al. 2002).
InlA is an 88 kDa surface protein covalently linked to
the cell wall and binds to the eukaryotic receptor, E-
cadherin, promoting bacterial internalisation into enter-
ocytes (Radoshevich and Cossart 2018) and bacterial
transcytosis across the intestinal barrier (Nikitas et al.
2011). InlA is abundantly expressed and uniformly dis-
tributed on the surface of L. monocytogenes cells and is
anchored covalently to the cell wall, while InlB has short
chains with a helical region and a loop that allows spe-
cific recognition of molecules within the host cell
(Gessain et al. 2015).
InlA protein is highly antigenic and antibodies
(Mendonca et al. 2012) or aptamers (Ohk et al. 2010)
were developed for sensitive detection using immuno-
sensors. Antibodies were also developed against InlB
protein (65 kDa) and used in surface plasmon resonance
sensors (Leonard et al. 2005; Hearty et al. 2006). A com-
parative genomics approach was also used to generate
a PAb against InlB (Lathrop et al. 2008,2014).
Two studies have focussed on the use of antibodies
against variable and immunodominant regions of inter-
nalins. Hearty et al (Hearty et al. 2006) produced anti-
InlA antibody (MAb-2B3) that reacted strongly with
strains belonging to serotype 1/2a, but weakly with 1/
2c, 4a, and 4b. Since most of the listeriosis outbreaks
involve serotype 4b, the MAb-2B3 is not a desirable
candidate for use in food screening. Mendonca et al
(Mendonca et al. 2012) also produced MAb-2D12
against InlA and its characterisation showed high speci-
ficity for both L. monocytogenes and L. ivanovii.
Antibodies, bi-functional antibodies, and antibody frag-
ments against InlB have been used in the development
of immunology-based techniques useful in the detec-
tion of L. monocytogenes in human, environmental, and
food samples with suspected contamination (Hearty
et al. 2006; Tully et al. 2008; Lathrop et al. 2014; Owais
et al. 2014; Gene et al. 2015). Despite the similar protein
structures with InlA, InlB protein can elicit more specific
antibodies against the L. monocytogenes since it is
unique to its genome (Glaser et al. 2001; Hurley
et al. 2019).
2.3.3. Haemolysin, phospholipases and ActA
As mentioned above, internalised L. monocytogenes
secrete LLO, PC-PLC, and PI-PLC to disrupt the vacuolar
membrane and are released into the cytoplasm. L.
monocytogenes also express ActA that initiates host cell
actin polymerisation facilitating bacterial movement
from cell-to-cell (Tilney and Portnoy 1989;K
uhn and
Enninga 2020). These virulence proteins have been
expressed in vitro as recombinant antigens because
they have immunogenic sequences that are useful in
6 L. LOPES-LUZ ET AL.
the generation of antibodies. Antibodies generated
against a sequence of 17 amino acids from the N-ter-
minal region of LLO (58 kDa) have shown high specifi-
city for L. monocytogenes in the immunodetection of
the bacterium (Day and Hammack 2019). Due to its
high immunogenicity, LLO has also been used as an
antigen in the development of immunological
approaches for the diagnosis of listeriosis in both ani-
mals and humans (Rekha et al. 2006; Shoukat et al.
2013; Suryawanshi et al. 2017).
Antibodies generated against PI-PLC (2932 kDa)
(Poussin and Goldfine 2005) and ActA (8892 kDa)
(Paoli and Brewster 2007; Lathrop et al. 2008,2014;
Travier and Lecuit 2014) found application in L. monocy-
togenes detection. PI-PLC is highly conserved among
the Listeria genus and other non-Listeria species, such
as Bacillus cereus, B. anthracis,Staphylococcus aureus,
and Clostridium spp., presenting 5156% identity
among these species (Daugherty and Low 1993) and
inducing high cross-reactivity in diagnostic tests
(Suryawanshi et al. 2017). These anti-PI-PLC antibodies
possibly will have cross-reactivity in assays for the
detection of L. monocytogenes in food. Highly specific
PAb against ActA, on the other hand, have the potential
to be used in immunological assays for specific detec-
tion of L. monocytogenes (Lathrop et al. 2014). Coupling
anti-ActA single-chain antibody with magnetic beads
was also effective for the separation of L. monocyto-
genes from food matrices (Paoli et al. 2007).
2.3.4. Autolysins
Among the autolysins, a protein of 60 kDa (P60, also
known as cell wall hydrolase - CWH) is highly conserved
among the Listeria genus (80%90%) (Bubert et al.
1992; Wieckowska-Szakiel et al. 2002; Yu et al. 2004)
and the generated antibodies are highly reactive to
each member. Thus, the L. monocytogenes detection
using MAb against the P60 protein (CWH) structure
showed high cross-reactivity (Wang et al. 2017).
However, the expression of a peptide composed of 11
amino acids unique to the P60 protein has allowed the
generation of highly specific antibodies that are suit-
able for the detection of L. monocytogenes (Etty et al.
2019,2020).
Antibodies were also generated against another
autolysin amidase, called peptidoglycan muramidase or
N-acetylmuramidase (encoded by murA)inL. monocyto-
genes (Bhunia et al. 1991; Bhunia and Johnson 1992;
Carroll et al. 2003). These MAbs were reactive against
antigenic epitopes that are highly selective for L. mono-
cytogenes with very little or no reaction with other
Listeria spp. (Bhunia et al. 1991; Bhunia and Johnson
1992). These antibodies have been used on various
immunosensor platforms (Table 2) for the detection of
this pathogen from food (Geng et al. 2004; Koo et al.
2009; Ohk et al. 2010; Ohk and Bhunia 2013; Rodr
ıguez-
Lorenzo et al. 2019). Besides, the MAb-C11E9 was also
used for the delivery of nanocargo loaded with the anti-
cancer drug on the surface of L. monocytogenes target-
ing solid tumours (Akin et al. 2007).
2.3.5. Miscellaneous proteins
Cell surface proteins also play an essential role as spe-
cific antibody-inducing molecules (Zhang et al. 2016).
Surface autolysin (IspC; 86 kDa) has immunodominant
epitopes for L. monocytogenes serotype 4b (Ronholm
et al. 2013) and Fructose 1,6-Bisphosphate Aldolase
(FAB; 30 kDa) has strong immunogenicity (Mendonc¸a
et al. 2016). Listeria adhesion protein B (LapB; 184 kDa)
that is absent in non-pathogenic species, has the
potential to generate species-specific antibodies (Boivin
et al. 2016). Thus, IspC, FAB, LapB, and LAP are useful
surface proteins for the development of immunoassays
for the detection of L. monocytogenes.
3. Listeria monocytogenes detection:
immunological approaches
Antibodies generated against the key antigens of L.
monocytogenes have been useful in the development of
new immunological techniques for listeriosis control
and prevention (Yu et al. 2004; Hearty et al. 2006; Etty
et al. 2020). Techniques based on applied immunology
and biotechnology have shown quick, simple, and inex-
pensive applications (Li, Jing, et al. 2018; Lv et al. 2019).
MAbs and PAbs are the essential key elements for
immunological assay development such as ELISA,
immunofluorescence, lateral flow immunoassay, and
immunosensors (Banada and Bhunia 2008;
Jeyaletchumi et al. 2010; Law et al. 2014;V
alimaa et al.
2015; Xu et al. 2021). Important improvements based
on biotechnology and antibody engineering have been
performed, such as techniques based on single-chain
antibody (scFv) fragments and heavy-chain antibodies
(V
H
H fragments). scFv and V
H
H antibodies have similar
antigen-binding affinities as whole antibodies, however,
those can be produced in large quantities using micro-
bial expression systems and have easy handling and
extensive application range (Paoli et al. 2004; Liu,
Xiong, et al. 2017; King et al. 2018).
CRITICAL REVIEWS IN MICROBIOLOGY 7
3.1. Elisa and immunofluorescence-based methods
to detect Listeria
ELISA and immunofluorescence assays have been
widely used in the diagnosis of many diseases by
detecting specific antigens or antibodies. The results
are obtained through the enzymatic reaction of chrom-
ogens and fluorescent chemical compounds in ELISA
and immunofluorescence, respectively (Boonham et al.
2014; Diercks et al. 2017). The application of these
immunological assays has expanded to detect harmful
substances found in foods such as mycotoxins
(Leszczy
nska et al. 2013), allergens (Ito et al. 2016), bac-
terial toxins or antigens, and whole cells of gastro-
enteric bacteria (Zhao et al. 2017). Thus, these assays
have become an important tool in the prophylaxis of
several diseases, including listeriosis. Rapid detection
and identification of L. monocytogenes antigens in food
samples or food preparation/production premises can
help introduce proper Listeria control strategy and
potentially reduce listeriosis outbreaks (Cho et al. 2015;
Wang et al. 2015).
In the last decades, several methodologies based on
ELISA or immunofluorescence assays have been devel-
oped (Table 2). Notably, these immunoassays have
reduced the time for pathogen detection procedures
from 4 to 5 days to just 1 day compared to conventional
methodologies (Cavaiuolo et al. 2013; Bhunia 2014).
Regarding specificity, enzymatic and fluorescent-based
immunoassays have reached important improvements
for the accurate detection of L. monocytogenes in sam-
ples with suspected contamination. The production of
highly specific MAbs against L. monocytogenes antigens
ensures the specificity, especially the antibodies pro-
duced against epitopes of InlA (Mendonca et al. 2012),
InlB (Karamonov
a et al. 2003) or LLO (Prommajan et al.
2016) and recombinant p60 protein, that excluded con-
served regions between the genus Listeria (Coutu et al.
2014). PAbs can be used as efficient capture molecules
due to the heterogeneity of recognition while MAbs
assure the specificity of the method (Zhang et al. 2016).
In ELISA-based methods, the assay format can also
influence the specificity. Karamonov
a et al.
(Karamonov
a et al. 2003) demonstrated that the sand-
wich-type format, in which bacterial cells are free, was
more specific than the competitive format that keeps
bacteria immobilised. Also, both formats have distinct
antibody binding kinetics that can influence specificity.
In the sandwich-type format, it was possible to
decrease the limit of detection (LOD) through the use
of antibodies such as an anti-InlB antibody
(Karamonov
a et al. 2003), adaptations with piezoelectric
cantilever sensor (Sharma and Mutharasan 2013), use of
p60 recombinant antigens (Coutu et al. 2014) or use of
silica nanoparticles carrying polyacrylic acid (Chen
et al. 2015).
Tu et al. (Tu et al. 2016) showed that the use of V
H
H
fragments in indirect ELISA could decrease costs and
time for the detection of L. monocytogenes in food com-
pared to the use of conventional MAb or PAb. Liu,
Xiong, et al. (2017) showed that the application of scFv
fragments as a capture agent was sufficient to increase
the sensitivity of the assays, in addition to being a key
reagent of easy manipulation and with wide
application.
Although ELISA and immunofluorescence-based
methods are simpler than molecular techniques, they
are not considered rapid (Hearty et al. 2006; Jiang et al.
2019) because of the various washing and incubation
steps, in addition to the need for equipment to read or
visualise the reaction. These steps make the technique
more expensive, complex, and time-consuming (Bruno
et al. 2015; Chen et al. 2015). Thus, ELISA and immuno-
fluorescence assays may have limitations in the emer-
gency screening of food or production facilities with
suspected contamination by L. monocytogenes.
3.2. Lateral flow immunochromatographic-based
methods: rapid tests
Rapid tests are widely used in diagnostic medicine to
detect specific antibodies or pathogens and are charac-
terised by the availability of results in a few minutes (5
to 30 min) (Koczula and Gallotta 2016). In addition to
speed, rapid tests are inexpensive, simple, and easy to
perform. Thus, rapid tests have been applied in several
other areas such as environmental (Singh et al. 2015),
veterinary medicine (Jain et al. 2018), quality control in
industries, and food safety through the detection of
toxins, allergens (Ito et al. 2016) or even pathogens
such as L. monocytogenes (Cho and Irudayaraj 2013; Shi
et al. 2015; Wang et al. 2017). A rapid test that can diag-
nose a disease or detect pathogens in places with lim-
ited resources or when no pre-treatment or complex
processing of suspect samples is required, is classified
as a point-of-care test (POCT) (Koczula and
Gallotta 2016).
The lateral flow assay, also known as lateral flow
immunochromatographic assay (LFIA), is the immuno-
logical technique most used as a rapid test or POCT.
The assay principle is based on the sample containing
the analyte of interest introduction in the sample pad.
The sample moves laterally by the action of capillarity
through different zones of a strip. The first zone is the
conjugated pad which contains molecules conjugated
8 L. LOPES-LUZ ET AL.
Table 2. Immune approaches based on ELISA, immunofluorescence and immunosensors developed to detect L. monocytogenes
and other foodborne pathogens.
Immunoassay platforms Detected species
LOD (CFU/mL or CFU/
g)
(only L. monocytogenes) Food matrix
Improvement,
combination or antibody/
label used References
Multicolour and
ultrasensitive ELISA
E. coli O157:H7
Salmonella serotype
Choleraesuis
L. monocytogenes
7.0 10
1
in food
sample
(no enrichment)
milk Multicolour platform /
multicolour
fluorescence
hybridisation chain
reaction concatemers
(Lv et al. 2019)
Sandwich assay L. monocytogenes 6.5 in food sample
(48 h enrichment)
spiked pork and
pasteurised milk
Well-oriented single-chain
Fv (scFv)
antibody fragment
(Liu, Xiong, et al. 2017)
Dot-ELISA L. monocytogenes 1.0 10
6
in food
sample
(5 h enrichment)
pork meat and
fresh vegetables
MAb LMF3-238 (Prommajan
et al. 2016)
Sandwich assay L. monocytogenes 1.0 10
4
in food
sample
(no enrichment)
pasteurised milk Heavy-chain antibodies
(VHHs) (species-
specific nanobodies)
(Tu et al. 2016)
Plasmonic ELISA L. monocytogenes 8.0 10
1
in food
sample
(no enrichment)
spiked Lettuce Silica nanoparticles
carrying
polyacrylic acid
(Chen et al. 2015)
Sandwich assay Cross-reactivity
between:
L. monocytogenes
L. innocua
E. coli
S. Typhimurium
S. aureus
1.0 10
3
in pure
culture
(no enrichment)
Antibodies against
recombinant
p60 protein
(Coutu et al. 2014)
Indirect assay L. monocytogenes
E. coli O157
10
3
in food sample
(no enrichment)
cucumber Commercial antibodies
from Abcam: anti-E.
coli O157 (ab20976)
and anti- L.
monocytogenes
(LZA2) (ab11439)
(Cavaiuolo et al. 2013)
Sandwich assay L. monocytogenes 10
3
in food sample
(no enrichment)
milk A novel piezoelectric
cantilever sensor
(Sharma and
Mutharasan 2013)
Competitive fluorescence
immunoassay
L. monocytogenes 1.0 in pure culture
(18 h enrichment)
MAbs against
recombinant
p60 protein
(Beauchamp
et al. 2012)
Sandwich assay L. monocytogenes 5.0 in food sample
(24 h enrichment)
milk Anti-internalin B antibody (Karamonov
a
et al. 2003)
SERS spectroscopy
combined with
microfluidic
L. monocytogenes 110
5
in pure culture
(no enrichment)
MAb-C11E9/
aminopeptidase
(66 kDa)
(Rodr
ıguez-Lorenzo
et al. 2019)
Custom-built multichannel
surface plasmon
resonance
(SPR) biosensor
L. monocytogenes
E. coli O157H:7
S. Enteritidis
28 CFU/25 g in food
sample
(24 h enrichment)
chicken PAb anti-Listeria (Zhang, Tsuji,
et al. 2017)
Multiplex fibre
optic biosensor
L. monocytogenes,
E. coli O157:H7
S. enterica
10
3
in food sample
(24 h enrichment)
ready-to-eat
meat samples
PAb-66 anti-Listeria
MAb-C11E9 anti-
aminopeptidase
(66 kDa)
(Ohk and Bhunia 2013)
Fibre optic immunosensor
coupled with
immunomagnetic
separation
L. monocytogenes
L. ivanovii
310
2
in food sample
(1618 h
of enrichment)
hotdog and
soft cheese
MAb-2D12 anti-InlA (Mendonca et al. 2012)
Antibodyaptamer
functionalised fibre-
optic biosensor
L. monocytogenes 10
3
in pure culture
and 10
2
CFU/25 g in
food samples
(18 h enrichment)
ready-to-eat
meat products
Aptamer-A8 anti-InlA (Ohk et al. 2010)
Surface plasmon
resonance
(SPR) biosensor
L. monocytogenes 210
6
in pure culture
(no enrichment)
scFv antibody/ActA (Nanduri et al. 2007)
Automated fibre
optic biosensor
L. monocytogenes 210
3
in PBS and
510
5
in food sample
(20 h enrichment)
Frankfurter sample PAb anti-Listeria
PAb-66 anti-Listeria
(Nanduri et al. 2006)
Fibre-optic immunosensor L. monocytogenes 4.310
3
in pure
culture
(20 h enrichment)
Hot dog and Bologna MAb-C11E9 anti-
aminopeptidase
(66 kDa)
(Geng et al. 2004)
Resonant mirror biosensor L. monocytogenes 110
6
in pure culture
(no enrichment)
MAb-C11E9 anti-
aminopeptidase
(66 kDa)
(Lathrop et al. 2003)
CRITICAL REVIEWS IN MICROBIOLOGY 9
to coloured or fluorescent labels (Hsieh et al. 2017). In
conventional methods, colloidal gold nanoparticles
conjugated to specific antigens or antibodies that act
as detection agents are commonly used (Mao et al.
2009). Then, there is the detection zone formed by por-
ous membranes, such as the nitrocellulose (NC) mem-
brane, containing specific biological reagents
immobilised in test and control lines (Koczula and
Gallotta 2016). Polyclonal antibodies are widely used in
the test lines as a capture agent for the conjugate ana-
lyte (Banerjee and Jaiswal 2018; Jacinto et al. 2018). The
recognition of the analyte in the sample promotes
staining in the test line and the result can be read with
the naked eyes. The test operation is certified by
observing colour in the control line (Sajid et al. 2015;
Hsieh et al. 2017).
To ensure food safety and for foodborne pathogen
and toxin detection, the development of LFIA as a rapid
and inexpensive test has exploded in recent decades
(Table 3). LFIA is an essential detection tool in the food
industry and has aided the industry to eliminate and
prevent contamination by L. monocytogenes in process-
ing environments and in food products (Phraephaisarn
et al. 2017). Specifically, the greatest advantage of LFIA
compared to ELISA and immunofluorescence-based
methods are that a person with minimal technical
knowledge can perform the test without the need for a
laboratory or specialised equipment (Wang et al. 2015).
To be used as a tool to prevent listeriosis outbreaks,
an LFIA should be able to detect the lowest possible
number of bacterial cells per gram or millilitre. The min-
imum infection dose varies widely (<100 cells to 10
11
cells) among neonates, elderly, immunocompromised,
and healthy individuals (Sim et al. 2002; Ooi and Lorber
2005; Pouillot et al. 2016). The LFIA available for the
detection in pure culture or food samples have low
accuracy and the LOD of commercial tests ranges from
110
5
to 1 10
6
CFU/mL (Kitao et al. 2010; Cho and
Irudayaraj 2013). Combining LFIA with molecular tech-
niques, such as recombinase polymerase amplification
lateral flow (RPA-LF), Du et al. (Du et al. 2018) reported
LOD to be 1.5 10
1
CFU/mL in pure culture within
15 min: 10 min for DNA amplification and 5 min for visu-
alisation of the amplicons. In different food samples,
after 6 h of enrichment, the researchers reported the
LOD to be 1236 CFU/mL while Liu, Du, et al. (2017)
reported a LOD of 22 to 36 CFU/mL in food sample
within 30 min.
Application of LFIA in food samples can be limited
by the types of food matrix used, which often affects
the biological activity of reagents and bacterial anti-
gens, decreasing the sensitivity of conventional tests (Li
et al. 2017). Bacterial separation and concentration
methodologies have been used as sample pre-treat-
ment to reduce the food matrix interference. Li et al.
(Li, Li, et al. 2018) developed a nucleic acid lateral flow
biosensor with an immunomagnetic separation plat-
form based on a biotin-streptavidin system that
resulted in a LOD of 3.5 10
3
in pure culture and LOD
of 3.5 10
4
CFU/g in lettuce, both after 6 h of enrich-
ment. Biotinylated antibodies show better capture cap-
acity in an immunoseparation system due to the high
affinity of streptavidin for biotin. In this system, the
binding sites of antibodies coupled to magnetic nano-
particles are targeted. Combining lateral-flow chroma-
tography enzyme immunoassay platform with
immunomagnetic separation and concentration, Cho
and Irudayaraj (Cho and Irudayaraj 2013) found a LOD
of 10
2
CFU/mL in pure culture and artificially contami-
nated milk samples in 2 h. In addition to reducing the
interference of the food matrix, immunomagnetic sep-
aration or concentration decreases other steps of the
pre-treatment process, maintains the biological activity
of the samples, and can also be used to emit signals for
quantitative analysis when the magnetic nanoparticles
are conjugated to monoclonal antibodies (Mendonca
et al. 2012; Cho and Irudayaraj 2013; Shi et al. 2015).
Employing a pathogen enrichment device (PED)
coupled with lateral-flow immunoassay, Hahm et al.
(Hahm et al. 2015) reported LOD of 3.2 10
2
to
5.6 10
4
for L. monocytogenes from food samples. The
LOD values varied depending on the initial inoculation
and the enrichment time (824 h), and the LOD was cal-
culated by counting the bacteria on MOX agar plates.
Combining surface-enhanced Raman scattering
(SERS) with bacteriophage amplification, Stambach
et al. (Stambach et al. 2015) demonstrated a LOD of
510
4
CFU/mL to 1 10
7
CFU/mL, after enrichment of
28 h. Using a similar approach, Wu (Wu 2019)
reported LOD of 75 CFU/mL in pure culture and about
200 CFU/mL in artificially contaminated milk spiked
with different concentrations of L. monocytogenes. They
used SERS-based LFIA for the simultaneous detection of
L. monocytogenes and Salmonella enterica serovar
Typhimurium.
3.3. Immunosensors
Immunological approaches based on biosensors also
use antibodies as key molecules for the recognition of
analytes in different types of samples. Thus, biosensors
that use antibodies are called immunosensors. In this
technique, the antigen-antibody bond is converted into
a measurable physical signal through a signal
10 L. LOPES-LUZ ET AL.
Table 3. Lateral flow-based methods for L. monocytogenes detection and their technical features.
Lateral flow-based methods/labelling Target species
LOD (CFU/mL or CFU/
g) in pure culture j
runtime (only L.
monocytogenes)
LOD (CFU/mL or CFU/g) in food matrix
jruntime (only L. monocytogenes) Improvement or labelling or antibody References
SERS-based lateral flow
immunochromatographic assay
L. monocytogenes
Salmonella
Typhimurium
75 j229 CFU in milk jA novel SERS using two different Raman
reporters:
anti-L. monocytogenes antibody and
DTNB-modified AuNPs
anti-S. Typhimurium antibody and 4-
MBA-modified AuNPs
(Wu 2019)
Nucleic acid lateral flow biosensor L. monocytogenes 3.5 10
3
j6 h 3.5 10
4
in lettuce j6 h Combination: PCR-based single specific
nucleic acid sequence amplification
with nucleic acid lateral flow
biosensor; pre-treatment platform:
immunomagnetic separation based on
the streptavidin-biotin system
(Li, Li, et al. 2018)
Lateral flow strip L. monocytogenes 1.5 10
1
j15 min 1236 in milk, chicken breast, bread,
cereal milk, milk tea, and latte
coffee j15 min after 6 h
of enrichment
Recombinase polymerase amplification (Du et al. 2018)
SERS-based lateral flow strip biosensor L. monocytogenes
Salmonella serotype
Enteritidis
19 j5 min 22 to 36 in milk, chicken breast and
beef j530 min
Recombinase polymerase amplification (Liu, Du, et al. 2017)
Fluorescence
immunochromatographic assay
L. monocytogenes 4.0 10
4
to 4.0 10
5
j110
4
in sausage, pork and milk 100
times enriched j3h
Immunomagnetic separation and
polyclonal antibodies generated from
whole killed L. monocytogenes cells
(Li, Jing, et al. 2018)
Gold nanoparticle-based paper sensor Listeria sp. 1 10
3
to 1 10
4
j1 to 9 in milk j10 min after
8 h enrichment
Monoclonal antibody against P60 protein (Wang et al. 2017)
Lateral flow immunochromatography
assay coupled to pathogen enrichment
device (PED)
Escherichia coli
O157:H7, Salmonella
enterica,L
monocytogenes
3.2 10
2
to 5.6 10
4
in spinach,
ground beef, hotdogs, and e.g.gs j
824 h of enrichment
Pathogen enrichment device (PED)
containing a growth chamber, filters,
and an ion-exchange cartridge to
deliver bacteria directly onto the
detection platforms
(Hahm et al. 2015)
SERS-lateral flow immunochromatography L. monocytogenes 510
4
to 1 10
7
j28h
Bacteriophage amplification (phage A511) (Stambach et al. 2015)
Superparamagnetic lateral flow
immunoassay
L. monocytogenes 110
4
j20 min Various food sources, including
chicken, fish, flammulina
velutipes, etc.
Superparamagnetic particles and a pair of
monoclonal antibodies
(Shi et al. 2015)
Lateral-flow enzyme
immunoconcentration
L. monocytogenes 95 j2 h 97 in reduced-fat milk j2 h Immunomagnetic separation and
concentration step
(Cho and
Irudayaraj 2013)
Nucleic acid lateral flow immunoassay L. monocytogenes
Listeria spp.
110
5
in dairy products, ready-to-eat
salad, sprouted seeds, nutrition for
children, etc j15 min after 24 h
(enrichment) and 3.5 h (isolation of
template DNA and a PCR
amplification step)
A duplex PCR (two labelled primer sets) (Bla
zkov
a et al. 2009)
Runtime not mentioned.
Not mentioned.
CRITICAL REVIEWS IN MICROBIOLOGY 11
transducer coupled to the biosensor (Banada and
Bhunia 2008; Bhunia 2008; Singh and Bhunia 2018;Xu
et al. 2021). Compared to lateral flow-based methods,
immunosensors also have flexibility in use, many are
portable and allow the simultaneous detection of sev-
eral biomolecules. The effectiveness of an immunosen-
sor depends on the high specificity and sensitivity that
are directly related to the type, quality, and reactivity of
the antibodies used and other combined techniques
(Ohk et al. 2010; Smolsky et al. 2017). Various immuno-
sensor platforms are developed for specific detection of
L. monocytogenes which are summarised in Table 2.
Fibre optic biosensor utilises sandwich immunoassay
configuration and the specificity of the assay depends
on the antibody pair used and the assay sensitivity on
the generation of the evanescent wave signal. Fibre
optic sensor has been developed using antibodies to N-
acetyl muramidase (MAb-C11E9) (Geng et al. 2004) and
InlA (Mendonca et al. 2012) and the assay was highly
specific and the LOD varied from 10
2
10
4
CFU/mL.
Polyclonal anti-Listeria antibodies have heterogeneous
recognition of pathogen antigens and are useful as cap-
ture agents when immobilised on the fibre optic bio-
sensor platform (Nanduri et al. 2006,2007; Zhang, Tsuji,
et al. 2017). MAb-C11E9 application has improved the
specificity of immunosensors based on SERS
(Rodr
ıguez-Lorenzo et al. 2019), multiplex fibre optic
biosensor (Ohk and Bhunia 2013), and resonant mirror
biosensor (Lathrop et al. 2003). Although MAb-C11E9
reacts weakly with L. innocua antigens under certain
conditions of cultivation and bacterial lysis (Lathrop
et al. 2003), it was demonstrated that the application of
the MAb-C11E9 in the SERS-based method can distin-
guish L. monocytogenes from L. innocua (Rodr
ıguez-
Lorenzo et al. 2019).
The SERS-based method is a quantitative, stable, sim-
ple, and specific technique that allows improving L.
monocytogenes detection and has been extensively
studied. This technique is used to detect isolated mole-
cules on surfaces with high sensitivity through Raman
dispersion (Smolsky et al. 2017; Liu, Du, et al. 2017;
Oravec et al. 2018;). Recently, gold nanostars conju-
gated to MAb-C11E9 were used for the detection of L.
monocytogenes using SERS (Rodr
ıguez-Lorenzo et al.
2019). A surface plasmon resonance sensor was also
developed for the detection of L. monocytogenes using
MAb (Lathrop et al. 2003; Nanduri et al. 2006) or phage-
displayed single chain antibodies (Nanduri et al. 2007).
Zhang et al (Zhang, Tsuji, et al. 2017) developed a
multichannel SPR using PAb for the detection of L.
monocytogenes,S. Enteritidis, and E. coli O157:H7. After
24 h of enrichment, the three pathogens were detected
at a concentration of >10
5
CFU/mL in pure culture, and
LOD was estimated to be 2.3 10
5
CFU/mL for L. mono-
cytogenes in a chicken sample.
The use of highly specific MAbs, PAbs, or other
reagents based on antibody engineering is essential for
developing highly specific immunosensors. An aptamer
specific for internalin A was used in combination with
polyclonal anti-Listeria P66 in a sandwich format for
detection of L. monocytogenes in a fibre-optic biosensor
(Ohk et al. 2010). It is important to note that immuno-
logical approaches that use some type of pre-treatment
or separation process with magnetic nanoparticles
improve the sensitivity of the assays. However, these
combined approaches require additional more complex
analyses when compared to results interpreted with the
naked eye (Fisher et al. 2009; Uusitalo et al. 2016).
4. Conclusions and future perspectives
Comprehensive knowledge of the pathogenesis and
the role of each virulence factor in pathogenicity pro-
vides a solid foundation for developing a robust diag-
nostic or detection tool for a pathogen. Targeting a
virulence factor as an antigen also helps to develop an
assay that can detect only the pathogenic species (ex.
L. monocytogenes) within a genus that may have non-
pathogenic members (ex. L. innocua,L. welshimeri,L.
seeligeri.L. marthii, and others) that share high genetic
sequence similarities. In this review we provide updated
information on the pathogenesis of L. monocytogenes,
the immunogenicity of virulence determinants and
their expression in food or food testing reagents (such
as in enrichment broths) that are key to the generation
and selection of highly specific monoclonal and poly-
clonal antibodies for immunological assay
development.
The immunological assays include ELISA, immuno-
fluorescence, lateral flow immunochromatography, and
immunosensors. Though immunological methods are
perceived as less sensitive than molecular-based meth-
ods (such as PCR), their affordability, ease of use, rapid
results, and portability (point-of-care use) make them
highly attractive. Furthermore, the limit of detection
varies among the immunoassay platforms: a traditional
ELISA requires a minimum of 1 10
6
CFU, and commer-
cial lateral flow-based assays require about 1 10
7
CFU.
The application of improved fluorogenic or colorimetric
substrates in ELISA and nanoparticles and fluorophores
in lateral flow-based assays can improve the assay sen-
sitivity significantly, i.e. a reduced detection limit.
Remarkably, the most improvement in immunoassays
in recent years has been observed in biosensor
12 L. LOPES-LUZ ET AL.
(immunosensor) platforms, where a detection limit of
110
2
110
3
CFU was achieved for L. monocyto-
genes. However, such sensitive assays are more vulner-
able to inhibitors, background microflora, and food
matrix thus sample preparation needs extreme care
which, therefore, may prolong the assay time (i.e. sam-
ple-to-result) and cost per test.
It is generally accepted that most food products,
especially those that are ready-to-eat are free of L.
monocytogenes or its level is very low and a majority of
aforementioned assay platforms are unable to provide
test results when applied to food samples directly.
Therefore, an enrichment step, either growth-based or
physical concentration method (IMS, centrifugation) is
needed to meet the detection threshold. The growth-
based enrichment method is common since it helps
resuscitate injured or stressed bacterial cells, increase
bacterial numbers, dilute inhibitors and reduce back-
ground microbial competitors providing reliable assay
results, but prolongs the assay time. However, the
length of enrichment can be shortened with an assay
that has a very low detection limit. On the other hand,
IMS is useful but samples containing high levels of food
particles or low pH may interfere with bacterial enrich-
ment efforts. Irrespective of detection platforms used, a
highly efficient sample preparation step is critical for
success especially for those that promote resuscitation
and growth and remove interfering agents from the
test samples to meet the detection threshold for
that assay.
Several immunoassay platforms were presented in
this review, as well as improvements that have been
studied and applied in combination with these
immunological techniques. This combination has
increased the specificity and sensitivity of the techni-
ques, making them important tools in listeriosis diagno-
sis, control, and prevention. However, the
methodologies still face challenges in the removal of
food matrices and inhibitors as part of efficient sample
preparation. Furthermore, due to the low sensitivity of
most tests, sample enrichment steps are imperative to
increase bacterial concentration to a detectable level.
Thus, there is a need for further investigation on meth-
odology optimisation and the selection of more specific
antibodies for the development of a more effective,
fast, and accessible method.
Disclosure statement
The authors declare no conflict of interest.
Funding
The work in the authorslaboratories was supported by
funds from the Brazilian Coordination for the Improvement
of Higher Education Personnel [CAPES grant # 88882.385457/
2007-01 20192023] and by the U.S. Department of
Agriculture, Agricultural Research Service, under Agreement
No. 59-8072-6-001. Any opinions, findings, conclusions, or
recommendations expressed in this publication are those of
the author(s) and do not necessarily reflect the view of the
U.S. Department of Agriculture.
References
Akin D, Sturgis J, Ragheb K, Sherman D, Burkholder K,
Robinson JP, Bhunia AK, Mohammed S, Bashir R. 2007.
Bacteria-mediated delivery of nanoparticles and cargo into
cells. Nat Nanotechnol. 2(7):441449.
Asano K, Sashinami H, Osanai A, Asano Y, Nakane A. 2011.
Autolysin amidase of Listeria monocytogenes promotes effi-
cient colonization of mouse hepatocytes and enhances
host immune response. Int J Med Microbiol. 301(6):
480487.
Auriemma C, Viscardi M, Tafuri S, Pavone LM, Capuano F,
Rinaldi L, Della Morte R, Iovane G, Staiano N. 2010.
Integrin receptors play a role in the internalin B-depend-
ent entry of Listeria monocytogenes into host cells. Cell
Mol Biol Lett. 15(3):496506.
Bai X, Liu D, Xu L, Tenguria S, Drolia R, Gallina NLF, Cox AD,
Koo O-K, Bhunia AK. 2021. Biofilm-isolated Listeria monocy-
togenes exhibits reduced systemic dissemination at the
early (1224 h) stage of infection in a mouse model. NPJ
Biofilm Microbi. 7(1):18.
Banada PP, Bhunia AK. 2008. Antibodies and immunoassays
for detection of bacterial pathogens. In: Zourob M, Elwary
S, Turner A, editors. Principles of bacterial detection: bio-
sensors, recognition receptors and microsystems.
Manchester: Cambridge University; p. 567602.
Banerjee R, Jaiswal A. 2018. Recent advances in nanoparticle-
based lateral flow immunoassay as a point-of-care diag-
nostic tool for infectious agents and diseases. Analyst.
143(9):19701996.
Beauchamp S, DAuria S, Pennacchio A, Lacroix M. 2012. A
new competitive fluorescence immunoassay for detection
of Listeria monocytogenes. Anal Methods. 4(12):41874192.
Becattini S, Littmann ER, Carter RA, Kim SG, Morjaria SM, Ling
L, Gyaltshen Y, Fontana E, Taur Y, Leiner IM, et al. 2017.
Commensal microbes provide first line defense against
Listeria monocytogenes infection. J Exp Med. 214(7):
19731989.
Bhunia AK, Ball PH, Fuad AT, Kurz BW, Emerson JW, Johnson
MG. 1991. Development and characterization of a mono-
clonal antibody specific for Listeria monocytogenes and
Listeria innocua. Infect Immun. 59(9):31763184.
Bhunia AK, Johnson MG. 1992. Monoclonal antibody specific
for Listeria monocytogenes associated with a 66-kilodalton
cell surface antigen. Appl Environ Microbiol. 58(6):
19241929.
Bhunia AK. 1997. Antibodies to Listeria monocytogenes. Crit
Rev Microbiol. 23(2):77107.
CRITICAL REVIEWS IN MICROBIOLOGY 13
Bhunia AK. 2008. Biosensors and bio-based methods for the
separation and detection of foodborne pathogens. Adv
Food Nutr Res. 54:144.
Bhunia AK. 2014. One day to one hour: how quickly can
foodborne pathogens be detected? Future Microbiol. 9(8):
935946.
Bhunia AK. 2018. Listeria monocytogenes. In: Bhunia AK, edi-
tor. Foodborne microbial pathogens: mechanisms and
pathogenesis. New York (NY): Springer; p. 229248.
Bierne H, Cossart P. 2002. InIB, a surface protein of Listeria
monocytogenes that behaves as an invasin and a growth
factor. J Cell Sci. 115(17):33573367.
Bierne H, Milohanic E, Kortebi M. 2018. To be cytosolic or
vacuolar: the double life of Listeria monocytogenes. Front
Cell Infect Microbiol. 8:136.
Bla
zkov
a M, Koets M, Rauch P, van Amerongen A. 2009.
Development of a nucleic acid lateral flow immunoassay
for simultaneous detection of Listeria spp. and
Listeriamonocytogenes in food. Eur Food Res Technol.
229(6):867874.
Boivin T, Elmgren C, Brooks BW, Huang H, Pagotto F, Lin M.
2016. Expression of surface protein LapB by a wide spec-
trum of Listeria monocytogenes serotypes as demonstrated
with anti-LapB monoclonal antibodies. Appl Environ
Microbiol. 82(22):67686778.
Boonham N, Kreuze J, Winter S, van der Vlugt R, Bergervoet
J, Tomlinson J, Mumford R. 2014. Methods in virus diag-
nostics: from ELISA to next generation sequencing. Virus
Res. 186:2031.
Bou Ghanem EN, Jones GS, Myers-Morales T, Patil PD,
Hidayatullah AN, DOrazio SE. 2012. InlA promotes dissem-
ination of Listeria monocytogenes to the mesenteric lymph
nodes during food borne infection of mice. PLoS Pathog.
8(11):e1003015.
Bruno JG, Phillips T, Montez T, Garcia A, Sivils JC, Mayo MW,
Greis A. 2015. Development of a fluorescent enzyme-
linked DNA aptamer-magnetic bead sandwich assay and
portable fluorometer for sensitive and rapid Listeria detec-
tion. J Fluoresc. 25(1):173183.
Bubert A, Kuhn M, Goebel W, Kohler S. 1992. Structural and
functional properties of the p60 proteins from different
Listeria species. J Bacteriol. 174(24):81668171.
Buchanan RL, Gorris LGM, Hayman MM, Jackson TC, Whiting
RC. 2017. A review of Listeria monocytogenes: an update
on outbreaks, virulence, dose-response, ecology, and risk
assessments. Food Control. 75:113.
Burkholder KM, Bhunia AK. 2010. Listeria monocytogenes uses
Listeria adhesion protein (LAP) to promote bacterial trans-
epithelial translocation, and induces expression of LAP
receptor Hsp60. IAI. 78(12):50625073.
Burkholder KM, Kim K-P, Mishra K, Medina S, Hahm B-K, Kim
H, Bhunia AK. 2009. Expression of LAP, a SecA2-dependent
secretory protein, is induced under anaerobic environ-
ment. Microbes Infect. 11(1011):859867.
Cabanes D, Dehoux P, Dussurget O, Frangeul L, Cossart P.
2002. Surface proteins and the pathogenic potential of
Listeria monocytogenes. Trends Microbiol. 10(5):238245.
Cabanes D, Sousa S, Cebria A, Lecuit M, Garcia-del Portillo F,
Cossart P. 2005. Gp96 is a receptor for a novel Listeria
monocytogenes virulence factor, Vip, a surface protein.
Embo J. 24(15):28272838.
Camejo A, Carvalho F, Reis O, Leitao E, Sousa S, Cabanes D.
2011. The arsenal of virulence factors deployed by Listeria
monocytogenes to promote its cell infection cycle.
Virulence. 2(5):379394.
Carroll SA, Hain T, Technow U, Darji A, Pashalidis P, Joseph
SW, Chakraborty T. 2003. Identification and characteriza-
tion of a peptidoglycan hydrolase, MurA, of Listeria mono-
cytogenes, a muramidase needed for cell separation. J
Bacteriol. 185(23):68016808.
Cavaiuolo M, Paramithiotis S, Drosinos EH, Ferrante A. 2013.
Development and optimization of an ELISA based method
to detect Listeria monocytogenes and Escherichia coli O157
in fresh vegetables. Anal Methods. 5(18):46224627.
Chen J-Q, Healey S, Regan P, Laksanalamai P, Hu Z. 2017.
PCR-based methodologies for detection and characteriza-
tion of Listeria monocytogenes and Listeria ivanovii in foods
and environmental sources. Food Sci Human Wellness.
6(2):3959.
Chen R, Huang X, Xu H, Xiong Y, Li Y. 2015. Plasmonic
enzyme-linked immunosorbent assay using nanospherical
brushes as a catalase container for colorimetric detection
of ultralow concentrations of Listeria monocytogenes. ACS
Appl Mater Interfaces. 7(51):2863228639.
Cheng C, Dong Z, Han X, Sun J, Wang H, Jiang L, Yang Y, Ma
T, Chen Z, Yu J, et al. 2017. Listeria monocytogenes 10403S
arginine repressor ArgR finely tunes arginine metabolism
regulation under acidic conditions. Front Microbiol. 8:112.
Cheng MI, Chen C, Engstr
om P, Portnoy DA, Mitchell G. 2018.
Actin-based motility allows Listeria monocytogenes to
avoid autophagy in the macrophage cytosol. Cell
Microbiol. 20(9):e12854.
Cho IH, Bhunia A, Irudayaraj J. 2015. Rapid pathogen detec-
tion by lateral-flow immunochromatographic assay with
gold nanoparticle-assisted enzyme signal amplification. Int
J Food Microbiol. 206:6066.
Cho I-H, Irudayaraj J. 2013. Lateral-flow enzyme immunocon-
centration for rapid detection of Listeria monocytogenes.
Anal Bioanal Chem. 405(10):33133319.
Coelho C, Brown L, Maryam M, Vij R, Smith DFQ, Burnet MC,
Kyle JE, Heyman HM, Ramirez J, Prados-Rosales R, et al.
2019. Listeria monocytogenes virulence factors, including
listeriolysin O, are secreted in biologically active extracellu-
lar vesicles. J Biol Chem. 294(4):12021217.
Cossart P, Helenius A. 2014. Endocytosis of viruses and bac-
teria. Cold Spring Harbor Perspect Biol. 6(8):a016972.
Cotter PD, Draper LA, Lawton EM, Daly KM, Groeger DS,
Casey PG, Ross RP, Hill C. 2008. Listeriolysin S, a novel
peptide haemolysin associated with a subset of lineage I
Listeria monocytogenes. PLoS Pathog. 4(9):e1000144.
Coutu JV, Morissette C, Auria SD, Lacroix M. 2014.
Development of a highly specific sandwich ELISA for the
detection of Listeria monocytogenes, an important food-
borne pathogen. Microbiol Res Int. 2:4652.
Craig AM, Dotters-Katz S, Kuller JA, Thompson JL. 2019.
Listeriosis in pregnancy: a review. Obstet Gynecol Survey.
74(6):362368.
Daugherty S, Low MG. 1993. Cloning, expression, and muta-
genesis of phosphatidylinositol-specific phospholipase C
from Staphylococcus aureus: a potential staphylococcal
virulence factor. Infect Immun. 61(12):50785089.
David DJ, Pagliuso A, Radoshevich L, Nahori MA, Cossart P.
2018. Lmo1656 is a secreted virulence factor of Listeria
14 L. LOPES-LUZ ET AL.
monocytogenes that interacts with the sorting nexin 6-BAR
complex. J Biol Chem. 293(24):92659276.
Day JB, Hammack TS. 2019. Immuno-detection and differenti-
ation of Listeria monocytogenes and Listeria ivanovii in
stone fruits. J Appl Microbiol. 127(6):18481858.
Diercks GF, Pas HH, Jonkman MF. 2017. Immunofluorescence
of autoimmune bullous diseases. Surg Pathol Clin. 10(2):
505512.
Dramsi S, Dehoux P, Lebrun M, Goossens PL, Cossart P. 1997.
Identification of four new members of the internalin multi-
gene family of Listeria monocytogenes EGD. Infect
Immun. 65(5):16151625. DOI:10.1128/IAI.65.5.1615-1625.
1997.
Drolia R, Amalaradjou MAR, Ryan V, Tenguria S, Liu D, Bai X,
Xu L, Singh AK, Cox AD, Bernal-Crespo V, et al. 2020.
Receptor-targeted engineered probiotics mitigate lethal
Listeria infection. Nat Commun. 11(1):6344.
Drolia R, Bhunia AK. 2019. Crossing the intestinal barrier via
Listeria adhesion protein and internalin A. Trends
Microbiol. 27(5):408425.
Drolia R, Tenguria S, Durkes AC, Turner JR, Bhunia AK. 2018.
Listeria adhesion protein induces intestinal epithelial bar-
rier dysfunction for bacterial translocation. Cell Host
Microbe. 23(4):470484.
Du XJ, Zang YX, Liu HB, Li P, Wang S. 2018. Recombinase
polymerase amplification combined with lateral flow strip
for Listeria monocytogenes detection in food. J Food Sci.
83(4):10411047.
Dussurget O, Cabanes D, Dehoux P, Lecuit M, Buchrieser C,
Glaser P, Cossart P. 2002. Listeria monocytogenes bile salt
hydrolase is a PrfA-regulated virulence factor involved in
the intestinal and hepatic phases of listeriosis. Mol
Microbiol. 45(4):10951106.
Dussurget O, Pizarro-Cerda J, Cossart P. 2004. Molecular
determinants of Listeria monocytogenes virulence. Annu
Rev Microbiol. 58(1):587610.
Etty MC, DAuria S, Fraschini C, Salmieri S, Lacroix M. 2019.
Effect of the optimized selective enrichment medium on
the expression of the p60 protein used as Listeria monocy-
togenes antigen in specific sandwich ELISA. Res Microbiol.
170(45):182191.
Etty MC, DAuria S, Shankar S, Salmieri S, Coutu J, Baraketi A,
Jamshidan M, Fraschini C, Lacroix M. 2020. New immobil-
ization method of anti-PepD monoclonal antibodies for
the detection of Listeria monocytogenes p60 protein part
A: optimization of a crosslinked film support based on chi-
tosan and cellulose nanocrystals (CNC). Reactive
Functional Polymers. 146:104313104313.
Ferreira V, Wiedmann M, Teixeira P, Stasiewicz MJ. 2014.
Listeria monocytogenes persistence in food-associated envi-
ronments: epidemiology, strain characteristics, and impli-
cations for public health. J Food Prot. 77(1):150170.
Fisher M, Atiya-Nasagi Y, Simon I, Gordin M, Mechaly A,
Yitzhaki S. 2009. A combined immunomagnetic separation
and lateral flow method for a sensitive on-site detection
of Bacillus anthracis spores-assessment in water and dairy
products. Lett Appl Microbiol. 48(4):413418.
Gahan C, Hill C. 2014. Listeria monocytogenes: survival and
adaptation in the gastrointestinal tract. Front Cell Infect
Microbiol. 4:9.
Gandhi M, Chikindas ML. 2007. Listeria: a foodborne patho-
gen that knows how to survive. Int J Food Microbiol.
113(1):115.
Gasanov U, Hughes D, Hansbro PM. 2005. Methods for the
isolation and identification of Listeria spp. and Listeria
monocytogenes: a review. FEMS Microbiol Rev. 29(5):
851875.
Gelb
ı
cov
a T, Pant˚
u
cek R, Karp
ı
skov
a R. 2016. Virulence factors
and resistance to antimicrobials in Listeria monocytogenes
serotype 1/2c isolated from food. J Appl Microbiol. 121(2):
569576.
Gene RW, Kumaran J, Aroche C, van Faassen H, Hall JC,
MacKenzie CR, Arbabi-Ghahroudi M. 2015. High affinity
anti-Internalin B VHH antibody fragments isolated from
naturally and artificially immunized repertoires. J Immunol
Methods. 416:2939.
Geng T, Kim KP, Gomez R, Sherman DM, Bashir R, Ladisch
MR, Bhunia AK. 2003. Expression of cellular antigens of
Listeria monocytogenes that react with monoclonal anti-
bodies C11E9 and EM-7G1 under acid-, salt- or tempera-
ture-induced stress environments. J Appl Microbiol. 95(4):
762772.
Geng T, Morgan MT, Bhunia AK. 2004. Detection of low levels
of Listeria monocytogenes cells by using a fiber-optic
immunosensor. Appl Environ Microbiol. 70(10):61386146.
Gessain G, Tsai Y-H, Travier L, Bonazzi M, Grayo S, Cossart P,
Charlier C, Disson O, Lecuit M. 2015. PI3-kinase activation
is critical for host barrier permissiveness to Listeria mono-
cytogenes. J Exp Med. 212(2):165183.
Ghosh P, Halvorsen EM, Ammendolia DA, Mor-Vaknin N,
ORiordan MX, Brumell JH, Markovitz DM, Higgins DE.
2018. Invasion of the brain by Listeria monocytogenes is
mediated by InlF and host cell vimentin. mBio. 9(1):
e00160-18.
Glaser P, Frangeul L, Buchrieser C, Rusniok C, Amend A,
Baquero F, Berche P, Bloecker H, Brandt P, Chakraborty T,
et al. 2001. Comparative genomics of Listeria species.
Science. 294(5543):849852.
Gonzalez-Escobedo G, Gunn JS. 2013. Gallbladder epithelium
as a niche for chronic salmonella carriage. Infect Immun.
81(8):29202930.
Grundling A, Gonzalez MD, Higgins DE. 2003. Requirement of
the Listeria monocytogenes broad-range phospholipase PC-
PLC during infection of human epithelial cells. J Bacteriol.
185(21):62956307.
Hahm BK, Bhunia AK. 2006. Effect of environmental stresses
on antibody-based detection of Escherichia coli O157:H7,
Salmonella enterica serotype Enteritidis and Listeria mono-
cytogenes. J Appl Microbiol. 100(5):10171027.
Hahm BK, Kim H, Singh AK, Bhunia AK. 2015. Pathogen
enrichment device (PED) enables one-step growth, enrich-
ment and separation of pathogen from food matrices for
detection using bioanalytical platforms. J Microbiol
Methods. 117:6473.
Hearty S, Leonard P, Quinn J, OKennedy R. 2006. Production,
characterisation and potential application of a novel
monoclonal antibody for rapid identification of virulent
Listeria monocytogenes. J Microbiol Methods. 66(2):
294312.
Heo SA, Nannapaneni R, Story RP, Johnson MG. 2007.
Characterization of new hybridoma clones producing
monoclonal antibodies reactive against both live and
CRITICAL REVIEWS IN MICROBIOLOGY 15
heat-killed Listeria monocytogenes. J Food Sci. 72(1):M008-
15.
Hoge J, Yan I, J
anner N, Schumacher V, Chalaris A, Steinmetz
OM, Engel DR, Scheller J, Rose-John S, Mittr
ucker H-W.
2013. IL-6 controls the Innate immune response against
Listeria monocytogenes via classical IL-6 signaling. J
Immunol. 190(2):703711.
Horn N, Bhunia AK. 2018. Food-associated stress primes
foodborne pathogens for the gastrointestinal phase of
infection. Front Microbiol. 9:1962.
Hsieh HV, Dantzler JL, Weigl BH. 2017. Analytical tools to
improve optimization procedures for lateral flow assays.
Diagnostics. 7(2):29.
Hurley D, Luque-Sastre L, Parker CT, Huynh S, Eshwar AK,
Nguyen SV, Andrews N, Moura A, Fox EM, Jordan K, et al.
2019. Whole-genome sequencing-based characterization
of 100 Listeria monocytogenes isolates collected from food
processing environments over a four-year period. Appl
Environ Sci. 4(4):114.
Ito K, Yamamoto T, Oyama Y, Tsuruma R, Saito E, Saito Y,
Ozu T, Honjoh T, Adachi R, Sakai S, et al. 2016. Food aller-
gen analysis for processed food using a novel extraction
method to eliminate harmful reagents for both ELISA and
lateral-flow tests. Anal Bioanal Chem. 408(22):59735984.
Jacinto MJ, Trabuco JRC, Vu BV, Garvey G, Khodadady M,
Azevedo AM, Aires-Barros MR, Chang L, Kourentzi K,
Litvinov D, Willson RC. 2018. Enhancement of lateral flow
assay performance by electromagnetic relocation of
reporter particles. PLoS One. 13(1):e0186782.
Jagadeesan B, Fleishman Littlejohn AE, Amalaradjou MAR,
Singh AK, Mishra KK, La D, Kihara D, Bhunia AK. 2011. N-
Terminal Gly
224
- Gly
411
domain in Listeria adhesion pro-
tein interacts with host receptor Hsp60. PLoS One. 6(6):
e20694.
Jagadeesan B, Koo OK, Kim KP, Burkholder KM, Mishra KK,
Aroonnual A, Bhunia AK. 2010. LAP, an alcohol acetalde-
hyde dehydrogenase enzyme in Listeria promotes bacterial
adhesion to enterocyte-like Caco-2 cells only in patho-
genic species. Microbiology. 156(9):27822795.
Jain B, Lambe U, Tewari A, Kadian SK, Prasad M. 2018.
Development of a rapid test for detection of foot-and-
mouth disease virus specific antibodies using gold nano-
particles. Virusdisease. 29(2):192198.
Jaradat ZW, Bhunia AK. 2002. Glucose and nutrient concen-
trations affect the expression of a 104-kilodalton Listeria
adhesion protein in Listeria monocytogenes. Appl Environ
Microbiol. 68(10):48764883.
Jeyaletchumi P, Tunung R, Margaret SP, Son R, Farinazleen
MG, Cheah YK. 2010. Detection of Listeria monocytogenes
in foods. Int Food Res J. 17(1):111.
Jiang N, Ahmed R, Damayantharan M,
Unal B, Butt H, Yetisen
AK. 2019. Lateral and vertical flow assays for point-of-care
diagnostics. Adv Healthcare Mater. 8(14):
19002441900219.
Johansson J, Freitag NE. 2019. Regulation of Listeria monocy-
togenes virulence. In: Fischetti VA, Novick RP, Ferretti JJ,
Portnoy, DA, Braunstein M, Rood JI, editor. Gram-positive
pathogens. Hoboken (NJ): John Wiley & Sons, Inc.; p.
836850.
Karamonov
a L, Bla
zkov
a M, Fukal L, Rauch P, Greifov
aM,
Hor
akov
a K, Tom
a
ska M, Roubal P, Brett GM, Wyatt GM.
2003. Development of an ELISA specific for Listeria
monocytogenes using a polyclonal antibody raised against
a cell extract containing internalin B. Food Agri Immunol.
15(34):167182.
Karatzas K-AG, Suur L, OByrne CP. 2012. Characterization of
the intracellular glutamate decarboxylase system: analysis
of its function, transcription, and role in the acid resist-
ance of various strains of Listeria monocytogenes. Appl
Environ Microbiol. 78(10):35713579.
Kim KP, Jagadeesan B, Burkholder KM, Jaradat ZW, Wampler
JL, Lathrop AA, Morgan MT, Bhunia AK. 2006. Adhesion
characteristics of Listeria adhesion protein (LAP)-express-
ing Escherichia coli to Caco-2 cells and of recombinant
LAP to eukaryotic receptor Hsp60 as examined in a surface
plasmon resonance sensor. FEMS Microbiol Lett. 256(2):
324332.
Kim KP, Singh AK, Bai X, Leprun L, Bhunia AK. 2015. Novel
PCR assays complement laser biosensor-based method
and facilitate Listeria species detection from food. Sensors.
15(9):2267222691.
King MT, Huh I, Shenai A, Brooks TM, Brooks CL. 2018.
Structural basis of V H H-mediated neutralization of the
food-borne pathogen Listeria monocytogenes. J Biol Chem.
293(35):1362613635.
Kitao T, Miyoshi-Akiyama T, Shimada K, Tanaka M, Narahara
K, Saito N, Kirikae T. 2010. Development of an
immunochromatographic assay for the rapid detection of
AAC(6)-Iae-producing multidrug-resistant Pseudomonas
aeruginosa. J Antimicrob Chemother. 65(7):13821386.
Koczula KM, Gallotta A. 2016. Lateral flow assays. Essays
Biochem. 60(1):111120.
Koo OK, Aroonnual A, Bhunia AK. 2011. Human heat-shock
protein 60 receptor-coated paramagnetic beads show
improved capture of Listeria monocytogenes in the pres-
ence of other Listeria in food. J Appl Microbiol. 111(1):
93104.
Koo OK, Liu Y, Shuaib S, Bhattacharya S, Ladisch MR, Bashir
R, Bhunia AK. 2009. Targeted capture of pathogenic bac-
teria using a mammalian cell receptor coupled with
dielectrophoresis on a biochip. Anal Chem. 81(8):
30943101.
Kortebi M, Milohanic E, Mitchell G, P
echoux C, Prevost M-C,
Cossart P, Bierne H. 2017. Listeria monocytogenes switches
from dissemination to persistence by adopting a vacuolar
lifestyle in epithelial cells. PLoS Pathog. 13(11):e1006734.
K
uhn S, Enninga J. 2020. The actin comet guides the way:
how Listeria actin subversion has impacted cell biology,
infection biology and structural biology. Cell Microbiol.
22(4):e13190.
Lamont RF, Sobel J, Mazaki-Tovi S, Kusanovic JP, Vaisbuch E,
Kim SK, Uldbjerg N, Romero R. 2011. Listeriosis in human
pregnancy: a systematic review. J Perinatal Med. 39(3):
227236.
Lathrop AA, Bailey TW, Kwang-Pyo K, Bhunia AK. 2014.
Pathogen-specific antigen target for production of anti-
bodies produced by comparative genomics. Antibody
Technol J. 4:13.
Lathrop AA, Banada PP, Bhunia AK. 2008. Differential expres-
sion of InlB and ActA in Listeria monocytogenes in selective
and nonselective enrichment broths. J Appl Microbiol.
104(3):627639.
Lathrop AA, Jaradat ZW, Haley T, Bhunia AK. 2003.
Characterization and application of a Listeria
16 L. LOPES-LUZ ET AL.
monocytogenes reactive monoclonal antibody C11E9 in a
resonant mirror biosensor. J Immunol Methods. 281(12):
119128.
Law JW-F, Ab Mutalib N-S, Chan K-G, Lee L-H. 2014. Rapid
methods for the detection of foodborne bacterial patho-
gens: principles, applications, advantages and limitations.
Front Microbiol. 5:770770.
Le Monnier A, Abachin E, Beretti JL, Berche P, Kayal S. 2011.
Diagnosis of Listeria monocytogenes meningoencephalitis
by real-time PCR for the hly gene. J Clin Microbiol. 49(11):
39173923.
Lebreton A, Lakisic G, Job V, Fritsch L, Tham TN, Camejo A,
Mattei PJ, Regnault B, Nahori MA, Cabanes D, et al. 2011.
A bacterial protein targets the BAHD1 chromatin complex
to stimulate type III interferon response. Science.
331(6022):13191321.
Leclercq A, Moura A, Vales G, Tessaud-Rita N, Aguilhon C,
Lecuit M. 2019. Listeria thailandensis sp. nov. Int J Syst
Evol Microbiol. 69(1):7481.
Lee B-H, Cole S, Badel-Berchoux S, Guillier L, Felix B,
Krezdorn N, H
ebraud M, Bernardi T, Sultan I, Piveteau P.
2019. Biofilm formation of Listeria monocytogenes strains
under food processing environments and pan-genome-
wide association study. Front Microbiol. 10:2698.
Leonard P, Hearty S, Wyatt G, Quinn J, OKennedy R. 2005.
Development of a surface plasmon resonance-based
immunoassay for Listeria monocytogenes. J Food Prot.
68(4):728735.
Leszczy
nska J, MasŁowska J, Owczarek A, Kucharska U. 2013.
Determination of aflatoxins in food products by the ELISA
method. Czech J Food Sci. 19(No. 1):812.
Li F, Li F, Luo D, Lai W, Xiong Y, Xu H. 2018. Biotin-exposure-
based immunomagnetic separation coupled with nucleic
acid lateral flow biosensor for visibly detecting viable
Listeria monocytogenes. Anal Chim Acta. 1017:4856.
Li J, Jing L, Song Y, Zhang J, Chen Q, Wang B, Xia X, Han Q.
2018. Rapid detection of rongalite via a sandwich lateral
flow strip assay using a pair of aptamers. Nanoscale Res
Lett. 13(1):296.
Li Q, Zhang S, Cai Y, Yang Y, Hu F, Liu X, He X. 2017. Rapid
detection of Listeria monocytogenes using fluorescence
immunochromatographic assay combined with immuno-
magnetic separation technique. Int J Food Sci Technol.
52(7):15591566.
Lind
en SK, Bierne H, Sabet C, Png CW, Florin TH, McGuckin
MA, Cossart P. 2008. Listeria monocytogenes internalins
bind to the human intestinal mucin MUC2. Arch Microbiol.
190(1):101104.
Lingala SMD, Ghany M. 2015. The p60 and NamA autolysins
from Listeria monocytogenes contribute to host coloniza-
tion and induction of protective memory. Cell Microbiol.
17(2):147163.
Liu A, Xiong Q, Shen L, Li W, Zeng Z, Li C, Liu S, Liu Y, Han
G. 2017. A sandwich-type ELISA for the detection of
Listeria monocytogenes using the well-oriented single
chain Fv antibody fragment. Food Control. 79:156161.
Liu H, Du X-J, Zang Y-X, Li P, Wang S. 2017. SERS-based lat-
eral flow strip biosensor for simultaneous detection of
Listeria monocytogenes and Salmonella enterica serotype
enteritidis. J Agric Food Chem. 65(47):1029010299.
Liu Y, Orsi RH, Gaballa A, Wiedmann M, Boor KJ, Guariglia-
Oropeza V. 2019. Systematic review of the Listeria
monocytogenes rB regulon supports a role in stress
response, virulence and metabolism. Future Microbiol.
14(9):801828.
Lungu B, Ricke SC, Johnson MG. 2009. Growth, survival, pro-
liferation and pathogenesis of Listeria monocytogenes
under low oxygen or anaerobic conditions: a review.
Anaerobe. 15(12):717.
Luo X, Cai X. 2012. A combined use of autolysin p60 and lis-
teriolysin O antigens induces high protective immune
responses against Listeria monocytogenes infection. Curr
Microbiol. 65(6):813818.
Lv X, Huang Y, Liu D, Liu C, Shan S, Li G. Duan M, Lai W.
2019. Multicolor and ultrasensitive ELISA based on fluores-
cence hybrid chain reaction for simultaneous detection of
pathogens. J Agric Food Chem. 67(33):93909398.
Malica L, Zhanga X, Brassarda D, Climea L, Daouda J,
Luebbertb C, Barrereb V, Boutina A, Bidawidb S, Farberb J,
et al. 2015. Polymer-based microfluidic chip for rapid and
efficient immunomagnetic capture and release of Listeria
monocytogenes. Lab Chip. 0:13.
Mao X, Xu H, Zeng Q, Zeng L, Liu G. 2009. Molecular bea-
con-functionalized gold nanoparticles as probes in dry-
reagent strip biosensor for DNA analysis. Chem Commun.
81(21):30653067.
Markkula A, Lindstr
om M, Korkeala H. 2011. Listeria monocy-
togenes serotypes 1/2c and 3c possess inlH. Foodborne
Pathog Dis. 8(10):11251129.
Mendonca M, Conrad N, Conceicao F, Moreira A, da Silva W,
Aleixo J, Bhunia A. 2012. Highly specific fiber optic immu-
nosensor coupled with immunomagnetic separation for
detection of low levels of Listeria monocytogenes and L.
ivanovii. BMC Microbiol. 12(1):275.
Mendonc¸a M, Moreira GMSG, Conceic¸~
ao FR, Hust M,
Mendonc¸a KS, Moreira
^
AN, Franc¸a RC, Silva WPD, Bhunia
AK, Aleixo JAG. 2016. Fructose 1,6-bisphosphate aldolase,
a novel immunogenic surface protein on Listeria species.
PLoS One. 11(8):e016054420.
Meyer C, Fredriksson-Ahomaa M, Sperner B, M
artlbauer E.
2011. Detection of Listeria monocytogenes in pork and
beef using the VIDASV
RLMO2 automated enzyme linked
immunoassay method. Meat Sci. 88(3):594596.
Nanduri V, Bhunia AK, Tu SI, Paoli GC, Brewster JD. 2007. SPR
biosensor for the detection of L. monocytogenes using
phage-displayed antibody. Biosens Bioelectron. 23(2):
248252.
Nanduri V, Kim G, Morgan MT, Ess D, Hahm BK, Kothapalli A,
Valadez A, Geng T, Bhunia AK. 2006. Antibody immobiliza-
tion on waveguides using a flow-through system shows
improved Listeria monocytogenes detection in an auto-
mated fiber optic biosensor: RAPTOR (TM). Sensors. 6(8):
808822.
Nannapaneni R, Story R, Bhunia AK, Johnson MG. 1998.
Unstable expression and thermal instability of a species-
specific cell surface epitope associated with a 66-kilodal-
ton antigen recognized by monoclonal antibody EM-7G1
within serotypes of Listeria monocytogenes grown in non-
selective and selective broths. Appl Environ Microbiol.
64(8):30703074.
Nikitas G, Deschamps C, Disson O, Niault T, Cossart P, Lecuit
M. 2011. Transcytosis of Listeria monocytogenes across the
intestinal barrier upon specific targeting of goblet cell
accessible E-cadherin. J Exp Med. 208(11):22632277.
CRITICAL REVIEWS IN MICROBIOLOGY 17
N
u~
nez-Montero K, Leclercq A, Moura A, Vales G, Peraza J,
Pizarro-Cerd
a J, Lecuit M. 2018. Listeria costaricensis sp.
nov. Int J Sys Evol Microbiol. 68(3):844850.
Ohk S-H, Bhunia AK. 2013. Multiplex fiber optic biosensor for
detection of Listeria monocytogenes,Escherichia coli O157:
H7 and Salmonella enterica from ready-to-eat meat sam-
ples. Food Microbiol. 33(2):166171.
Ohk SH, Koo OK, Sen T, Yamamoto CM, Bhunia AK. 2010.
Antibody-aptamer functionalized fibre-optic biosensor for
specific detection of Listeria monocytogenes from food. J
Appl Microbiol. 109(3):808817.
Ooi ST, Lorber B. 2005. Gastroenteritis due to Listeria mono-
cytogenes. Clin Infect Dis. 40(9):13271332.
Oravec M, Sasinkov V, Tomanova K, Gal L, Parciova S, Huck
CW. 2018. In-situ surface-enhanced Raman scattering and
FT-Raman spectroscopy of black prints. Vib Spectrosc. 94:
1621.
Ortega FE, Koslover EF, Theriot JA. 2019. Listeria monocyto-
genes cell-to-cell spread in epithelia is heterogeneous and
dominated by rare pioneer bacteria. Elife. 8:126.
Ortega FE, Rengarajan M, Chavez N, Radhakrishnan P,
Gloerich M, Bianchini J, Siemers K, Luckett WS, Lauer P,
Nelson WJ, et al. 2017. Adhesion to the host cell surface is
sufficient to mediate Listeria monocytogenes entry into epi-
thelial cells. Mol Biol Cell. 28(22):29452957.
Osanai A, Li SJ, Asano K, Sashinami H, Hu DL, Nakane A.
2013. Fibronectin-binding protein, FbpA, is the adhesin
responsible for pathogenesis of Listeria monocytogenes
infection. Microbiol Immunol. 57(4):253262.
Owais M, Kazmi S, Tufail S, Zubair S. 2014. An alternative
chemical redox method for the production of bispecific
antibodies: implication in rapid detection of food borne
pathogens. PLoS One. 9(3):e9125513.
P
agelow D, Chhatbar C, Beineke A, Liu X, Nerlich A, Van
Vorst K, Rohde M, Kalinke U, F
orster R, Halle S, et al.
2018. The olfactory epithelium as a port of entry in neo-
natal neurolisteriosis. Nat Commun. 9(1):4269.DOI:10.1038/
s41467-018-06668-2. PMC: 30323282.
Pandiripally VK, Westbrook DG, Sunki GR, Bhunia AK. 1999.
Surface protein p104 is involved in adhesion of Listeria
monocytogenes to human intestinal cell line, Caco-2. J
Med Microbiol. 48(2):117124.
Pang X, Wong C, Chung H-J, Yuk H-G. 2019. Biofilm forma-
tion of Listeria monocytogenes and its resistance to quater-
nary ammonium compounds in a simulated salmon
processing environment. Food Control. 98:200208.
Paoli GC, Brewster JD. 2007. A Listeria monocytogenes-specific
phage-displayed antibody fragment recognizes a cell sur-
face protein whose expression is regulated by physio-
logical conditions. J Rapid Methods Auto Microbiol. 15(1):
7791.
Paoli GC, Chen C-Y, Brewster JD. 2004. Single-chain Fv anti-
body with specificity for Listeria monocytogenes.J
Immunol Methods. 289(12):147155.
Paoli GC, Kleina LG, Brewster JD. 2007. Development of
Listeria monocytogenes-specific immunomagnetic beads
using a single-chain antibody fragment. Foodborne
Pathog Dis. 4(1):7483.
Pattabiraman G, Palasiewicz K, Visvabharathy L, Freitag NE,
Ucker DS. 2017. Apoptotic cells enhance pathogenesis of
Listeria monocytogenes. Microb Pathog. 105:218225.
Paudyal R, Barnes RH, Karatzas KAG. 2018. A novel approach
in acidic disinfection through inhibition of acid resistance
mechanisms; maleic acid-mediated inhibition of glutamate
decarboxylase activity enhances acid sensitivity of Listeria
monocytogenes. Food Microbiol. 69:96104.
Phelps CC, Vadia S, Arnett E, Tan Y, Zhang X, Pathak-Sharma
S, Gavrilin MA, Seveau S. 2018. Relative roles of listeriolysin
O, InlA, and InlB in Listeria monocytogenes uptake by host
cells. Infect Immun. 86(10):e00555-18.
Phraephaisarn C, Khumthong R, Takahashi H, Ohshima C,
Kodama K, Techaruvichit P, Vesaratchavest M, Taharnklaew
R, Keeratipibul S. 2017. A novel biomarker for detection of
Listeria species in food processing factory. Food Control.
73:10321038.
Pickard JM, Zeng MY, Caruso R, N
u~
nez G. 2017. Gut micro-
biota: role in pathogen colonization, immune responses,
and inflammatory disease. Immunol Rev. 279(1):7089.
Pouillot R, Klontz KC, Chen Y, Burall LS, Macarisin D, Doyle M,
Bally KM, Strain E, Datta AR, Hammack TS, et al. 2016.
Infectious dose of Listeria monocytogenes in outbreak
linked to ice cream, United States, 2015. Emerg Infect Dis.
22(12):21132119.
Poussin MA, Goldfine H. 2005. Involvement of Listeria mono-
cytogenes phosphatidylinositol-specific phospholipase C
and host protein kinase C in permeabilization of the
macrophage phagosome. Infect Immun. 73(7):44104413.
Prommajan K, Palaga T, Rengpipat S. 2016. Detection of
Listeria monocytogenes in foods using monoclonal anti-
body by dot-ELISA and multiplex-PCR. Int Food Res J.
23(6):27022709.
Pron B, Boumaila C, Jaubert F, Sarnacki S, Monnet J, Berche
P, Gaillard J. 1998. Comprehensive study of the intestinal
stage of listeriosis in a rat ligated ileal loop system. Infect
Immun. 66(2):747755.
Quereda JJ, Dussurget O, Nahori M-A, Ghozlane A, Volant S,
Dillies M-A, Regnault B, Kennedy S, Mondot S, Villoing B,
et al. 2016. Bacteriocin from epidemic Listeria strains alters
the host intestinal microbiota to favor infection. Proc Natl
Acad Sci USA. 113(20):57065711.
Quereda JJ, Rodr
ıguez-G
omez IM, Meza-Torres J, G
omez-
Laguna J, Nahori MA, Dussurget O, Carrasco L, Cossart P,
Pizarro-Cerd
a J. 2019. Reassessing the role of internalin B
in Listeria monocytogenes virulence using the epidemic
strain F2365. Clin Microbiol Infect. 25(2):252.e1e4.
Radoshevich L, Cossart P. 2018. Listeria monocytogenes:
towards a complete picture of its physiology and patho-
genesis. Nat Rev Microbiol. 16(1):3246.
Rajabian T, Gavicherla B, Heisig M, M
uller-Altrock S, Goebel
W, Gray-Owen SD, Ireton K. 2009. The bacterial virulence
factor InlC perturbs apical cell junctions and promotes
cell-to-cell spread of Listeria. Nat Cell Biol. 11(10):
12121218.
Rekha VB, Malik SVS, Chaudhari SP, Barbuddhe SB. 2006.
Listeriolysin O-based diagnosis of Listeria monocytogenes
infection in experimentally and naturally infected goats.
Small Ruminant Res. 66(13):7075.
Ribet D, Cossart P. 2015. How bacterial pathogens colonize
their hosts and invade deeper tissues. Microbes Infect.
17(3):173183.
Roberts BN, Chakravarty D, Gardner JC, Ricke SC, Donaldson
JR. 2020. Listeria monocytogenes response to anaerobic
environments. Pathogens. 9(3):210.
18 L. LOPES-LUZ ET AL.
Rodr
ıguez-Lorenzo L, Garrido-Maestu A, Bhunia AK, Espi~
na B,
Prado M, Di
eguez L, Abalde-Cela S. 2019. Gold Nanostars
for the detection of foodborne pathogens via surface-
enhanced Raman scattering combined with microfluidics.
ACS Appl Nano Mater. 2(10):60816086.
Ronholm J, van Faassen H, MacKenzie R, Zhang Z, Cao X, Lin
M. 2013. Monoclonal antibodies recognizing the surface
autolysin IspC of Listeria monocytogenes serotype 4b: epi-
tope localization, kinetic characterization, and cross-reac-
tion studies. PLoS One. 8(2):e55098.
Ryan S, Begley M, Gahan CGM, Hill C. 2009. Molecular charac-
terization of the arginine deiminase system in Listeria
monocytogenes: regulation and role in acid tolerance.
Environ Microbiol. 11(2):432445.
Sajid M, Kawde AN, Daud M. 2015. Designs, formats and
applications of lateral flow assay: a literature review. J
Saudi Chem Soc. 19(6):689705.
Santiago NI, Zipf A, Bhunia AK. 1999. Influence of tempera-
ture and growth phase on expression of a 104-kilodalton
Listeria adhesion protein in Listeria monocytogenes. Appl
Environ Microbiol. 65(6):27652769.
Schlech WF. 2019. Epidemiology and clinical manifestations
of Listeria monocytogenes infection. In: Fischetti VA, Novick
RP, Ferretti JJ, Portnoy DA. Braunstein M, Rood JI, editors.
Gram-positive pathogens. Hoboken (NJ): John Wiley &
Sons, Inc.; p. 793802.
Sharma H, Mutharasan R. 2013. Rapid and sensitive immuno-
detection of Listeria monocytogenes in milk using a novel
piezoelectric cantilever sensor. Biosens Bioelectron. 45(1):
158162.
Shen Y, Naujokas M, Park M, Ireton K. 2000. InIB-dependent
internalization of Listeria is mediated by the Met receptor
tyrosine kinase . Cell. 103(3):501510.
Shi C, Hohl TM, Ingrid L, Equinda MJ, Fan X, Pamer EG. 2011.
Ly6G þneutrophils are dispensable for defense against
systemic Listeria monocytogenes infection. J Immunol.
187(10):52935298.
Shi L, Wu F, Wen Y, Zhao F, Xiang J, Ma L. 2015. A novel
method to detect Listeria monocytogenes via superpara-
magnetic lateral flow immunoassay. Anal Bioanal Chem.
407(2):529535.
Shoukat S, Malik SVS, Rawool DB, Kumar A, Kumar S,
Shrivastava S, Das DP, Das S, Barbuddhe SB. 2013.
Comparison of indirect based ELISA by employing purified
LLO and its synthetic peptides and cultural method for
diagnosis of ovine listeriosis. Small Ruminant Res. 113(1):
301306.
Sim J, Hood D, Finnie L, Wilson M, Graham C, Brett M,
Hudson JA. 2002. Series of incidents of Listeria monocyto-
genes non-invasive febrile gastroenteritis involving ready-
to-eat meats. Lett Appl Microbiol. 35(5):409413.
Singh AK, Bhunia AK. 2018. Optical Biosensors in Foodborne
Pathogen Detection. 443. https://www.taylorfrancis.com/
chapters/edit/10.1201/9780429429934-21/optical-biosen-
sors-foodborne-pathogen-detection-atul-singh-arun-bhunia
Singh J, Sharma S, Nara S. 2015. Evaluation of gold nanopar-
ticle based lateral flow assays for diagnosis of enterobac-
teriaceae members in food and water. Food Chem. 170:
470483.
Smith AM, Tau NP, Smouse SL, Allam M, Ismail A, Ramalwa
NR, Disenyeng B, Ngomane M, Thomas J. 2019. Outbreak
of Listeria monocytogenes in South Africa, 20172018:
laboratory activities and experiences associated with
whole-genome sequencing analysis of isolates. Foodborne
Pathog Dis. 16(7):524530.
Smolsky J, Kaur S, Hayashi C, Batra SK, Krasnoslobodtsev AV.
2017. Immunoassay technologies for detection of disease
biomarkers. Biosensors. 7(4):721.
Sobyanin K, Sysolyatina E, Krivozubov M, Chalenko Y,
Karyagina A, Ermolaeva S, Calendar R. 2017. Naturally
occurring InlB variants that support intragastric Listeria
monocytogenes infection in mice. FEMS Microbiol Lett.
364(3):fnx011.
Stambach NR, Carr SA, Cox CR, Voorhees KJ. 2015. Rapid
detection of Listeria by bacteriophage amplification and
SERS-lateral flow immunochromatography. Viruses. 7(12):
66316641.
Suryawanshi RD, Malik SVS, Jayarao B, Chaudhari SP, Savage
E, Vergis J, Kurkure NV, Barbuddhe SB, Rawool DB. 2017.
Comparative diagnostic efficacy of recombinant LLO and
PI-PLC-based ELISAs for detection of listeriosis in animals.
J Microbiol Methods. 137:4045.
Swaminathan B, Gerner-Smidt P. 2007. The epidemiology of
human listeriosis. Microbes Infect. 9(10):12361243.
Tasara T, Stephan R. 2006. Cold stress tolerance of Listeria
monocytogenes: a review of molecular adaptive mecha-
nisms and food safety implications. J Food Prot. 69(6):
14731484.
Theriot JA, Mitchison TJ, Tilney LG, Portnoy DA. 1992. The
rate of actin-based motility of intracellular Listeria monocy-
togenes equals the rate of actin polymerization. Nature.
357(6375):257260.
Tilney L, Portnoy D. 1989. Actin filaments and the growth,
movement, and spread of the intracellular bacterial para-
site, Listeria monocytogenes. J Cell Biol. 109(4 Pt 1):
15971608.
Travier L, Guadagnini S, Gouin E, Dufour A, Chenal-
Francisque V, Cossart P, Olivo-Marin J-C, Ghigo J-M, Disson
O, Lecuit M. 2013. ActA promotes Listeria monocytogenes
aggregation, intestinal colonization and carriage. PLoS
Pathog. 9(1):e1003131.
Travier L, Lecuit M. 2014. Listeria monocytogenes ActA: a new
function for a classicvirulence factor. Curr Opin
Microbiol. 17(1):5360.
Tu Z, Chen Q, Li Y, Xiong Y, Xu Y, Hu N, Tao Y. 2016.
Identification and characterization of species-specific nano-
bodies for the detection of Listeria monocytogenes in milk.
Anal Biochem. 493:17.
Tully E, Higson SP, OKennedy R. 2008. The development of a
labelessimmunosensor for the detection of Listeria
monocytogenes cell surface protein, Internalin B. Biosens
Bioelectron. 23(6):906912.
Uusitalo S, K
ogler M, V
alimaa AL, Popov A, Ryabchikov Y,
Kontturi V, Siitonen S, Pet
aj
a J, Virtanen T, Laitinen R,
et al. 2016. Detection of Listeria innocua on roll-to-roll pro-
duced SERS substrates with gold nanoparticles. RSC Adv.
6(67):6298162989.
V
alimaa AL, Tilsala-Timisj
arvi A, Virtanen E. 2015. Rapid
detection and identification methods for Listeria monocy-
togenes in the food chain A review. Food Control. 55:
103114.
Wang G, Zhao H, Zheng B, Li D, Yuan Y, Han Q, Tian Z,
Zhang J. 2019. TLR2 promotes monocyte/macrophage
recruitment Into the liver and microabscess formation to
CRITICAL REVIEWS IN MICROBIOLOGY 19
limit the spread of Listeria monocytogenes. Front Immunol.
10(June):112.
Wang J, Ray AJ, Hammons SR, Oliver HF. 2015. Persistent and
transient Listeria monocytogenes strains from retail deli
environments vary in their ability to adhere and form bio-
films and rarely have inlA premature stop codons.
Foodborne Pathog Dis. 12(2):151158.
Wang W, Liu L, Song S, Xu L, Kuang H, Zhu J, Xu C. 2017.
Identification and quantification of eight Listeria monocyto-
genes serotypes from Listeria spp. using a gold nanopar-
ticle-based lateral flow assay. Microchim Acta. 184(3):
715724.
Wang Y, Salazar JK. 2016. Culture-independent rapid detec-
tion methods for bacterial pathogens and toxins in food
matrices. Compr Rev Food Sci Food Saf. 15(1):183205.
Wang W, Liu L, Song S, Xu L, Zhu J, Kuang H. 2017. Gold
nanoparticle-based paper sensor for multiple detection of
12 Listeria spp. by P60-mediated monoclonal antibody.
Food Ag Immunol. 28(2):274287.
Weller D, Andrus A, Wiedmann M, den Bakker HC. 2015.
Listeria booriae sp. nov. and Listeria newyorkensis sp. nov.,
from food processing environments in the USA. Int J Syst
Evol Microbiol. 65(Pt 1):286292.
Wieckowska-Szakiel M, Bubert A, Rozalski M, Krajewska U,
Rudnicka W, Rozalska B. 2002. Colony-blot assay with anti-
p60 antibodies as a method for quick identification of
Listeria in food. Int J Food Microbiol. 72(12):6371.
Wolfe B. 2017. Acute fetal demise with first trimester mater-
nal infection resulting from Listeria monocytogenes in a
nonhuman primate model. mBio. 8(1):e01938-16.
Wu Z. 2019. Simultaneous detection of Listeria monocyto-
genes and Salmonella Typhimurium by a SERS-based lat-
eral flow immunochromatographic assay. Food Anal
Methods. 12(5):10861091.
Xu L, Bai X, Bhunia AK. 2021. Current state of biosensors
development and their application in foodborne pathogen
detection. J Food Prot. DOI:10.4315/JFP-20-464.
Yeung PS, Zagorski N, Marquis H. 2005. The metalloprotease
of Listeria monocytogenes controls cell wall translocation
of the broad-range phospholipase C. J Bacteriol. 187(8):
26012608.
Yu K-Y, Noh Y, Chung M, Park H-J, Lee N, Youn M, Jung BY,
Youn B-S. 2004. Use of monoclonal antibodies that recog-
nize p60 for identification of Listeria monocytogenes. Clin
Diagn Lab Immunol. 11(3):446451.
Zhang CXY, Brooks BW, Huang H, Pagotto F, Lin M. 2016.
Identification of surface protein biomarkers of Listeria
monocytogenes via bioinformatics and antibody-based pro-
tein detection tools. Appl Environ Microbiol. 82(17):
54655476.
Zhang T, Abel S, Abel Zur Wiesch P, Sasabe J, Davis BM,
Higgins DE, Waldor MK. 2017. Deciphering the landscape
of host barriers to Listeria monocytogenes infection. Proc
Natl Acad Sci U S A. 114(24):63346339.
Zhang X, Tsuji S, Kitaoka H, Kobayashi H, Tamai M, Honjoh
KI, Miyamoto T. 2017. Simultaneous detection of
Escherichia coli O157:H7, Salmonella Enteritidis, and Listeria
monocytogenes at a very low level using simultaneous
enrichment broth and multichannel SPR biosensor. J Food
Sci. 82(10):23572363.
Zhao Y, Zhu A, Tang J, Tang C, Chen J. 2017. Identification
and measurement of staphylococcal enterotoxin M from
Staphylococcus aureus isolate associated with staphylococ-
cal food poisoning. Lett Appl Microbiol. 65(1):2734.
Zheng L, Li M, Li F, Zhang M. 2018. Paper-based chemilumin-
escence enzyme-linked immunosorbent assay enhanced
by biotin-streptavidin system for high-sensitivity C-reactive
protein detection. Anal Biochem. 559:8690.
20 L. LOPES-LUZ ET AL.
... To prevent further outbreaks of listeriosis, a limit of detection that is closer to the regulatory minimum infectious dose is required. 23 Hence, rapid detection methods need to be sensitive enough to immediately detect pathogens in small numbers present in different kinds of foods as well as in biological samples. In a previous study, we have developed a colorimetric probe for detecting L. monocytogenes using a monoclonal L. monocytogenes antibody (mAb-Lis) attached to amino-terminated oligo (ethylene glycol)-capped gold nanoparticles (NH 2 -TEG-AuNPs) in a covalently fixed "end-on" orientation. ...
... According to the LOD, the fixed "end-on" Lis-mAb-NH-TEG-AuNPs LFS offers enhanced sensitivity, reaching levels close to the minimum infectious dose of less than 100 CFU/mL. 1,23 As a result, our proposed LFS shows great potential as a POCT for detecting L. monocytogenes in readyto-eat raw food, aligning with the regulatory standards proposed by health surveillance agencies. 8,27 The cutoff value for color intensity was calculated according to the previous report 28,29 (mean of blank + 3SD of blank) and found to be 316.17 ...
... Especially, it exhibits a strong sensitivity in the detection of L. monocytogenes in human blood samples, which is close to the accepted minimum infectious dose. 1,23 In addition, the presence of three distinct LOD in these three samples suggests that the composition of the matrix in each sample could influence the colorimetric LFS. 41−43 Particularly in the milk sample, the LOD is higher compared to human blood and mushroom. ...
Article
Full-text available
In this study, the covalently fixed “end-on” orientation of a monoclonal Listeria monocytogenes antibody (mAb-Lis) to amino terminated oligo (ethylene glycol)-capped gold nanoparticles (NH2-TEG-AuNPs) was used to fabricate an in-house lateral flow strip (LFS), namely, the fixed “end-on” Lis-mAb-NH-TEG-AuNPs LFS. The aim was to evaluate the performance of the fixed “end-on” Lis-mAb-NH-TEG-AuNPs LFS in detecting L. monocytogenes. The proposed LFS enabled the sensitive detection of L. monocytogenes in 15 min with a visual limit of detection of 10² CFU/mL. Quantitative analysis indicated an LOD at 10 CFU/mL. The fixed “end-on” Lis-mAb-NH-TEG-AuNPs LFS showed no cross-reactivity with other pathogenic bacteria and practical performance across different food matrices, including human blood, milk, and mushroom samples. Furthermore, the clinical performance of the fixed “end-on” Lis-mAb-NH-TEG-AuNPs LFS for detecting L. monocytogenes was evaluated by using 12 clinical samples validated by the hemoculture method. It demonstrated excellent concordance with the reference methods, with no false-positive or false-negative results observed. Therefore, the fixed “end-on” Lis-mAb-NH-TEG-AuNPs LFS serves as a promising candidate for a point-of-care test (POCT), enabling the rapid, precise, and highly sensitive detection of L. monocytogenes in clinical samples and contaminated food.
... Listeriosis is characterized as a zoonotic disease resulting from the ingestion of contaminated food by L. monocytogenes. Systemic dissemination of pathogens from the gastrointestinal tract depends on their ability to overcome barriers such as the intestinal, blood-brain, and placental barriers [1,9]. Listeriosis is characterized by septisis and central nervous system infections, occurring primarily in immunocompromised hosts, the elderly, and pregnant women, as well as localized infections anatomically rare. ...
... Listeriosis is characterized by septisis and central nervous system infections, occurring primarily in immunocompromised hosts, the elderly, and pregnant women, as well as localized infections anatomically rare. Gastroenteritis is caused by healthy individuals when the ingested contaminated ready-to-eat foods such as hotdogs, cheeses (unpasteurised milk), smoked fish, ice cream, patés, cantaloupe, apple, and vegetables [9,10]. Although morbidity is very low in the normal population, these epidemics are characterized by high hospitalization and mortality rates, especially in high-risk groups with hospitalization rates higher than 95% in these cases [1,10]. ...
... Listeria species, more specific L. monocytogenes is a ubiquitous bacterium ( Figure 1) known for its adaptability, including antibiotic resistance genes and biofilm formation [2,10]. Its resistance to adverse environmental conditions such as high salt concentration, temperature range low pH and oxygen-limiting conditions, allows it to spread through food and multiply on various surfaces [9]. ...
Article
Full-text available
Antibiotics play an important role in veterinary medicine and serve as important tools to maintain animal health and ensure food safety. However, heavy use of antibiotics in animal production can lead to increased antimicrobial resistance from livestock to humans. Foodborne pathogens are a major public health and food safety problem. Listeria monocytogenes cause severe diseases and outbreaks associated to the consumption of contaminated food products, in humans. In the treatment of infections, L. mo-nocytogenes are susceptible to several antimicrobial agents, however , several recent studies have already reported cases of strains resistant to several classes of antibiotics, such as ampicillin, cefota-xime, tetracyclines, sulfonamides, β-lactams, and penicillin among livestock animals, but also the emergence of multi-resistant strains in these environments have also been described in several recent studies. This review focuses on the occurrence and prevalence L. monocytogenes in livestock and derived food-products and strives to provide information on prevalence of L. monocytogenes in livestock animals, and derived food products, and describe the main antimicrobial resistance and genomic analysis in strains associated and isolated from regions worldwide.
... Other groups of key listerial virulence factors, outside of LIPI-1, include internalins. The most important ones are inlA and inlB, but many have been identified (including inlC, inlJ, inlH, inlK, inlL, inlF and inlP) [8,9,12,13]. Other known listerial virulence factors include, for example, invasion-associated protein (encoded by iap), flagellin (encoded by flaA), a general stress-response regulator, called sigma factor B, (encoded by sigB), Listeriamucin-binding invasin A, bile salt hydrolase, and cell invasion LPXTG protein, ClpP, a heat shock protein that is involved in intracellular growth or fibronectin-binding protein [7,8,10,12], to just name a few. ...
... The most important ones are inlA and inlB, but many have been identified (including inlC, inlJ, inlH, inlK, inlL, inlF and inlP) [8,9,12,13]. Other known listerial virulence factors include, for example, invasion-associated protein (encoded by iap), flagellin (encoded by flaA), a general stress-response regulator, called sigma factor B, (encoded by sigB), Listeriamucin-binding invasin A, bile salt hydrolase, and cell invasion LPXTG protein, ClpP, a heat shock protein that is involved in intracellular growth or fibronectin-binding protein [7,8,10,12], to just name a few. However, not all L. monocytogenes isolates harbor all discovered virulence genes [14][15][16]. ...
Article
Full-text available
Listeria monocytogenes is a human pathogen that has the ability to cause listeriosis, a disease with possible fatal outcomes. The typical route of infection is ingestion of the bacteria with contaminated food. In this study, 13 virulence-associated genes were examined with PCR in the genomes of 153 L. monocytogenes isolates collected from meat products and processing environments in Poland. All isolates possessed genes from LIPI-1—hly, actA, plcA, plcB and mpl—as well as four internalins: inlA, inlB, inlC, inlJ. Invasion-associated protein iap, as well as genes prfA and sigB, encoding regulatory proteins, were also detected in all isolates. Gene flaA, encoding flagellin, was detected in 113 (74%) isolates. This was the only gene that was not detected in all isolates, as its presence is serotype-dependent. Gene actA showed polymorphism with longer and shorter variants in PCR amplicons. Two isolates were characterized by truncated inlB genes, lacking 141 bp in their sequence, which was confirmed by gene sequencing. All isolates were positive in hemolysis assays, proving the synthesis of functional PrfA and Hly proteins. Four genotypes of L. monocytogenes based on actA polymorphism and two genotypes based on inlB polymorphism were distinguished within the isolates’ collection.
... This Gram-positive facultative intracellular bacterium belongs to the genus Listeria, which comprises a total of 18 species with only two pathogenic species, i.e., L. monocytogenes and Listeria ivanovii [7]. The fundamental spread of L. monocytogenes from the gastrointestinal tract relies on its capacity to cross digestive, blood-brain, and placental boundaries and its infection is mediated by numerous virulence factors [8]. Diverse Listeria determinants are well-known to play a significant role in L. monocytogenes' pathogenicity. ...
... Diverse Listeria determinants are well-known to play a significant role in L. monocytogenes' pathogenicity. These virulence potentials are also influenced by temperature, the presence or absence of oxygen, osmotic stress, and pH [8]. ...
Article
Full-text available
Listeria monocytogenes, a foodborne pathogen causing listeriosis, poses substantial societal, economic, and public health challenges due to its resistance, persistence, and biofilm formation in the food industry. Exploring subinhibitory concentrations of compounds to target virulence inhibition and increase susceptibility to adverse conditions presents a promising strategy to mitigate its impact of L. monocytogenes and unveils new potential applications. Thus, this study aims to explore the effect of linalool on virulence factors of L. monocytogenes and potential use in the reduction in its tolerance to stressful conditions. This action was analysed considering the use of two sub-inhibitory concentrations of linalool, 0.312 and 0.625 mg/mL. We found that even with the lowest tested concentrations, a 65% inhibition of violacein production by Chromobacterium violaceum, 55% inhibition in biofilm formation by L. monocytogenes and 62% reduction on haemolysis caused by this bacterium were observed. In addition to its impact on virulence factors, linalool diminished the tolerance to osmotic stress (up to 4.3 log reduction after 24 h with 12% NaCl), as well as to high (up to 3.8 log reduction after 15 min at 55 °C) and low temperatures (up to 4.6 log reduction after 84 days with 12% NaCl at 4 °C). Thus, this study paves the way to further investigation into the potential utilization of linalool to mitigate the threat posed by L. monocytogenes in the field of food safety and public health.
Article
Full-text available
The Amplified Luminescent Proximity Homogenous Assay-linked Immunosorbent Assay (AlphaLISA) is known for detecting various protein targets; however, its ability to detect nucleic acid sequences is not well established. Here, the capabilities of the AlphaLISA technology were expanded to include direct detection of DNA (aka: oligo-Alpha) and was applied to the detection of Listeria monocytogenes. Parameters were defined that allowed the newly developed oligo-Alpha to differentiate L. monocytogenes from other Listeria species through the use of only a single nucleotide polymorphism within the 16S rDNA region. Investigations into the applicability of this assay with different matrices demonstrated its utility in both milk and juice. One remarkable feature of the oligo-Alpha is that greater sensitivity could be achieved through the use of multiple acceptor oligos compared to only a single acceptor oligo, even when only a single donor oligo was employed. Additional acceptor oligos were easily incorporated into the assay and a tenfold change in the detection limit was readily achieved, with detection limits of 250 attomole of target being recorded. In summary, replacement of antibodies with oligonucleotides allows us to take advantage of genotypic difference(s), which both expands its repertoire of biological markers and furthers its use as a diagnostic tool.
Article
Full-text available
High-pressure processing (HPP) is currently one of the leading methods of non-thermal food preservation as an alternative to traditional methods based on thermal processing. The application of HPP involves the simultaneous action of a combination of several factors—pressure values (100–600 MPa), time of operation (a few–several minutes), and temperature of operation (room temperature or lower)—using a liquid medium responsible for pressure transfer. The combination of these three factors results in the inactivation of microorganisms, thus extending food shelf life and improving the food’s microbiological safety. HPP can provide high value for the sensory and quality characteristics of products and reduce the population of pathogenic microorganisms such as L. monocytogenes to the required safety level. Nevertheless, the technology is not without impact on the cellular response of pathogens. L. monocytogenes cells surviving the HPP treatment may have multiple damages, which may impact the activation of mechanisms involved in the repair of cellular damage, increased virulence, or antibiotic resistance, as well as an increased expression of genes encoding pathogenicity and antibiotic resistance. This review has demonstrated that HPP is a technology that can reduce L. monocytogenes cells to below detection levels, thus indicating the potential to provide the desired level of safety. However, problems have been noted related to the possibilities of cell recovery during storage and changes in virulence and antibiotic resistance due to the activation of gene expression mechanisms, and the lack of a sufficient number of studies explaining these changes has been reported.
Article
Full-text available
Foodborne disease outbreaks continue to be a major public health and food safety concern. Testing products promptly can protect consumers from foodborne diseases by ensuring the safety of food before retail distribution. Fast, sensitive, and accurate detection tools are in great demand. Therefore, various approaches have been explored recently to find a more effective way to incorporate antibodies, oligonucleotides, phages, and mammalian cells as signal transducers and analyte recognition probes on biosensor platforms. The ultimate goal is to achieve high specificity and low detection limits (1 to 100 bacterial cells or piconanogram concentrations of toxins). Advancements in mammalian cell–based and bacteriophage-based sensors have produced sensors that detect low levels of pathogens and differentiate live from dead cells. Combinations of biotechnology platforms have increased the practical utility and application of biosensors for detection of foodborne pathogens. However, further rigorous testing of biosensors with complex food matrices is needed to ensure the utility of these sensors for point-of-care needs and outbreak investigations. HIGHLIGHTS
Article
Full-text available
Environmental cues promote microbial biofilm formation and physiological and genetic heterogeneity. In food production facilities, biofilms produced by pathogens are a major source for food contamination; however, the pathogenesis of biofilm-isolated sessile cells is not well understood. We investigated the pathogenesis of sessile Listeria monocytogenes ( Lm ) using cell culture and mouse models. Lm sessile cells express reduced levels of the lap , inlA, hly , prfA , and sigB and show reduced adhesion, invasion, translocation, and cytotoxicity in the cell culture model than the planktonic cells. Oral challenge of C57BL/6 mice with food, clinical, or murinized-InlA (InlA m ) strains reveals that at 12 and 24 h post-infection (hpi), Lm burdens are lower in tissues of mice infected with sessile cells than those infected with planktonic cells. However, these differences are negligible at 48 hpi. Besides, the expressions of inlA and lap mRNA in sessile Lm from intestinal content are about 6.0- and 280-fold higher than the sessle inoculum, respectively, suggesting sessile Lm can still upregulate virulence genes shortly after ingestion (12 h). Similarly, exposure to simulated gastric fluid (SGF, pH 3) and intestinal fluid (SIF, pH 7) for 13 h shows equal reduction in sessile and planktonic cell counts, but induces LAP and InlA expression and pathogenic phenotypes. Our data show that the virulence of biofilm-isolated Lm is temporarily attenuated and can be upregulated in mice during the early stage (12–24 hpi) but fully restored at a later stage (48 hpi) of infection. Our study further demonstrates that in vitro cell culture assay is unreliable; therefore, an animal model is essential for studying the pathogenesis of biofilm-isolated bacteria.
Article
Full-text available
Probiotic bacteria reduce the intestinal colonization of pathogens. Yet, their use in preventing fatal infection caused by foodborne Listeria monocytogenes (Lm), is inconsistent. Here, we bioengineered Lactobacillusprobiotics (BLP) to express the Listeria adhesion protein (LAP) from a non-pathogenic Listeria (L. innocua) and a pathogenic Listeria (Lm) on the surface of Lactobacillus casei. The BLP strains colonize the intestine, reduce Lm mucosal colonization and systemic dissemination, and protect mice from lethal infection. The BLP competitively excludes Lm by occupying the surface presented LAP receptor, heat shock protein 60 and ameliorates the Lm-induced intestinal barrier dysfunction by blocking the nuclear factor-κB and myosin light chain kinase-mediated redistribution of the major epithelial junctional proteins. Additionally, the BLP increases intestinal immunomodulatory functions by recruiting FOXP3⁺T cells, CD11c⁺ dendritic cells and natural killer cells. Engineering a probiotic strain with an adhesion protein from a non-pathogenic bacterium provides a new paradigm to exclude pathogens and amplify their inherent health benefits.
Article
Full-text available
Listeria monocytogenes is a Gram-positive facultative anaerobic bacterium that is responsible for the disease, listeriosis. It is particularly lethal in pregnant women, the fetus, the elderly and the immunocompromised. The pathogen survives and replicates over a wide range of temperatures (4 to 42 °C), pH, salt and oxygen concentrations. Because it can withstand various environments, L. monocytogenes is a major concern in food processing industries, especially in dairy products and ready-to-eat fruits, vegetables and deli meats. The environment in which the pathogen is exposed can influence the expression of virulence genes. For instance, studies have shown that variations in oxygen availability can impact resistance to stressors. Further investigation is needed to understand the essential genes required for the growth of L. monocytogenes in anaerobic conditions. Therefore, the purpose of this review is to highlight the data on L. monocytogenes under known environmental stresses in anaerobic environments and to focus on gaps in knowledge that may be advantageous to study in order to better understand the pathogenicity of the bacterium.
Article
Full-text available
Concerns about food contamination by Listeria monocytogenes are on the rise with increasing consumption of ready-to-eat foods. Biofilm production of L. monocytogenes is presumed to be one of the ways that confer its increased resistance and persistence in the food chain. In this study, a collection of isolates from foods and food processing environments (FPEs) representing persistent, prevalent, and rarely detected genotypes was evaluated for biofilm forming capacities including adhesion and sessile biomass production under diverse environmental conditions. The quantity of sessile biomass varied according to growth conditions, lineage, serotype as well as genotype but association of clonal complex (CC) 26 genotype with biofilm production was evidenced under cold temperature. In general, relative biofilm productivity of each strain varied inconsistently across growth conditions. Under our experimental conditions, there were no clear associations between biofilm formation efficiency and persistent or prevalent genotypes. Distinct extrinsic factors affected specific steps of biofilm formation. Sudden nutrient deprivation enhanced cellular adhesion while a prolonged nutrient deficiency impeded biofilm maturation. Salt addition increased biofilm production, moreover, nutrient limitation supplemented by salt significantly stimulated biofilm formation. Pan-genome-wide association study (Pan-GWAS) assessed genetic composition with regard to biofilm phenotypes for the first time. The number of reported genes differed depending on the growth conditions and the number of common genes was low. However, a broad overview of the ontology contents revealed similar patterns regardless of the conditions. Functional analysis showed that functions related to transformation/competence and surface proteins including Internalins were highly enriched.
Article
The discovery of the role of ActA to polymerise actin at one pole of Listeria monocytogenes represents a key event in the field of cellular microbiology. It uncovered much more than the molecular principle behind actin-based motility of Listeria within the cytosol of infected cells, and it changed the way how actin dynamics could be studied and eventually understood. The ActA discovery took place at a time when cell biology, biochemistry and microbiology came together in a very fruitful fashion. Here, we provide an overview of the science that took place around this event. Then, we outline the wide array of research fields that have been impacted by this finding. This ranges from structural and biophysical investigations on actin and its dynamics, the role of actin polymerisation during infection with different pathogens, to actin-dynamics during various pathologies. Like a comet in the sky, Pascale Cossart's work on ActA has inspired and will inspire generations of (life) scientists.
Chapter
Listeria monocytogenes is a Gram-positive motile facultative anaerobe that inhabits a broad ecologic niche (1-3). With selective media it can be readily isolated from soil, water, and vegetation, including raw produce designated for human consumption without further processing (4, 5). Newer chromogenic media may offer some advantages in the detection of contaminated foodstuffs (6, 7). Surface contamination of meat and vegetables is relatively common, with up to 15% of these foods harboring the organism. In addition, the organism is a transient inhabitant of both animal and human gastrointestinal tracts (8-10), and intermittent carriage suggests frequent exposure. The gut is the source for the organism in invasive listeriosis when it occurs, and the virulence factor ActA may promote carriage (11). The organism is psychrophilic and enjoys a competitive advantage against other Gram-positive and Gram-negative microorganisms in cold environments, such as refrigerators. It may also be amplified in spoiled food products, particularly when spoilage leads to increased alkalinity. Feeding of spoiled silage with a high pH has resulted in epidemics of listeriosis in sheep and cattle (12).
Article
Herein, we demonstrated the potential of surface-enhanced Raman scattering (SERS) spectroscopy combined with microfluidics for the detection and discrimination of foodborne pathogens. SERS-tagged gold nanostars (GNSs) were functionalized with a monoclonal antibody specific for Listeria monocytogenes. In the presence of L. monocytogenes, a SERS signal corresponding to the SERS tag paired to the antibody was detected in real time and in continuous flow, enabling the discrimination of L. monocytogenes and L. innocua in just 100 s. To the best of our knowledge, this is the first time that SERS tags have been used for the in-flow detection of living organisms.
Article
Aims: The aim of this study was to develop a rapid detection and differentiation method for pathogenic Listeria species in stone fruits. Methods and results: We utilized activated charcoal enrichment media (ACM) to induce overexpression and hypersecretion of pathogenic Listeria virulence proteins which can subsequently be detected via immunoblot analysis. Plum and nectarine slices spiked with either L. monocytogenes or L. ivanovii were incubated in pre-enrichment broth followed by enrichment in ACM. Secreted proteins were precipitated and subjected to SDS-PAGE and immunoblot analysis using a combination of L. monocytogenes specific antibody (α-listeriolysin O) and antibody specific for both L. monocytogenes and L. ivanovii (α-Internalin C). As few as 1 CFU g-1 of L. monocytogenes in plum and nectarine were detected whereas a detection limit of 10 CFU g-1 was achieved for L. ivanovii in each food tested following a 20 hour enrichment. Non-pathogenic Listeria species and non-Listeria bacterial pathogens tested were negative. Conclusions: These results demonstrate the highly sensitive and specific nature of the detection method for pathogenic Listeria in stone fruits using activated charcoal enrichment as well as the capability to discriminate between L. monocytogenes and L. ivanovii. Significance and impact of study: This method is the first to identify and differentiate L. monocytogenes and L. ivanovii in select stone fruit enrichments within 24 hours using immunological techniques. The rapidity and sensitivity of the method could aid in the reduction of exposure to the public in the event of an outbreak and expedite the administration of appropriate antibiotics to infected individuals.