DOI QR코드

DOI QR Code

Advances in Rapid Detection Methods for Foodborne Pathogens

  • Zhao, Xihong (Key Laboratory for Green Chemical Process of Ministry of Education, School of Chemical Engineering and Pharmacy, Wuhan Institute of Technology) ;
  • Lin, Chii-Wann (Institute of Biomedical Engineering, National Taiwan University) ;
  • Wang, Jun (Department of Food Science and Biotechnology and Institute of Bioscience and Biotechnology, Kangwon National University) ;
  • Oh, Deog Hwan (Department of Food Science and Biotechnology and Institute of Bioscience and Biotechnology, Kangwon National University)
  • Received : 2013.10.07
  • Accepted : 2013.12.22
  • Published : 2014.03.28

Abstract

Food safety is increasingly becoming an important public health issue, as foodborne diseases present a widespread and growing public health problem in both developed and developing countries. The rapid and precise monitoring and detection of foodborne pathogens are some of the most effective ways to control and prevent human foodborne infections. Traditional microbiological detection and identification methods for foodborne pathogens are well known to be time consuming and laborious as they are increasingly being perceived as insufficient to meet the demands of rapid food testing. Recently, various kinds of rapid detection, identification, and monitoring methods have been developed for foodborne pathogens, including nucleic-acid-based methods, immunological methods, and biosensor-based methods, etc. This article reviews the principles, characteristics, and applications of recent rapid detection methods for foodborne pathogens.

Keywords

Introduction

Foodborne pathogens are microorganisms (i.e., bacteria, viruses, and fungi) as well as a number of parasites, which are capable of infecting humans via contaminated food or water [24]. In particular, foodborne bacteria such as Escherichia coli O157:H7, Salmonella enterica, Staphylococcus aureus, Listeria monocytogenes, Campylobacter jejuni, Bacillus cereus, and other Shiga-toxin producing E. coli strains (non-O157 STEC), and Vibrio spp. are leading causes of foodborne diseases. In recent years, diseases caused by foodborne pathogens have become an important public health problem in the world, producing a significant rate of morbidity and mortality [72]. The global incidence of foodborne disease is difficult to estimate, but it has been reported that roughly 1 in 6 Americans in the United States (or 48 million people) gets sick, 128,000 are hospitalized, and 3,000 die of foodborne diseases annually according to CDC 2011 Estimates [9]. A great proportion of these cases can be attributed to the contamination of food and drinking water. Additionally, diarrhea is a major cause of malnutrition in infants and young children [110]. Although there are 31 pathogens that have been identified as causing foodborne illnesses, Norovirus, Salmonella, Campylobacter, Staphylococcus aureus, Listeria monocytogenes, Clostridium perfringens, Toxoplasma gondii, and Escherichia coli O157:H7 have been generally found to be responsible for the vast majority of illnesses, hospitalizations, and deaths [9, 100].

The high prevalence of foodborne diseases in many developing countries suggests major underlying food safety problems; therefore, it is important to detect foodborne pathogens in order to reduce foodborne disease occurrence. Traditional methods for the detection of bacterial pathogens from foods depend on culturing the organisms on agar plates; it is a time-consuming process, taking 2-3 days for initial results, and up to more than 1 week for confirming the specific pathogenic microorganisms. It is obvious that culture and colony counting methods are inadequate. In order to prevent the spread of infectious diseases, ensure the food safety, and thereby to protect public health, there is an ever-increasing demand for more rapid methods of foodborne pathogen detection.

There is a wide variety of microorganisms that are able to produce toxins causing foodborne diseases, mainly S. aureus, Vibrio cholerae, Clostridium botulinum, C. perfringens, Bacillus cereus, and E. coli O157 [27]. Existing literatures report numerous methods developed for the detection of toxin-related genes and their toxin products. The detection methods of toxin-related genes are nucleic-acid-based methods, such as molecular amplification and hybridization probing. The detection methods of toxin products rely primarily on immunological assays, such as ELISA, lateral flow immunoassay and agglutination tests, and bioassays such as mouse neutralization testing and cytotoxicity assays in tissue culture, as well as biosensor-based assay [27, 90].

Recently, many researchers are focusing on the progress of rapid methods for foodborne pathogens. Novel molecular techniques for pathogens are being developed on various aspects of detection, such as sensitivity, rapidity, and selectivity, discrimination of the viable cell, and also suitability for in situ analysis. The immunological methods permit the rapid and sensitive analysis of a range of pathogens and toxins, especially with potential for on-site analysis. The emerging biosensor methods can detect foodborne pathogens in a much shorter time with sensitivity and selectivity comparable to the conventional methods and can potentially be used in the future as stand-alone devices for on-site monitoring. According to the main principle, these rapid detection methods can be classified into the following categories: nucleic-acid-based methods, immunological methods, and biosensor-based methods. The purpose of this paper is to review such rapid methods of detection and identification and to discuss some of the more recent and novel methods for the characterization of foodborne pathogens.

 

Nucleic-Acid-Based Methods

One of the advantages of nucleic-acid-based food pathogen detection assays is the high level of specificity, as they detect specific nucleic acid sequences in the target organism by hybridizing them to a short synthetic oligonucleotide complementary to the specific nucleic acid sequence. Several different types of nucleic-acid-based assays, including amplification, hybridization, microarrays, and biochips, have been developed for use as rapid methods to detect foodborne pathogens [92].

Simple PCR Method

PCR is the most well-known and established nucleic acid amplification technique for detecting pathogenic microorganisms [19]. In this method, double-stranded DNA is denatured into single strands, and specific primers or single-stranded (ss) oligonucleotides anneal to these DNA strands, followed by extension of the primers complementary to the singlestranded DNA, with a thermostable DNA polymerase. These steps are repeated, resulting in doubling of the initial number of target sequences with each cycle. This quantity of the products of amplification can be visualized as a band on an ethidium-bromide-stained electrophoresis gel. Identification based on PCR amplification of target genes by sequencing is considered to be a reliable technique when properly developed and validated for a certain species. With the distinct advantages of rapidity, specificity, sensitivity, and less samples over culture-based methods, many PCR assays for the detection and validation of foodborne bacteria and viruses in food have been developed and applied in food samples [38].

PCR is also used for toxins detection by amplifying specific genes that encode bacterial toxins. PCR methods for toxin detection have been developed for a number of bacterial species, including V. cholera, B. cereus, E. coli, and S. aureus. In addition to PCR, a number of gene-specific hybridization probes have been designed and used for the detection of toxin genes in foodborne pathogens [77].

Multiplex PCR

Simultaneous amplification of more than one locus is required for a rapid detection of multiple microorganisms in a single reaction. It is a methodology referred to as multiplex PCR (mPCR), in which several specific primer sets are combined into a single PCR assay [10]. Apparently, the design of the primers is a key factor in the development of a multiplex PCR assay. There may be some interaction between the multiple primer sets, so the primer concentrations may have to be adjusted in order to generate reliable yields of all the PCR products. Meanwhile, the primer sets should be designed with a similar annealing temperature, while providing a method to distinguish between amplicons following thermal cycling. Today, mPCR can also be useful to define the structure of certain microbial communities and to evaluate community dynamics, such as during fermentation or in response to environmental variations. Kong et al. [46] described a rapid mPCR method allowing for the simultaneous detection of six commonly encountered waterborne pathogens in a single tube. The target genes used were the aerolysin (aero) gene of Aeromonas hydrophila, the invasion plasmid antigen H (ipaH) gene of Shigella flexneri, the attachment invasion locus (ail) gene of Yersinia enterocolitica, the invasion plasmid antigen B (ipaB) gene of Salmonella Typhimurium, the enterotoxin extracellular secretion protein (epsM) gene of Vibrio cholerae, and a species-specific region of the 16S-23S rDNA (Vpara) gene of Vibrio parahaemolyticus.

Park et al. [75] established a mPCR assay for the simultaneous detection of Escherichia coli O157:H7, Salmonella spp., Staphylococcus aureus, and Listeria monocytogenes, in one tube. The mPCR employed the Escherichia coli O157:H7 specific primer Stx2A, Salmonella spp. specific primer Its, S. aureus specific primer Cap8A-B, and Listeria monocytogenes specific primer Hly. Amplification with these primers produced products of 553, 312, 405, and 210 bp, respectively. Recently, Mukhopadhyay et al. [66] used fliCh7 and iap gene-specific primers to establish a multiplex-PCR assay for the simultaneous detection of Escherichia coli O157:H7 and Listeria monocytogenes. Meanwhile, they developed a modified method of enrichment and harvesting, leading to a highly sensitive and rapid single-reaction PCR detection of both pathogens.

Quantitative PCR

Quantitative PCR (qPCR), also called real-time PCR, is an approach capable of continuously monitoring the PCR product formation throughout the reaction; it offers rapid, simultaneous amplification and sequence-specific-based detection of target genes and is increasingly being applied in food microbiology [41]. Using this method allows quantifying one specific microorganism in food and studying its behavior as a consequence of the influence of the environment (i.e., food composition, temperature, pH, oxygen, etc.) by studying expression of suitable target genes. Moreover, the real-time monitoring of the process means no need for post-amplification treatment of the samples, such as gel electrophoresis, reducing the time of analysis. Gomez et al. [33] developed a qPCR to quantify the total aerobic bacteria and fungi on fresh produce, using as reference the centrifugation water (CW) that comes up during processing instead of the food matrix itself. On average, 35% of the natural bacterial population and 64% of inoculated bacteria were recovered in the CW. Enumeration of cell number by qPCR did not differ significantly from plate assay and therefore, may replace it. This method could be an alternative to plate assays in order to get reliable information about the aerobic bacterial load of fresh-cut commodities in less than 5 h.

Derzelle et al. [22] developed a multiplex qPCR assay capable of detecting all known stx gene variants, including the highly divergent subtype stx2f, and evaluated its performance in combination with two different internal amplification controls. The new screening method was tested with artificially and naturally contaminated food samples and compared with two stx-specific assays used routinely in their laboratory: a PCR-ELISA method and a real-time PCR system, which followed the recommendations from the International Organization for Standardization Technical Specification (ISO/TS) 13136 defining a method for the detection of the main pathogenic Shiga-toxin producing Escherichia coli (STEC) in foodstuffs. The results showed that the newly developed qPCR method performed equally as well as the PCR-ELISA test and the stx-IAC realtime PCR test when applied to the same 353 naturally contaminated test portions (99.7% concordance).

Fusco et al. [28] developed a TaqMan and a SYBR Green real time PCR assay for reliable identification and quantitative detection of S. aureus strains harboring the enterotoxin gene cluster, regardless of their variants. Using optimized qPCR conditions, the assay was able to quantitatively detect at least about 1 × 103 and 1 × 104 CFU of the pathogen per milliliter raw milk (10 and 100 CFU equivalents of egc+ S. aureus per reaction mixture) by the SYBR Green and TaqMan qPCR assay, respectively.

Recently developed qPCR assays eliminated the post-PCR step by means of real-time monitoring of the PCR product generation, and the multiplex PCR approach has been implemented in the qPCR using a set of TaqMan probes labeled with different fluorescent dyes. Kim et al. [45] developed and proposed a multiplex qPCR assay for the simultaneous detection of V. cholerae, V. parahaemolyticus, and V. vulnificus, using zot, vmrA, and vuuA as target genes, respectively. The overall procedure took approximately 12 h, including the enrichment culture period; it yielded a method that was faster, simpler, and less costly than conventional culture-based methods. Using enrichment culture with alkaline peptone water and optimized ultiplex qPCR assay, they achieved a practical maximum sensitivity (100 CFU/g food homogenate) for each target species in all food matrices tested. Therefore, the method was shown to achieve a maximum sensitivity that meets the FDA guidelines (104 CFU/g) for acceptable levels of V. cholerae, V. parahaemolyticus, and V. vulnificus in seafood.

Isothermal Amplification

Although PCR has been widely used in foodborne pathogens, it requires thermocycling to separate the double strands of DNA; this has limited its application in the lowresource settings. During the past two decades, many novel methods have been developed to amplify nucleic acids under isothermal conditions. These methods include loopmediated isothermal amplification (LAMP), nucleic acid sequence-based amplification (NASBA), rolling circle amplification (RCA), and strand displacement amplification (SDA). Isothermal amplification has simpler hardware requirements than PCR, as it does not require a thermal cycling system, and may even work with a simple water bath setup. Isothermal amplification techniques have better tolerance than PCR to some inhibitory materials that affect the molecular amplification efficiency [30].

Loop-mediated isothermal amplification. Most recently, a novel nucleic acid amplification method known as LAMP has been demonstrated as a rapid, low-cost, easy operating, highly sensitive, and specific detection method applied in several fields [70]. This method relies on an autocycling strand displacement DNA synthesis performed by the Bst DNA polymerase large fragment, which is different from PCR in that 4-6 primers are used to target 6-8 specific regions of the target gene. The amplification is performed under isothermal conditions between 59℃ and 65℃, and the amplicons are mixtures of many different sizes of stem loop DNAs with several inverted repeats of the target sequence and cauliflower-like structures with multiple loops. The reaction can be accelerated with additional one or two loop primers. LAMP reactions usually result in about 103-fold or higher levels of amplification product with stem-loop DNAs in 60 min than conventional PCR. LAMP products can be observed with the naked eye by employing SYBR Green I dye instead of conventional gel electrophoresis analysis; the color of the solution changes to green in the presence of LAMP amplicons, whereas it remains orange for mixtures with no amplification.

The first foodborne pathogen application of the LAMP method was for the detection of stxA2 in Escherichia coliO157: H7 cells [62]. The mild permeabilization conditions and low isothermal temperatures used in the in situ LAMP method caused less cell damage than in situ PCR. The results showed that higher-contrast images were obtained with this method than with in situ PCR. Chen et al. [13] developed and evaluated a LAMP assay for identification and direct detection of acidophilic thermophilic bacteria (ATB) contaminants in pure juices. The LAMP method could detect 2.25 × 101 CFU/ml of ATB in juice samples within 2 h.

Recently, derivative LAMP assays, such as reversetranscription LAMP assay [12], multiplex LAMP assay [40], in situ LAMP assay [38, 116], and real-time reversetranscription LAMP assay [55], have been developed and employed for the detection of various foodborne pathogens, such as Bacillus anthracis [79], Vibrio parahaemolyticus [69, 113, 120], Staphylococcus aureus [114], Salmonella [13, 116, 119], Pseudomonas aeruginosa [121], Escherichia coli O157 [71, 118], and Listeria monocytogenes [105].

Nucleic acid sequence-based amplification. NASBA is an isothermal amplification reaction for the detection of RNA or DNA, which was developed after PCR had begun gaining widespread attention [18]. The reaction typically consists of three enzymes, including T7 RNA polymerase, RNase H, and avian myeloblastosis virus (AMV) reverse transcriptase (RT), all of which act together to amplify sequences from an original single-stranded RNA template. The reaction also includes buffering agents and two specific primers and takes place at approximately 41℃ [35].

NASBA is specific for target RNA or DNA sequences and has been gaining popularity owing to its wide range of applications for pathogen detection in clinical, environmental, and food samples [4]. Simpkins et al. [93] showed that NASBA can selectively amplify mRNA sequences from Salmonella enterica in a background of genomic DNA and demonstrated that NASBA could be a great means of assessing cell viability. Min and Baeumner [65] developed a NASBA assay for the detection of viable Escherichia coli. Baeumner et al. [5] confirmed the assay’s specificity for viable Escherichia coli by demonstrating that heat-killed cells did not produce a signal above the background of the instrumentation. Churruca et al. [17] developed a NASBA assay based on molecular beacons used for real-time detection of Campylobacter jejuni and Campylobacter coli in samples of chicken meat.

Real-time NASBA was proven to be the basis of sensitive and specific assays for detection, quantification, and analysis of RNA (and, in one case, DNA) targets [99]. Molecular beacons were used to generate fluorescence signals with NASBA assays for the detection of Vibrio cholerae [29]. More examples of the application of nucleic-acid-based techniques in food and other samples are presented in Table 1.

Table 1.aEnrichment.

 

Immunological Methods

Immunological detection based on antigen-antibody bindings is widely used for determining foodborne pathogens. These assays rely mainly on the specific binding of an antibody to an antigen. A variety of antibodies have been employed in different assay types for the detection of foodborne pathogens and microbial toxins. The suitability of the antigen-antibody complex depends mainly on the antibodies’ specificity. In order to ensure the reliable detection of foodborne pathogens using antibody-based methods, the influence of stress on antibody reactions should be thoroughly examined and understood first, as the physiological activities in cells are often altered in response to a stress [36]. Most polyclonal antibodies, derived from either rabbit or goat serum, contain a collection of antibodies with different cellular origins and, therefore, somewhat different specificities. Monoclonal antibodies are often more useful than polyclonal antibodies for specific detection of a molecule, since they provide an indefinite supply of a single antibody. With the development of monoclonal antibodies, immunological detection of microbial contamination has become more specific, sensitive, reproducible, and reliable, as many commercial immunological assays are available for the detection of a wide variety of microbes and their products [50].

Enzyme-Linked Immunosorbent Assay

One of the most widely used immunological assays for foodborne pathogens detection is enzyme-linked immunosorbent assay (ELISA), which is a very accurate and sensitive method for detecting antigens or haptens [101]. Traditional ELISA typically involves chromogenic reporters and substrates that produce some kind of observable color change to indicate the presence of antigen or analyte. The most powerful ELISA format is called the “sandwich” assay, because the antigen from the enrichment cultures to be measured is bound between two primary antibodies: the capture antibody and the detection antibody. The sandwich format is used because it is sensitive and robust. The walls of wells in microtiter plates are the most commonly used solid support; however, ELISAs have also been designed using dipsticks, paddles, membranes, pipet tips, and other solid matrices [26]. Bolton et al. [6] described the BIOLINE Salmonella ELISA test for Salmonella spp., which was a rapid, easy, and convenient assay for the detection of Salmonella in foods and feeds. The limit of detection of the ELISA test kit was as low as 1 CFU/25 g sample with at least 4 of the 20 matrixes tested, and was found to be applicable to all sample types tested. The BIOLINE Salmonella ELISA test kit was granted AOAC-RI performance tested status.

Many foodborne toxins detection rely mainly on the presence of immunological reactions that are used to detect toxins. ELISA, the most commonly used in toxins detection, has been generated for staphylococcal enterotoxins A, B, C, and E and found to have detection levels of less than 0.5 μg/100 g in ground beef. ELISA has also been employed for the detection of botulinum toxins and enterotoxins produced by E. coli [27].

Lateral Flow Immunoassay

Although ELISA has been widely used in many laboratories, this method still requires various equipments and trained personnel. Therefore, rapid and cheap, yet still reliable methods that can be conducted and interpreted at the site of the contamination are needed. More and more on-site immunological techniques based on lateral flow immunoassays such as dipstick, immunochromatography, and immunofiltration are gaining attention in the area of pathogen, mycotoxin, and disease detection in the food industry and medicine [67].

Lateral flow assays are a form of immunoassay in which the test sample flows along the solid substrate via capillary action. After the sample is applied to the test, it encounters a colored reagent (antibody or antigen labeled by colloidal latex or gold particles), which mixes with the sample and transits the substrate, encountering lines or zones that have been pretreated with an antibody or antigen. Depending on the analytes present in the sample, the colored reagent can become bound at the test line or zone [32]. Most lateral flow assays are basically designed to incorporate a visual response about 2-10 min after the application of the sample. Using these techniques allows simplifying the detection and minimizing the manipulations in order to provide accurate results with little or no instrumentation [25].

Delmulle et al. [21] developed an immunoassay-based lateral flow dipstick for the rapid detection of aflatoxin B1 in pig feed. The visual detection limit for aflatoxin B1 was 5 μg/kg. Jung et al. [42] developed a colloidal immunochromatographic strip for the detection of Escherichia coli O157:H7 in enriched samples, reporting that the minimum limit was 1.8 × 105 CFU/ml without enrichment and 1.8 CFU/ml after enrichment. Avidin and streptavidin are widely used in (strept)avidin-biotin system technology, which is based on their tight biotin-binding capability. These biotin-(strept)avidin-based methods enable both a signal amplification and a reduction in background activity, resulting in suitable analytical techniques to be used in many fields [84].

The labels used in lateral flow immunoassay are mainly colloidal gold and monodisperse latex, labeled with colored, fluorescent, or magnetic tags. Dyed latexes and paramagnetic particles are available from a variety of sources, including Bangs Laboratories, Dynal, Merck/Estapor, and Magsphere. Magnetic, latex, metal, and semiconductor particles on the nanometer scale have unique optical, electronic, and structural properties that can be used in a variety of detection applications [78].

With the development of semiqualitative and qualitative assays [2] as well as autoreading technologies, such as the Biosite Triage and Response Biomedical RAMP systems, Magnetic Assay Reader (MAR), Cozart’s DDS, or Rapiscan products such as American BioMedica Corporation’s Rapid Reader for their Rapid Screen, lateral flow immunoassay will be applied in ways that have the potential to create entirely new paradigms in high-sensitivity point-of-need testing on-site.

Immunomagnetic Separation Assay

Immunomagnetic separation (IMS), a procedure that utilizes immunomagnetic beads (IMBs) as capturing reagents, has been developed for microbial isolation and identification. IMS is analogous to selective cultural enrichment, whereby the growth of other pathogen is suppressed while the target pathogen is allowed to grow. The separation process consists of two fundamental steps; first, the target cells are mixed with immunomagnetic particles for incubation of less than 1 h and separated by an appropriate magnetic separator; then, the magnetic complex is washed several times to remove the contaminants [60]. The use of IMS in assays is increasing because magnetic handling is fast, efficient, and only slightly affects the target analytes. Furthermore, various bioreactive molecules can be conjugated to the IMB surface for the immunoprecipitation, isolation, and identification of biomolecules (such as cells, pathogens, and proteins), or to improve the resolution of magnetic resonance imaging (MRI) [97]. Unlike the several days necessary to perform the conventional microbiological method and additional workup to elucidate the microbial status of any suspected colonies (more than 500 CFU in plate), DeCory et al. [20] developed and optimized a protocol for the rapid detection of Escherichia coli O157:H7 in aqueous samples by a combined immunomagnetic beadimmunoliposome (IMB/IL) fluorescence assay within a single 8 h work shift. The assay was able to identify samples containing Escherichia coli O157:H7 with 100% accuracy. The results highlighted the possible benefits of using immunomagnetic beads in combination with sulforhodamine B-encapsulating immunoliposomes for the rapid detection of Escherichia coli O157:H7 in aqueous samples.

The growing importance of mass spectrometry for the identification and characterization of protein toxins produced by foodborne pathogens is a result of the improved sensitivity and specificity of mass-spectrometrybased techniques, especially when these techniques are combined with affinity methods. Schlosser et al. [87] reported a novel method based on the use of immunoaffinity capture and matrix-assisted laser desorption ionization–time-of-flight mass spectrometry for selective purification and detection of staphylococcal enterotoxin B (SEB). In this method, an affinity molecular probe was prepared by immobilizing the anti-SEB antibody on the surface of paratoluene-sulfonyl-functionalized monodisperse magnetic particles and was used to selectively isolate SEB. Immobilization and affinity capture procedures were optimized to maximize the density of anti-SEB immunoglobulin G and the amount of captured SEB, respectively, on the surface of the magnetic beads. SEB could be detected directly “on beads” by placing the molecular probe on the matrix-assisted laser desorption ionization target plate or, alternatively, “off beads” after its acidic elution.

Other technologies relying on the Ab-Ag binding mechanism have also been developed and applied in detection of foodborne pathogens and toxins. Lee and Deininger [49] used IMS and ATP bioluminescence for the selective capture of target bacteria and their quantification, respectively. The method consisted of trapping bacteria on a filter, resuspending them in a small amount of buffer, and washing the suspension with an antibody-coated magnetic bead mixture specific to the bacterial species of interest. A detection limit of about 20 CFU/100 ml was achieved, which was well below the action limits of 300 CFU/ml (daily event), or a 30 day moving average of 126 CFU/100 ml set by US EPA. The entire procedure took less than 1 h to perform without an enrichment step. The study demonstrated that the system combining IMS with ATP bioluminescence was effective and expedient for detecting Escherichia coli in beach water.

 

Biosensor-Based Methods

Biosensors have recently been defined as analytical devices incorporating a biological material (e.g., tissue, microorganisms, organelles, cell receptors, enzymes, antibodies, nucleic acids, natural products, etc.), a biologically derived material (e.g., recombinant antibodies, engineered proteins, aptamers, etc.), or a biomimic (e.g., synthetic catalysts, combinatorial ligands, and imprinted polymers) intimately associated with or integrated within a physicochemical transducer or transducing microsystem, which may be optical, electrochemical, thermometric, piezoelectric, magnetic, or micromechanical [48]. Biosensors are devices for pathogen detection and generally consist of at least three elements, including a biological capture molecule (e.g., probes or antibodies), a method of converting capture molecule-target interactions into a signal, and a data output system [48, 100]. The greatest advantageous aspects of biosensors are those that enable fast or real-time detection, portability, and multi-pathogen detection for both field and laboratory analyses. The advantages of fast or real-time detection can provide almost immediate interactive information on the food materials, which enable users to take corrective measures before consumption or further contamination occurs [82]. It has been reported that biosensors have been developed and applied to the microbial analysis of foodborne pathogens, including Escherichia coli O157:H7, Staphylococcus aureus, Salmonella, and Listeria monocytogenes, as well as various microbial toxins such as staphylococcal enterotoxins and mycotoxins [3]. Different modes of biosensor-based foodborne pathogen detection are given in Table 2.

Table 2.Different modes of biosensor-based foodborne pathogen detection.

Optical Biosensors

Optical biosensors are a powerful alternative to conventional analytical techniques, due to their particularly high specification and sensitivity, as well as their small size and cost-effectiveness [58]. Biosensor detection typically relies on an enzyme system, which catalytically converts analytes into products that can be oxidized or reduced at a working electrode and maintained at a specific potential. One of the best advantages of this optical transducer is the low cost and the use of biodegradable electrodes. An optical biosensor is a compact analytical device containing a biological sensing element integrated or connected to an optical transducer system [23]. Optical biosensor technology can be classified into several subclasses based on absorption, reflection, refraction, Raman, infrared, chemiluminescence, dispersion, fluorescence, and phosphorescence. In the past decade, various kinds of optical biosensors for the rapid detection of pathogens, toxins, and contaminants in the food industry have been developed [100]. The main advantage of this technique is the real-time binding reaction detection, allowing kinetic evaluation of affinity interactions and, in addition, the low cost of the instrumentation required. Optical biosensors require a suitable spectrometer to record the spectral chemical properties of the analyte. A common method that employs the techniques of optical detection using reflectance spectroscopy for detection of foodborne pathogens is surface plasmon resonance (SPR).

SPR is a collective oscillation of free charges (conduction electrons) present at the interface of two media (metal-dielectric) with permittivities of opposite sign [1]. Receptors or antibodies initially immobilized on the surface of a thin film of precious metal, deposited on the reflecting surface of an optically transparent waveguide, are used to capture the various target pathogens. The sensing surface is located above or below a high-index resonant layer and a lowindex coupling layer. When a visible or near-infrared radiation (IR) is passed through the waveguide in the correct manner, the interaction of light with the electron cloud in the metal generates a strong resonance. Binding of the pathogen to the metal surface causes a shift in resonance to longer wavelengths, and the corresponding amount of shift reflects the concentration of bound pathogens [3]. The main drawbacks of current SPR technique lay in its complexity (specialized staff is required), high cost of equipment, and the large size of most currently available instruments. For this reason, the miniature SPR instrument and disposable cartridge and biochip were developed for real-time genetic detection in a cost-effective manner. This system included a disposable SPRLAMP cartridge made of PMMA with a PC prism and a simple SPR imaging system with temperature control for LAMP amplification [16].

Wang et al. [107] developed a SPR immunosensor for the detection of Escherichia coli O157:H7 by means of a new subtractive inhibition assay. Their results showed that the signal was inversely correlated with the concentration of E. coli O157:H7 cells in a range from 3.0 × 104 to 3.0 × 108 CFU/ml, where the limit of detection was 3.0 × 104 CFU/ml. The limit of detection subtractive inhibition assay method was reduced by one order of magnitude, compared with direct SPR by immobilizing antibodies on the chip and ELISA for E. coli O157:H7 (limit of detection: both 3.0 × 105 CFU/ml).

Several commercial instruments using SPR techniques are available from companies such as BIAcore and Biosensing Instruments Inc. Commercial SPR instruments have a detection limit of 105 CFU/ml for Listeria monocytogenes. The commercially available low-cost SPREETA SPR biosensor was reported to detect an E. coli O157:H7 enterotoxin, stx1, with a detection limit of 300 pmol compared with a bulk acoustic wave sensor [96].

Piezoelectric Biosensors

Piezoelectric biosensors, which are capable of sensitive detection of minute amounts of analytes according to a linear relationship between the deposited mass and its frequency response, are an effective alternative to established label-free optical sensors, such as surface plasmon resonance spectroscopy and interferometry [11]. Piezoelectric biosensors have been widely used, and their performance for studies of affine interactions was extensively referred.

Olfactory sensing of specific volatile organic compounds released by the bacterial pathogens is one of the more outstanding ways to determine contamination in food products. Sankaran et al. [86] used a computational simulation to determine the biomimetic peptide-based sensing material to be deposited on the quartz crystal microbalance (QCM) sensor for detecting specific gases (alcohols) at low concentrations in food samples. The results showed that the developed QCM sensors were sensitive to 1-hexanol as well as 1-pentanol as predicted by the simulation algorithm. The estimated lower detection limits of the QCM sensors for detecting 1-hexanol and 1-pentanol were 2-3 ppm and 3-5 ppm, respectively. This report demonstrated the applicability of a simulation-based peptide sequence that mimics the olfactory receptor for sensing specific gases.

Salmain et al. [85] successfully designed a direct, labelfree immunosensor for the rapid detection and quantification of staphylococcal enterotoxin A (SEA) in buffered solutions, using the quartz crystal microbalance with dissipation (QCM-D) as a transduction method. With the optimized sensing layer, a standard curve for the direct assay of SEA was established from QCM-D responses within a working range of 50-1,000 or 2,000 ng/ml, with a detection limit of 20 ng/ml. The total time for analysis was 15 min. The study indicated that such systems had a considerable amount of potential for the rapid and reliable detection of targets at trace amounts of pathogens in various environments.

Immunosensors

Immunosensors, which are based on specific antibodyantigen interactions, detect antigen binding to antibodies by immobilizing the reaction to the surface of a transducer, which converts surface change parameters into a detectable electric signal [32]. It is difficult to measure immunological reactions in real time owing to the diffusion limitations of antigens to immobilized antibodies, particularly for low levels of contaminants. However, most immunosensors produce results within 20-90 min, which is close to real time compared with conventional techniques and classical ELISAs. Moreover, the results of immunosensors are read via digital signals and are not as dependant on personal factors such as bias, fatigue, level of training, or visual disorders. However, this property is also shared by microtiter plate spectrophotometric immunoassays [98]. Chen [15] reported a new conductometric immunebiosensor for the detection of staphylococcal enterotoxin B (SEB) based on immobilization of horseradish peroxidase (HRP)-labeled SEB antibody (HRP-anti-SEB) onto a nanogold/ chitosan-multiwalled carbon nanotube (Au/CTS-MWNT)-functionalized biorecognition interface. The results showed that under optimal conditions, the proposed immunebiosensor exhibited a good conductometric response relative to SEB concentration in a linear range from 0.5 to 83.5 ng/ml, with a correlation coefficient of 0.998.

Electrochemical Biosensors

Electrochemical-based detection methods are further transduction-based systems that have been used for identifying and quantifying foodborne pathogens. Electrochemical biosensors can be classified into amperometric, potentiometric, impedimetric, and conductometric responses, based on observed parameters such as current, potential, impedance, and conductance, respectively [100]. Electrochemical biosensors developed for the simultaneous multiplexed analysis of foodborne pathogens primarily use electrochemical impedance spectroscopy as the transduction technique, thus providing label-free, on-line, high-throughput devices for bacterial detection [76]. Impedance spectroscopy is a powerful method for the study of conducting materials and interfaces. Through this technique, a cyclic function of small amplitude and variable frequency is applied to a transducer, and the resulting current is used to calculate the impedance at each of the probed frequencies [48]. Impedance biosensors for the detection of foodborne pathogens are based on the measurement of changes in the electrical properties of bacterial cells when they are attached to, or associated with, the electrodes [115]. Furthermore, the advantages in microfabrication techniques have enabled the use of microfabricated microarray electrodes for impedance detection, and the miniaturization of impedance microbiology into a chip assay.

Louie et al. [57] developed an impedance-based, fieldable biosensor system to detect Escherichia coli O157:H7 and Salmonella spp. The portable biosensor system used a variety of disposable analyte-specific sensor modules, each of which could be used to quantitatively determine specific analytes. The response for each sensor was rapid, and stable readings could be obtained in less than 1 min. Despite the development of a portable reagentless impedance biosensor that allows rapid detection of specific foodborne pathogens, however, no real foodborne or clinical sample application was considered.

There seems to be a lot of interest in the development of integrated biosensors for the detection of multiple biologically relevant species. A miniaturized biosensor device composed of a probe, sampler, detector, amplifier, and logic circuitry for monitoring infectious pathogens is an attractive alternative to existing instrumentation. Normal biosensors and biochips employ only one type of bioreceptor as probes (i.e., either nucleic acid, enzyme, or antibody probes). The multifunctional biochip (MFB) is an integrated multi-array biochip, designed by combining integrated circuit elements, an electro-optics excitation/detection system, and bioreceptor probes into a self-contained and integrated microdevice [103]. The MFB is a superior system that can detect multiple specific analytes simultaneously and offer information on both gene mutation (with DNA probes) and protein expression (with antibody probes) simultaneously. Vo-Dinh et al. [103] described a MFB, which used two different types of bioreceptors, including nucleic acid and antibody probes, on a single platform. The multifunctional capability of the MFB device for biomedical diagnostics was illustrated by the measurements of DNA probes specific to gene fragments of Bacillus anthracis and antibody probes targeted to Escherichia coli. The results showed that the calibration curves for monitoring pathogenic species illustrated the capability of the device for medical diagnostics and for quantitative detection of pathogenic agents.

Cao et al. [8] described a rapid and sensitive DNA target detection using enzyme amplified electrochemical detection based on a microchip. They employed a biotin-modified DNA, which reacted with avidin-conjugated horseradish peroxidase (avidin-HRP), in order to obtain the HRPlabeled DNA probe, and hybridized it with its complementary target. After hybridization, the mixture containing dsDNAHRP, excess ssDNA-HRP, and remaining avidin-HRP was separated. With this protocol, the limits of quantification for the hybridization assay of 21- and 39-mer DNA fragments were 8 × 10-12 M and 1.2 × 10-11 M, respectively. The method was applied satisfactorily in the analysis of Escherichia coli genomic DNA.

Wang et al. [104] described a new diagnostic assay for the rapid detection of methicillin-resistant Staphylococcus aureus (MRSA) by combining nucleic acid extraction and isothermal amplification of target nucleic acids in a magnetic beadbased microfluidic system. LAMP amplification of the target genes was performed via the incorporation of a built-in micro temperature control module, followed by spectrophotometric analysis of the optical density of the LAMP amplicons. The results showed that the limit of detection for MRSA in clinical samples was approximately 10 fg/ml by performing this diagnostic assay in the magnetic bead-based microfluidic system within 60 min.

 

Future Perspectives

Traditional foodborne pathogen detection methods, although sensitive enough, are often too time-consuming for practical use, taking days to a week to perform. Therefore, new methods that overcome this performance limitation are required. Recently, several methods have been explored and developed for the rapid detection of foodborne pathogens. However, most of them still require improvement in sensitivity, selectivity, or accuracy to be of any practical use.

Nucleic-acid-based methods have high sensitivity and require a shorter time than conventional culture-based techniques for detection of foodborne pathogens and toxins, but most of them require trained personnel and expensive instruments, which limit their use in a practical environment. The emerging isothermal amplification methods such as LAMP and NASBA may have a good prospect for detection of pathogens and toxins in resourcelimit settings. The development of nucleic-acid-based methods and immunological methods helped improve the time required to yield results. The specificity and the sensitivity of immunological methods depend on the binding strength of the specific antibody to its antigen, and they work well for food matrixes without interfering factors such as other non-target cells, DNA, and proteins. Biosensors-based methods are easy to perform without training, and yield results in real-time detection of foodborne pathogens and toxins with high sensitivity and selectivity comparable to the culture-based methods. However, they still need to be improved in food matrixes detection.

All assays available for food diagnostics require some degree of sample preparation, which is a very important factor for rapid and conventional detection methods, and also a bottleneck for the advanced rapid methods. More studies regarding the separation techniques of microorganisms from the food matrix are required, as well as for sample concentration prior to detection by immunological, nucleicacid-based, or biosensor assays. Preconcentration is the preferred choice, as it can enhance sensitivity several folds by increasing the number of target organisms per unit volume at a relatively low cost. Several available modes of preconcentration are used, including filtration, sizefractionation, centrifugation, and immunomagnetic separation, or combinations of these methods.

The possibilities of combining various rapid methods, including nucleic-acid-based methods, immunologicalbased methods, and biosensor-based methods should be further exploited. With the correct application of a number of these technologies simultaneously, broader ranging and more accurate technologies could be developed. Antibodies can be modified to capture specific cells, which may then be detected by a nucleic-acid-based method. Various nucleic acid amplified products can be quantified using immunoassays. The trend in immunoassays and nucleic-acid-based methods should result in the quantitative detection of microorganisms and the simultaneous determination of more than one pathogen or toxin. For immunological-based methods, further study of the application of biosensor chips may result in multiplex analyte assays. Biosensors must prove that they are capable of reaching at least the same detection levels as traditional methods (between 10 and 100 CFU/ml) in order to strengthen their appeal in food microbiology applications, not to mention the costeffectiveness and time efficiency. Despite the numerous research efforts made during the past decades and in recent years for foodborne pathogen detection, current technologies still entail room for improvement. Since foodborne pathogens are mostly present in very low numbers (<100 CFU/g) and in the presence of millions of other bacteria, they are not easily detected. Therefore, a detection method that is reliable, accurate, rapid, simple, sensitive, selective, and cost-effective would be ideal. Such methods of pathogen detection would offer a great commercial advantage in the food industry and related fields. Moreover, the trend of crossing various methods will generate novel devices or methodologies to strengthen the advantages of rapid detection methods.

In summary, there are a host of promising applications in the field of rapid and automated detection methods for foodborne pathogens. Given the broad applicability and the great potential of such methods, there is still a great chance for further developments in the near future.

References

  1. Abbas A, Linman MJ, Cheng QA. 2011. New trends in instrumental design for surface plasmon resonance-based biosensors. Biosens. Bioelectron. 26: 1815-1824. https://doi.org/10.1016/j.bios.2010.09.030
  2. Arora P, Sindhu A, Dilbaghi N, Chaudhury A. 2011. Biosensors as innovative tools for the detection of food borne pathogens. Biosens. Bioelectron. 28: 1-12. https://doi.org/10.1016/j.bios.2011.06.002
  3. Asiello PJ, Baeumner AJ. 2011. Miniaturized isothermal nucleic acid amplification, a review. Lab. Chip 11: 1420- 1430. https://doi.org/10.1039/c0lc00666a
  4. Baeumner AJ, Cohen RN, Miksic V, Min J. 2003. RNA biosensor for the rapid detection of viable Escherichia coli in drinking water. Biosens. Bioelectron. 18: 405-413. https://doi.org/10.1016/S0956-5663(02)00162-8
  5. Bolton FJ, Fritz E, Poynton S, Jensen T. 2000. Rapid enzyme-linked immunoassay for detection of Salmonella in food and feed products: performance testing program. J. AOAC Int. 83: 299-303.
  6. Campbell GA, Mutharasan R. 2006. Piezoelectric-excited millimeter-sized cantilever (PEMC) sensors detect Bacillus anthracis at 300 spores/ml. Biosens. Bioelectron. 21: 1684- 1692. https://doi.org/10.1016/j.bios.2005.08.001
  7. Cao W, S u M, Z hang S . 2010. Rap id and s ensitive DNA target detection using enzyme amplified electrochemical detection based on microchip. Electrophoresis 31: 659-665. https://doi.org/10.1002/elps.200900538
  8. CDC 2011. CDC Estimates of Foodborne Illness in the United States. [Online.] http://www.cdc.gov/foodborneburden/ 2011-foodborne-estimates.html
  9. Chamberlain JS, Gibbs RA, Ranier JE, Nguyen PN, Caskey CT. 1988. Deletion screening of the Duchenne muscular dystrophy locus via multiplex DNA amplification. Nucleic Acids Res. 16: 11141-11156. https://doi.org/10.1093/nar/16.23.11141
  10. Chen HM, Lin CW. 2007. Hydrogel-coated streptavidin piezoelectric biosensors and applications to selective detection of Strep-Tag displaying cells. Biotechnol. Prog. 23: 741-748.
  11. Chen HT, Zhang J, Sun DH, Ma LN, Liu XT, Cai XP, Liu YS. 2008. Development of reverse transcription loopmediated isothermal amplification for rapid detection of H9 avian influenza virus. J. Virol. Methods 151: 200-203. https://doi.org/10.1016/j.jviromet.2008.05.009
  12. Chen J, Ma XY, Yuan YW, Zhang W. 2011. Sensitive and rapid detection of Alicyclobacillus acidoterrestris u sing loop - mediated isothermal amplification. J. Sci. Food Agric. 91: 1070-1074. https://doi.org/10.1002/jsfa.4285
  13. Chen SY, Wang F, Beaulieu JC, Stein RE, Ge BL. 2011. Rapid detection of viable salmonellae in produce by coupling propidium monoazide with loop-mediated isothermal amplification. Appl. Environ. Microbiol. 77: 4008-4016. https://doi.org/10.1128/AEM.00354-11
  14. Chen ZG. 2008. Conductometric immunosensors for the detection of staphylococcal enterotoxin B based bioelectrocalytic reaction on micro-comb electrodes. Bioproc. Biosyst. Eng. 31: 345-350. https://doi.org/10.1007/s00449-007-0168-2
  15. Chuang TL, Wei SC, Lee SY, Lin CW. 2012. A polycarbonate based surface plasmon resonance sensing cartridge for high sensitivity HBV loop-mediated isothermal amplification. Biosens. Bioelectron. 32: 89-95. https://doi.org/10.1016/j.bios.2011.11.037
  16. Churruca E, Girbau C, Martinez I, Mateo E, Alonso R, Fernandez-Astorga A. 2007. Detection of Campylobacter jejuni and Campylobacter coli in chicken meat samples by real-time nucleic acid sequence-based amplification with molecular beacons. Int. J. Food Microbiol. 117: 85-90. https://doi.org/10.1016/j.ijfoodmicro.2007.02.007
  17. Compton J. 1991. Nucleic acid sequence-based amplification. Nature 350: 91-92. https://doi.org/10.1038/350091a0
  18. de Boer E, Beumer RR. 1999. Methodology for detection and typing of foodborne microorganisms. Int. J. Food Microbiol. 50: 119-130. https://doi.org/10.1016/S0168-1605(99)00081-1
  19. DeCory TR, Durst RA, Zimmerman SJ, Garringer LA, Paluca G, DeCory HH, Montagna RA. 2005. Development of an immunomagnetic bead-immunoliposome fluorescence assay for rapid detection of Escherichia coli O157:H7 in aqueous samples and comparison of the assay with a standard microbiological method. Appl. Environ. Microbiol. 71: 1856-1864. https://doi.org/10.1128/AEM.71.4.1856-1864.2005
  20. Delmulle BS, De Saeger SM, Sibanda L, Barna-Vetro I, Van Peteghem CH. 2005. Development of an immunoassaybased lateral flow dipstick for the rapid detection of aflatoxin B1 in pig feed. J. Agric. Food Chem. 53: 3364-3368. https://doi.org/10.1021/jf0404804
  21. Derzelle S, Grine A, Madic J, de Garam CP, Vingadassalon N, Dilasser F, et al. 2011. A quantitative PCR assay for the detection and quantification of Shiga toxin-producing Escherichia coli (STEC) in minced beef and dairy products. Int. J. Food Microbiol. 151: 44-51. https://doi.org/10.1016/j.ijfoodmicro.2011.07.039
  22. Dey D, Goswami T. 2011. Optical Biosensors: A revolution towards quantum nanoscale electronics device fabrication. J. Biomed. Biotechnol. 2011: 348218.
  23. Dongyou L. 2010. Molecular Detection of Foodborne Pathogens. CRC Press, Boca Raton.
  24. Dwivedi HP, Jaykus LA. 2011. Detection of pathogens in foods: the current state-of-the-art and future directions. Crit. Rev. Microbiol. 37: 40-63. https://doi.org/10.3109/1040841X.2010.506430
  25. Feng P. 1997. Impact of molecular biology on the detection of foodborne pathogens. Mol. Biotechnol. 7: 267-278. https://doi.org/10.1007/BF02740817
  26. Foley SL, Grant K. 2007. Molecular techniques of detection and discrimination of foodborne pathogens and their toxins, p. 485-510. In Simjee S (ed.). Infectious Disease: Foodborne Diseases. Humana Press Inc, Totowa, NJ.
  27. Fusco V, Quero GM, Morea M, Blaiotta G, Visconti A. 2011. Rapid and reliable identification of Staphylococcus aureus harbouring the enterotoxin gene cluster (egc) and quantitative detection in raw milk by real time PCR. Int. J. Food Microbiol. 144: 528-537. https://doi.org/10.1016/j.ijfoodmicro.2010.11.016
  28. Fykse EM, Skogan G, Davies W, Olsen JS, Blatny JM. 2007. Detection of Vibrio cholerae by real-time nucleic acid sequence-based amplification. Appl. Environ. Microbiol. 73: 1457-1466. https://doi.org/10.1128/AEM.01635-06
  29. Gill P, Ghaemi A. 2008. Nucleic acid isothermal amplification technologies: a review. Nucleosides Nucleotides Nucleic Acids 27: 224-243. https://doi.org/10.1080/15257770701845204
  30. Gill P, Ramezani R, Amiri MVP, Ghaemi A, Hashempour T, Eshraghi N, et al. 2006. Enzyme-linked immunosorbent assay of nucleic acid sequence-based amplification for molecular detection of M-tuberculosis. Biochem. Biophys. Res. Commun. 347: 1151-1157. https://doi.org/10.1016/j.bbrc.2006.07.039
  31. Gizeli E, Lowe CR. 1996. Immunosensors. Curr. Opin. Biotechnol. 7: 66-71. https://doi.org/10.1016/S0958-1669(96)80097-8
  32. Gomez P, Pagnon M, Egea-Cortines M, Artes F, Weiss J. 2010. A fast molecular nondestructive protocol for evaluating aerobic bacterial load on fresh-cut lettuce. Food Sci. Technol. Int. 16: 409-415. https://doi.org/10.1177/1082013210366882
  33. Grothaus GD, Bandla M, Currier T, Giroux R, Jenkins GR, Lipp M, et al. 2006. Immunoassay as an analytical tool in agricultural biotechnology. J. AOAC Int. 89: 913-928.
  34. Guatelli JC, Whitfield KM, Kwoh DY, Barringer KJ, Richman DD, Gingeras TR. 1990. Isothermal, in vitro amplification of nucleic acids by a multienzyme reaction modeled after retroviral replication. Proc. Natl. Acad. Sci. USA 87: 7797.
  35. Hahm BK, Bhunia AK. 2006. Effect of environmental stresses on antibody-based detection of Escherichia coli O157:H7, Salmonella enterica serotype Enteritidis and Listeria monocytogenes. J. Appl. Microbiol. 100: 1017-1027. https://doi.org/10.1111/j.1365-2672.2006.02814.x
  36. Hahn MA, Keng PC, Krauss TD. 2008. Flow cytometric analysis to detect pathogens in bacterial cell mixtures using semiconductor quantum dots. Anal. Chem. 80: 864-872. https://doi.org/10.1021/ac7018365
  37. Hill WE. 1996. The polymerase chain reaction: applications for the detection of foodborne pathogens. Crit. Rev. Food Sci. Nutr. 36: 123-173. https://doi.org/10.1080/10408399609527721
  38. Ikeda S, Takabe K, Inagaki M, Funakoshi N, Suzuki K. 2007. Detection of gene point mutation in paraffin sections using in situ loop-mediated isothermal amplification. Pathol. Int. 57: 594-599. https://doi.org/10.1111/j.1440-1827.2007.02144.x
  39. Iseki H, Alhassan A, Ohta N, Thekisoe OMM, Yokoyama N, Inoue N, et al. 2007. Development of a multiplex loopmediated isothermal amplification (mLAMP) method for the simultaneous detection of bovine Babesia parasites. J. Microbiol. Methods 71: 281-287. https://doi.org/10.1016/j.mimet.2007.09.019
  40. Jasson V, Jacxsens L, Luning P, Rajkovic A, Uyttendaele M. 2010. Alternative microbial methods: an overview and selection criteria. Food Microbiol. 27: 710-730. https://doi.org/10.1016/j.fm.2010.04.008
  41. Jung BY, Jung SC, K weon CH. 2005. Development of a rapid immunochromatographic strip for detection of Escherichia coli O157. J. Food Prot. 68: 2140-2143. https://doi.org/10.4315/0362-028X-68.10.2140
  42. Kawasaki S, Fratamico PM, Horikoshi N, Okada Y, Takeshita K, Sameshima T, Kawamoto S. 2009. Evaluation of a multiplex PCR system for simultaneous detection of Salmonella spp., listeria monocytogenes, and Escherichia coli O157:H7 in foods and in food subjected to freezing. Foodborne Pathog. Dis. 6: 81-89. https://doi.org/10.1089/fpd.2008.0153
  43. Kawasaki S, Horikoshi N, Okada Y, Takeshita K, Sameshima T, Kawamoto S. 2005. Multiplex PCR for simultaneous detection of Salmonella spp., Listeria monocytogenes, and Escherichia coli O157:H7 in meat samples. J. Food Prot. 68: 551-556. https://doi.org/10.4315/0362-028X-68.3.551
  44. Kim HJ, Lee HJ, Lee KH, Cho JC. 2012. Simultaneous detection of pathogenic Vibrio species using multiplex realtime PCR. Food Control 23: 491-498. https://doi.org/10.1016/j.foodcont.2011.08.019
  45. Kong RY, Lee SK, Law TW, Law SH, Wu RS. 2002. Rapid detection of six types of bacterial pathogens in marine waters by multiplex PCR. Water Res. 36: 2802-2812. https://doi.org/10.1016/S0043-1354(01)00503-6
  46. Lan YB, Wang SZ, Yin YG, Hoffmann WC, Zheng XZ. 2008. Using a surface plasmon resonance biosensor for rapid detection of Salmonella Typhimurium in chicken carcass. J. Bionic Eng. 5: 239-246. https://doi.org/10.1016/S1672-6529(08)60030-X
  47. Laura A, Gilda D, Claudio B, Cristina G, Gianfranco G. 2011. A lateral flow immunoassay for measuring ochratoxin A: development of a single system for maize, wheat and durum wheat. Food Control 22: 1965-1970. https://doi.org/10.1016/j.foodcont.2011.05.012
  48. Lazcka O, Del Campo FJ, Munoz FX. 2007. Pathogen detection: A perspective of traditional methods and biosensors. Biosens. Bioelectron. 22: 1205-1217. https://doi.org/10.1016/j.bios.2006.06.036
  49. Lee JY, Deininger RA. 2004. Detection of E. coli in beach water within 1 hour using immunomagnetic separation and ATP bioluminescence. Luminescence 19: 31-36. https://doi.org/10.1002/bio.753
  50. Leonard P, Hearty S, Brennan J, Dunne L, Quinn J, Chakraborty T, O'Kennedy R. 2003. Advances in biosensors for detection of pathogens in food and water. Enzyme Microb. Technol. 32: 3-13. https://doi.org/10.1016/S0141-0229(02)00232-6
  51. Li Y, Cheng P, Gong JH, Fang LC, Deng J, Liang WB, Zheng JS. 2012. Amperometric immunosensor for the detection of Escherichia coli O157:H7 in food specimens. Anal. Biochem. 421: 227-233. https://doi.org/10.1016/j.ab.2011.10.049
  52. Li Y, Mustapha A. 2004. Simultaneous detection of Escherichia coli O157:H7, Salmonella, and Shigella in apple cider and produce by a multiplex PCR. J. Food Prot. 67: 27-33. https://doi.org/10.4315/0362-028X-67.1.27
  53. Lin WS, Cheng CM, Van KT. 2010. A quantitative PCR assay for rapid detection of Shigella species in fresh produce. J. Food Prot. 73: 221-233. https://doi.org/10.4315/0362-028X-73.2.221
  54. Lin YH, Chen SH, Chuang YC, Lu YC, Shen TY, Chang CA, Lin CS. 2008. Disposable amperometric immunosensing strips fabricated by Au nanoparticles-modified screenprinted carbon electrodes for the detection of foodborne pathogen Escherichia coli O157:H7. Biosens Bioelectron. 23: 1832-1837. https://doi.org/10.1016/j.bios.2008.02.030
  55. Liu Y, Chuang CK, Chen WJ. 2009. In situ reversetranscription loop-mediated isothermal amplification (in situ RT-LAMP) for detection of Japanese encephalitis viral RNA in host cells. J. Clin. Virol. 46: 49-54. https://doi.org/10.1016/j.jcv.2009.06.010
  56. Loens K, Beck T, Goossens H, Ursi D, Overdijk M, Sillekens P, Ieven M. 2006. Development of conventional and real-time nucleic acid sequence-based amplification assays for detection of Chlamydophila pneumoniae in respiratory specimens. J. Clin. Microbiol. 44: 1241-1244. https://doi.org/10.1128/JCM.44.4.1241-1244.2006
  57. Louie AS, Marenchic IG, Whelan RH. 1998. A fieldable modular biosensor for use in detection of foodborne pathogens. Field Anal. Chem. Technol. 2: 371-377. https://doi.org/10.1002/(SICI)1520-6521(1998)2:6<371::AID-FACT7>3.0.CO;2-F
  58. Luo XL, Xu JJ, Zhao W, Chen HY. 2004. Glucose biosensor based on E NFET d op ed w ith $SiO_{2}$ nanoparticles. Sensors Actuat B Chem. 97: 249-255. https://doi.org/10.1016/j.snb.2003.08.024
  59. Magliulo M, Simoni P, Guardigli M, Michelini E, Luciani M, Lelli R, Roda A. 2007. A rapid multiplexed chemiluminescent immunoassay for the detection of Escherichia coli O157:H7, Yersinia enterocolitica, Salmonella typhimurium, and Listeria monocytogenes pathogen bacteria. J. Agric. Food Chem. 55: 4933-4939. https://doi.org/10.1021/jf063600b
  60. Mandal PK, Biswas AK, Choi K, Pal U. 2011. Methods for rapid detection of foodborne pathogens: an overview. Am. J. Food Technol. 6: 87-102. https://doi.org/10.3923/ajft.2011.87.102
  61. Martinon A, Wilkinson MG. 2011. Selection of optimal primer sets for use in a duplex sybr green-based, real-time polymerase chain reaction protocol for the detection of listeria monocytogenes and staphyloccocus aureus in foods. J. Food Saf. 31: 297-312. https://doi.org/10.1111/j.1745-4565.2011.00301.x
  62. Maruyama F, Kenzaka T, Yamaguchi N, Tani K, Nasu M. 2003. Detection of bacteria carrying the stx2 gene by in situ loop-mediated isothermal amplification. Appl. Environ. Microbiol. 69: 5023-5028. https://doi.org/10.1128/AEM.69.8.5023-5028.2003
  63. McEgan R, Fu TJ, Warriner K. 2009. Concentration and detection of Salmonella in mung bean sprout spent irrigation water by use of tangential flow filtration coupled with an amperometric flowthrough enzyme-linked immunosorbent assay. J. Food Prot. 72: 591-600. https://doi.org/10.4315/0362-028X-72.3.591
  64. Meeusen CA, Alocilja EC, Osburn WN. 2005. Detection of E. coli O157:H7 using a miniaturized surface plasmon resonance biosensor. Trans. ASAE 48: 2409-2416. https://doi.org/10.13031/2013.20067
  65. Min J, Baeumner AJ. 2002. Highly sensitive and specific detection of viable Escherichia coli in drinking water. Anal. Biochem. 303: 186-193. https://doi.org/10.1006/abio.2002.5593
  66. Mukhopadhyay A, Mukhopadhyay UK. 2007. Novel multiplex PCR approaches for the simultaneous detection of human pathogens: Escherichia coli O157:H7 and Listeria monocytogenes. J. Microbiol. Methods 68: 193-200. https://doi.org/10.1016/j.mimet.2006.07.009
  67. Muldoon MT, Teaney G, Li J, Onisk DV, Stave JW. 2007. Bacteriophage-based enrichment coupled to immunochromatographic strip-based detection for the determination of Salmonella in meat and poultry. J. Food Prot. 70: 2235-2242. https://doi.org/10.4315/0362-028X-70.10.2235
  68. Nam HM, Srinivasan V, Gillespie BE, Murinda SE, Oliver SP. 2005. Application of SYBR green real-time PCR assay for specific detection of Salmonella spp. in dairy farm environmental samples. Int. J. Food Microbiol. 102: 161-171. https://doi.org/10.1016/j.ijfoodmicro.2004.12.020
  69. Nemoto J, Ikedo M, Kojima T, Momoda T, Konuma H, Hara-Kudo Y. 2011. Development and evaluation of a loop-mediated isothermal amplification assay for rapid and sensitive detection of Vibrio parahaemolyticus. J. Food Prot. 74: 1462-1467. https://doi.org/10.4315/0362-028X.JFP-10-519
  70. Notomi T, Okayama H, Masubuchi H, Yonekawa T, Watanabe K, Amino N, Hase T. 2000. Loop-mediated isothermal amplification of DNA. Nucleic Acids Res. 28: E63. https://doi.org/10.1093/nar/28.12.e63
  71. Ohtsuka K, Tanaka M, Ohtsuka T, Takatori K, Hara-Kudo Y. 2010. Comparison of detection methods for Escherichia coli O157 in beef livers and carcasses. Foodborne Pathog. Dis. 7: 1563-1567. https://doi.org/10.1089/fpd.2010.0585
  72. Oliver SP, Jayarao BM, Almeida RA. 2005. Foodborne pathogens in milk and the dairy farm environment: food safety and public health implications. Foodborne Pathog. Dis. 2: 115-129. https://doi.org/10.1089/fpd.2005.2.115
  73. Ong KG, Zeng KF, Yang XP, Shankar K, Ruan CM, Grimes CA. 2006. Quantification of multiple bioagents with wireless, remote-query magnetoelastic microsensors. IEEE Sensors J. 6: 514-523.
  74. Pal S, Ying W, Alocilja EC, Downes FP. 2008. Sensitivity and specificity performance of a direct-charge transfer biosensor for detecting Bacillus cereus in selected food matrices Biosys Eng. 99: 461-468. https://doi.org/10.1016/j.biosystemseng.2007.11.015
  75. Park YS, Lee SR, Kim YG. 2006. Detection of Escherichia coli O157:H7, Salmonella spp., Staphylococcus aureus and Listeria monocytogenes in kimchi by multiplex polymerase chain reaction (mPCR). J. Microbiol. 44: 92-97.
  76. Pedrero M, Campuzano S, Pingarron JM. 2009. Electroanalytical sensors and devices for multiplexed detection of foodborne pathogen microorganisms. Sensors 9: 5503-5520. https://doi.org/10.3390/s90705503
  77. Planche T, Aghaizu A, Holliman R, Riley P, Poloniecki J, Breathnach A, Krishna S. 2008. Diagnosis of Clostridium difficile infection by toxin detection kits: a systematic review. Lancet Infect. Dis. 8: 777-784. https://doi.org/10.1016/S1473-3099(08)70233-0
  78. Posthuma-Trump ie GA, Korf J, van Amerongen A . 2009. Lateral flow (immuno) assay: its strengths, weaknesses, opportunities and threats. A literature survey. Anal. Bioanal. Chem. 393: 569-582. https://doi.org/10.1007/s00216-008-2287-2
  79. Qiao YM, Guo YC, Zhang XE, Zhou YF, Zhang ZP, Wei HP, et al. 2007. Loop-mediated isothermal amplification for rapid detection of Bacillus anthracis spores. Biotechnol. Lett. 29: 1939-1946. https://doi.org/10.1007/s10529-007-9472-9
  80. Rahman S, Lipert RJ, Porter MD. 2006. Rapid screening of pathogenic bacteria using solid phase concentration and diffuse reflectance spectroscopy. Anal. Chim. Acta 569: 83-90. https://doi.org/10.1016/j.aca.2006.03.083
  81. Rao VK, Rai GP, Agarwal GS, Suresh S. 2005. Amperometric immunosensor for detection of antibodies of Salmonella Typhi in patient serum. Anal. Chim. Acta 531: 173-177. https://doi.org/10.1016/j.aca.2004.10.015
  82. Rasooly A, Herold KE. 2006. Biosensors for the analysis of food- and waterborne pathogens and their toxins. J. AOAC Int. 89: 873-883.
  83. Rodriguez-Lazaro D, Jofre A, Aymerich T, Hugas M, Pla M. 2004. Rapid quantitative detection of Listeria monocytogenes in meat products by real-time PCR. Appl. Environ. Microbiol. 70: 6299-6301. https://doi.org/10.1128/AEM.70.10.6299-6301.2004
  84. Rosebrough SF, Hartley DF. 1996. Biochemical modification of streptavidin and avidin: in vitro and in vivo analysis. J. Nucl. Med. 37: 1380-1384.
  85. Salmain M, Ghasemi M, Boujday S, Spadavecchia J, Techer C, Val F, et al. 2011. Piezoelectric immunosensor for direct and rapid detection of staphylococcal enterotoxin A (SEA) at the ng level. Biosens. Bioelectron. 29: 140-144. https://doi.org/10.1016/j.bios.2011.08.007
  86. Sankaran S, Panigrahi S, Mallik S. 2011. Olfactory receptor based piezoelectric biosensors for detection of alcohols related to food safety applications. Sensors Actuat B Chem. 155: 8-18. https://doi.org/10.1016/j.snb.2010.08.003
  87. Schlosser G, Kacer P, Kuzma M, Szilagyi Z, Sorrentino A, Manzo C, et al. 2007. Coupling immunomagnetic separation on magnetic beads with matrix-assisted laser desorption ionization-time of flight mass spectrometry for detection of staphylococcal enterotoxin B. Appl. Environ. Microbiol. 73: 6945-6952. https://doi.org/10.1128/AEM.01136-07
  88. Seki M, Yamashita Y, Torigoe H, Tsuda H, Sato S, Maeno M. 2005. Loop-mediated isothermal amplification method targeting the lytA gene for detection of Streptococcus pneumoniae. J. Clin. Microbiol. 43: 1581-1586. https://doi.org/10.1128/JCM.43.4.1581-1586.2005
  89. Seo KH, Brackett RE, Hartman NF, Campbell DP. 1999. Development of a rapid response biosensor for detection of Salmonella typhimurium. J. Food Prot. 62: 431-437. https://doi.org/10.4315/0362-028X-62.5.431
  90. Sharma H, Mutharasan R. 2013. Review of biosensors for foodborne pathogens and toxins. Sensors Actuat. B Chem. 183: 535-549. https://doi.org/10.1016/j.snb.2013.03.137
  91. Shi XM, Long F, Suo B. 2010. Molecular methods for the detection and characterization of foodborne pathogens. Pure Appl. Chem. 82: 69-79. https://doi.org/10.1351/PAC-CON-09-02-07
  92. Si CY, Ye ZZ, Wang YX, Gai L, Wang JP, Ying YB. 2011. Rapid detection of Escherichia coli O157:H7 using surface plasmon resonance (SPR) biosensor. Spectrosc. Spectral Anal. 31: 2598-2601.
  93. Simpkins SA, Chan AB, Hays J , Popping B, Cook N. 2000. An RNA transcription-based amplification technique (NASBA) for the detection of viable Salmonella enterica. Lett. Appl. Microbiol. 30: 75-79. https://doi.org/10.1046/j.1472-765x.2000.00670.x
  94. Singh C, Agarwal GS, Rai GP, Singh L, Rao VK. 2005. Specific detection of Salmonella Typhi using renewable amperometric immunosensor. Electroanalysis 17: 2062-2067. https://doi.org/10.1002/elan.200403334
  95. Song TY, Toma C, Nakasone N, Iwanaga M. 2005. Sensitive and rapid detection of Shigella and enteroinvasive Escherichia coli by a loop-mediated isothermal amplification method. FEMS Microbiol. Lett. 243: 259-263. https://doi.org/10.1016/j.femsle.2004.12.014
  96. Spangler BD, Wilkinson EA, Murphy JT, Tyler BJ. 2001. Comparison of the Spreeta (R) surface plasmon resonance sensor and a quartz crystal microbalance for detection of Escherichia coli heat-labile enterotoxin. Anal. Chim. Acta 444: 149-161. https://doi.org/10.1016/S0003-2670(01)01156-4
  97. Stevens KA, Jaykus LA. 2004. Bacterial separation and concentration from complex sample matrices: a review. Crit. Rev. Microbiol. 30: 7-24. https://doi.org/10.1080/10408410490266410
  98. Tokarskyy O, Marshall DL. 2008. Immunosensors for rapid detection of Escherichia coli O157 : H7-perspectives for use in the meat processing industry. Food Microbiol. 25: 1-12.
  99. Tsaloglou MN, Bahi MM, Waugh EM, Morgan H, Mowlem M. 2011. On-chip real-time nucleic acid sequence-based amplification for RNA detection and amplification. Anal. Methods 3: 2127-2133. https://doi.org/10.1039/c1ay05164d
  100. Velusamy V, Arshak K, Korostynska O, Oliwa K, Adley C. 2010. An overview of foodborne pathogen detection: in the perspective of biosensors. Biotechnol. Adv. 28: 232-254. https://doi.org/10.1016/j.biotechadv.2009.12.004
  101. Vernozy-Rozand C, Mazuy-Cruchaudet C, Bavai C, Richard Y. 2004. Comparison of three immunological methods for detecting staphylococcal enterotoxins from food. Lett. Appl. Microbiol. 39: 490-494. https://doi.org/10.1111/j.1472-765X.2004.01602.x
  102. Verstraete K, Robyn J, Del-Favero J, De Rijk P, Joris MA, Herman L, et al. 2012. Evaluation of a multiplex-PCR detection in combination with an isolation method for STEC O26, O103, O111, O145 and sorbitol fermenting O157 in food. Food Microbiol. 29: 49-55. https://doi.org/10.1016/j.fm.2011.08.017
  103. Vo-Dinh T, Griffin G, Stokes DL, Wintenberg A. 2003. Multi-functional biochip for medical diagnostics and pathogen detection. Sensors Actuat B Chem. 90: 104-111. https://doi.org/10.1016/S0925-4005(03)00048-0
  104. Wang CH, Lien KY, Wu JJ, Lee GB. 2011. A magnetic bead-based assay for the rapid detection of methicillinresistant Staphylococcus aureus by using a microfluidic system with integrated loop-mediated isothermal amplification. Lab. Chip 11: 1521-1531. https://doi.org/10.1039/c0lc00430h
  105. Wang DG, Wang YZ, Wang JH, Zhang XW, Xiao FG. 2011. Rapid detection of viable Listeria monocytogenes in raw milk using loop-mediated isothermal amplification with the aid of ethidium monoazide. Milchwissenschaft Milk Sci. Int. 66: 426-429.
  106. Wang LX, Li Y, Mustapha A. 2007. Rapid and simultaneous quantitation of Escherichia coli O157:H7, Salmonella, and Shigella in ground beef by multiplex realtime PCR and immunomagnetic separation. J. Food Prot. 70: 1366-1372. https://doi.org/10.4315/0362-028X-70.6.1366
  107. Wang YX, Ye ZZ, Si CY, Ying YB. 2011. Subtractive inhibition assay for the detection of E. coli O157:H7 using surface plasmon resonance. Sensors 11: 2728-2739. https://doi.org/10.3390/s110302728
  108. Waswa J, Irudayaraj J, DebRoy C. 2007. Direct detection of E. coli O157:H7 in selected food systems by a surface plasmon resonance biosensor. LWT Food Sci. Technol. 40: 187-192. https://doi.org/10.1016/j.lwt.2005.11.001
  109. Waswa JW, Debroy C, Irudayaraj J. 2006. Rapid detection of Salmonella enteritidis and Escherichia coli using surface plasmon resonance biosensor. J. Food Process Eng. 29: 373-385. https://doi.org/10.1111/j.1745-4530.2006.00071.x
  110. World Health Organization. 2007. Food safety and foodborne illness. [Online.] http://www.who.int/mediacentre/factsheets/ fs237/en/
  111. Wu VCH, Chen SH, Lin CS. 2007. Real-time detection of Escherichia coli O157:H7 sequences using a circulating-flow system of quartz crystal microbalance. Biosens. Bioelectron. 22: 2967-2975. https://doi.org/10.1016/j.bios.2006.12.016
  112. Yamazaki W, Ishibashi M, Kawahara R, Inoue K. 2008. Development of a loop-mediated isothermal amplification assay for sensitive and rapid detection of Vibrio parahaemolyticus. BMC Microbiol. 8.
  113. Yamazaki W, Kumeda Y, Uemura R, Misawa N. 2011. Evaluation of a loop-mediated isothermal amplification assay for rapid and simple detection of Vibrio parahaemolyticus in naturally contaminated seafood samples. Food Microbiol. 28: 1238-1241. https://doi.org/10.1016/j.fm.2011.04.007
  114. Yang H, Ma XY, Zhang XZ, Wang Y, Zhang W. 2011. Development and evaluation of a loop-mediated isothermal amplification assay for the rapid detection of Staphylococcus aureus in food. Eur. Food Res. Technol. 232: 769-776. https://doi.org/10.1007/s00217-011-1442-8
  115. Yang L, Bashir R. 2008. Electrical/electrochemical impedance for rapid detection of foodborne pathogenic bacteria. Biotechnol. Adv. 26: 135-150. https://doi.org/10.1016/j.biotechadv.2007.10.003
  116. Ye YX, Wang B, Huang F, Song YS, Yan H, Alam MJ, et al. 2011. Application of in situ loop-mediated isothermal amplification method for detection of Salmonella in foods. Food Control 22: 438-444. https://doi.org/10.1016/j.foodcont.2010.09.023
  117. Yoo JH, Choi SM, Choi JH, Kwon EY, Park C, Shin WS. 2008. Construction of internal control for the quantitative assay of Aspergillus fumigatus using real-time nucleic acid sequence-based amplification. Diagn. Microbiol. Infect. Dis. 60: 121-124. https://doi.org/10.1016/j.diagmicrobio.2007.08.001
  118. Zhao XH, Li YM, Wang L, You LJ, Xu ZB, Li L, et al. 2010. Development and application of a loop-mediated isothermal amplification method on rapid detection Escherichia coli O157 strains from food samples. Mol. Biol. Rep. 37: 2183- 2188. https://doi.org/10.1007/s11033-009-9700-6
  119. Zhao XH, Wang L, Chu J, Li YY, Li YM, Xu ZB, et al. 2010. Development and application of a rapid and simple loopmediated isothermal amplification method for food-borne Salmonella detection. Food Sci. Biotechnol. 19: 1655-1659. https://doi.org/10.1007/s10068-010-0234-4
  120. Zhao XH, Wang L, Chu J, Li YY, Li YM, Xu ZB, et al. 2010. Rapid detection of Vibrio parahaemolyticus strains and virulent factors by loop-mediated isothermal amplification assays. Food Sci. Biotechnol. 19: 1191-1197. https://doi.org/10.1007/s10068-010-0170-3
  121. Zhao XH, Wang L, Li YM, Xu ZB, Li L, He XW, et al. 2011. Development and application of a loop-mediated isothermal amplification method on rapid detection of Pseudomonas aeruginosa strains. World J. Microbiol. Biotechnol. 27: 181-184. https://doi.org/10.1007/s11274-010-0429-0

Cited by

  1. Rapid methods for the detection of foodborne bacterial pathogens: principles, applications, advantages and limitations vol.5, pp.None, 2014, https://doi.org/10.3389/fmicb.2014.00770
  2. Recent trends in the detection of pathogenic Escherichia coli O157 : H7 vol.9, pp.3, 2014, https://doi.org/10.1007/s13206-015-9208-9
  3. Analytical characterization of IgG-cTpp and IgG-Mn-cTpp conjugates vol.19, pp.11, 2014, https://doi.org/10.1142/s1088424615500984
  4. An insight into the isolation, enumeration, and molecular detection of Listeria monocytogenes in food vol.6, pp.None, 2014, https://doi.org/10.3389/fmicb.2015.01227
  5. Phylogenetic Analyses of Shigella and Enteroinvasive Escherichia coli for the Identification of Molecular Epidemiological Markers: Whole-Genome Comparative Analysis Does Not Support Distinct Gener vol.6, pp.None, 2014, https://doi.org/10.3389/fmicb.2015.01573
  6. Diffusion and Persistence of Multidrug Resistant Salmonella Typhimurium Strains Phage Type DT120 in Southern Italy vol.2015, pp.None, 2014, https://doi.org/10.1155/2015/265042
  7. Phages in the global fruit and vegetable industry vol.118, pp.3, 2014, https://doi.org/10.1111/jam.12700
  8. Waterborne Pathogens: Detection Methods and Challenges vol.4, pp.2, 2014, https://doi.org/10.3390/pathogens4020307
  9. Foodomics for investigations of food toxins vol.4, pp.None, 2015, https://doi.org/10.1016/j.cofs.2015.05.004
  10. Development of a Loop‐Mediated Isothermal Amplification Assay Based on lmo0460 Sequence for Detection of Listeria monocytogenes vol.35, pp.3, 2014, https://doi.org/10.1111/jfs.12183
  11. A Review on Lateral Flow Test Strip for Food Safety vol.40, pp.3, 2014, https://doi.org/10.5307/jbe.2015.40.3.277
  12. Portable Nanoparticle-Based Sensors for Food Safety Assessment vol.15, pp.12, 2014, https://doi.org/10.3390/s151229826
  13. Novel Development of a qPCR Assay Based on the rpoB Gene for Rapid Detection of Cronobacter spp. vol.72, pp.4, 2014, https://doi.org/10.1007/s00284-015-0971-y
  14. Nucleic acid-based biotechnologies for food-borne pathogen detection using routine time-intensive culture-based methods and fast molecular diagnostics vol.25, pp.1, 2014, https://doi.org/10.1007/s10068-016-0002-1
  15. On-chip PMA labeling of foodborne pathogenic bacteria for viable qPCR and qLAMP detection vol.20, pp.8, 2016, https://doi.org/10.1007/s10404-016-1778-2
  16. Molecular characterization and phylogeny of Shiga toxin-producing E. coli (STEC) from imported beef meat in Malaysia vol.23, pp.17, 2014, https://doi.org/10.1007/s11356-016-6954-0
  17. Research advance in rapid detection of foodborne Staphylococcus aureus vol.30, pp.5, 2014, https://doi.org/10.1080/13102818.2016.1209433
  18. Detection of KRAS mutations of colorectal cancer with peptide-nucleic-acid-mediated real-time PCR clamping vol.30, pp.6, 2014, https://doi.org/10.1080/13102818.2016.1228479
  19. Rapid multiplex detection of 10 foodborne pathogens with an up-converting phosphor technology-based 10-channel lateral flow assay vol.6, pp.None, 2014, https://doi.org/10.1038/srep21342
  20. Porous Silicon-Based Biosensors: Towards Real-Time Optical Detection of Target Bacteria in the Food Industry vol.6, pp.None, 2014, https://doi.org/10.1038/srep38099
  21. The antimicrobial peptide aureocin A53 as an alternative agent for biopreservation of dairy products vol.121, pp.2, 2014, https://doi.org/10.1111/jam.13189
  22. Evaluation of Molecular Methods for Identification of Salmonella Serovars vol.54, pp.8, 2014, https://doi.org/10.1128/jcm.00262-16
  23. Potential Application of Fluorescence Imaging for Assessing Fecal Contamination of Soil and Compost Maturity vol.6, pp.9, 2014, https://doi.org/10.3390/app6090243
  24. Biosensors and their Applications in Food Safety: A Review vol.41, pp.3, 2016, https://doi.org/10.5307/jbe.2016.41.3.240
  25. Multiparametric Magneto-fluorescent Nanosensors for the Ultrasensitive Detection of Escherichia coli O157:H7 vol.2, pp.10, 2016, https://doi.org/10.1021/acsinfecdis.6b00108
  26. Multiplex PCR for simultaneous identification of E. coli O157:H7, Salmonella spp. and L. monocytogenes in food vol.6, pp.2, 2014, https://doi.org/10.1007/s13205-016-0523-6
  27. Nanotechnology in food safety and quality assessment: potentiality of nanoparticles in diagnosis of foodborne pathogens vol.13, pp.1, 2014, https://doi.org/10.1515/agricultura-2017-0004
  28. Biofilm formation and control strategies of foodborne pathogens: food safety perspectives vol.7, pp.58, 2014, https://doi.org/10.1039/c7ra02497e
  29. Rapid label-free visual detection of KRAS mutations using peptide nucleic acid and unmodified gold nanoparticles vol.7, pp.77, 2014, https://doi.org/10.1039/c7ra09088a
  30. Current Perspectives on Viable but Non-culturable State in Foodborne Pathogens vol.8, pp.None, 2014, https://doi.org/10.3389/fmicb.2017.00580
  31. Metagenomics: The Next Culture-Independent Game Changer vol.8, pp.None, 2014, https://doi.org/10.3389/fmicb.2017.01069
  32. A Novel Approach to Predict the Growth of Staphylococcus aureus on Rice Cake vol.8, pp.None, 2014, https://doi.org/10.3389/fmicb.2017.01140
  33. Microbial Quality, Safety, and Pathogen Detection by Using Quantitative PCR of Raw Salad Vegetables Sold in Dhanbad City, India vol.80, pp.1, 2014, https://doi.org/10.4315/0362-028x.jfp-16-223
  34. Multiplex PCR for detection of water-borne bacteria vol.17, pp.1, 2014, https://doi.org/10.2166/ws.2016.126
  35. How can nanosensors detect bacterial contamination before it ever reaches the dinner table? vol.12, pp.2, 2014, https://doi.org/10.2217/fmb-2016-0202
  36. Electrochemical Detection of Escherichia coli from Aqueous Samples Using Engineered Phages vol.89, pp.3, 2014, https://doi.org/10.1021/acs.analchem.6b03752
  37. Design and Elementary Evaluation of a Highly-Automated Fluorescence-Based Instrument System for On-Site Detection of Food-Borne Pathogens vol.17, pp.3, 2014, https://doi.org/10.3390/s17030442
  38. Bioelectrochemical Systems for Measuring Microbial Cellular Functions vol.29, pp.6, 2014, https://doi.org/10.1002/elan.201700110
  39. AC dielectrophoretic manipulation and electroporation of vaccinia virus using carbon nanoelectrode arrays vol.38, pp.11, 2014, https://doi.org/10.1002/elps.201600436
  40. Development of Simple Multiplex Real-Time PCR Assays for Foodborne Pathogens Detection and Identification On Lightcycler vol.40, pp.1, 2014, https://doi.org/10.1515/macvetrev-2017-0010
  41. Microcontact Imprinted Plasmonic Nanosensors: Powerful Tools in the Detection of Salmonella paratyphi vol.17, pp.6, 2014, https://doi.org/10.3390/s17061375
  42. Recombinase polymerase amplification: a promising point-of-care detection method for enteric viruses vol.12, pp.8, 2014, https://doi.org/10.2217/fvl-2017-0034
  43. Presenting a rapid method for detection of Bacillus cereus , Listeria monocytogenes and Campylobacter jejuni in food samples vol.20, pp.9, 2014, https://doi.org/10.22038/ijbms.2017.9275
  44. Evaluation of microplate immunocapture method for detection of Vibrio cholerae , Salmonella Typhi and Shigella flexneri from food vol.17, pp.None, 2014, https://doi.org/10.1186/s12866-017-1099-y
  45. Serogroup-level resolution of the “Super-7” Shiga toxin-producing Escherichia coli using nanopore single-molecule DNA sequencing vol.410, pp.22, 2018, https://doi.org/10.1007/s00216-018-0877-1
  46. Total Count of Salmonella typhimurium Coupled on Water Soluble CdSe Quantum Dots by Fluorescence Detection vol.47, pp.8, 2014, https://doi.org/10.1007/s11664-018-6347-x
  47. Au nanocluster-embedded chitosan nanocapsules as labels for the ultrasensitive fluorescence immunoassay of Escherichia coli O157:H7 vol.143, pp.17, 2014, https://doi.org/10.1039/c8an00987b
  48. Fabrication of an integrated polystyrene microdevice for pre-concentration and amplification ofEscherichia coliO157:H7 from raw milk vol.10, pp.42, 2018, https://doi.org/10.1039/c8ay01707g
  49. Direct enrichment of pathogens from physiological samples of high conductivity and viscosity using H-filter and positive dielectrophoresis vol.12, pp.1, 2018, https://doi.org/10.1063/1.5016413
  50. Advanced molecular diagnostic techniques for detection of food-borne pathogens: Current applications and future challenges vol.58, pp.1, 2014, https://doi.org/10.1080/10408398.2015.1126701
  51. The importance of lactic acid bacteria for the prevention of bacterial growth and their biogenic amines formation: A review vol.58, pp.10, 2014, https://doi.org/10.1080/10408398.2016.1277972
  52. Irrigation water quality and microbial safety of leafy greens in different vegetable production systems: A review vol.34, pp.4, 2014, https://doi.org/10.1080/87559129.2017.1289385
  53. Evaluation of Primers Detection Capabilities of the pef Salmonella typhimurium Gene and the fimC Eschericia coli Gene Using Real-Time PCR to Develop Rapid Detection of Food Poisoning Bacteria vol.1097, pp.None, 2018, https://doi.org/10.1088/1742-6596/1097/1/012049
  54. Predictive Modeling for the Growth of Salmonella Enteritidis in Chicken Juice by Real-Time Polymerase Chain Reaction vol.15, pp.7, 2018, https://doi.org/10.1089/fpd.2017.2392
  55. Detection of Colonized Pathogenic Bacteria from Food Handlers in Saudi Arabia vol.12, pp.3, 2014, https://doi.org/10.22207/jpam.12.3.32
  56. Detection of Foodborne Pathogens by Surface Enhanced Raman Spectroscopy vol.9, pp.None, 2014, https://doi.org/10.3389/fmicb.2018.01236
  57. Study the Features of 57 Confirmed CRISPR Loci in 38 Strains of Staphylococcus aureus vol.9, pp.None, 2014, https://doi.org/10.3389/fmicb.2018.01591
  58. Induction of Viable but Nonculturable Escherichia coli O157:H7 by Low Temperature and Its Resuscitation vol.9, pp.None, 2014, https://doi.org/10.3389/fmicb.2018.02728
  59. Surface-Enhanced Raman Scattering (SERS) With Silver Nano Substrates Synthesized by Microwave for Rapid Detection of Foodborne Pathogens vol.9, pp.None, 2014, https://doi.org/10.3389/fmicb.2018.02857
  60. Detection of viable but non-culturable Escherichia coli O157:H7 by PCR in combination with propidium monoazide vol.8, pp.1, 2014, https://doi.org/10.1007/s13205-017-1052-7
  61. Simultaneous detection of Escherichia coli O157:H7, Staphylococcus aureus and Salmonella by multiplex PCR in milk vol.8, pp.1, 2018, https://doi.org/10.1007/s13205-018-1086-5
  62. Plasmonic bacteria on a nanoporous mirror via hydrodynamic trapping for rapid identification of waterborne pathogens vol.7, pp.1, 2018, https://doi.org/10.1038/s41377-018-0071-4
  63. Occurrence and quantification of Shiga toxin-producing Escherichia coli from food matrices vol.11, pp.2, 2018, https://doi.org/10.14202/vetworld.2018.104-111
  64. Cronobacter Species in Powdered Infant Formula and Their Detection Methods vol.38, pp.2, 2014, https://doi.org/10.5851/kosfa.2018.38.2.376
  65. Development of a micromanipulation method for single cell isolation of prokaryotes and its application in food safety vol.13, pp.5, 2014, https://doi.org/10.1371/journal.pone.0198208
  66. Electrochemical Detection of E. coli O157:H7 in Water after Electrocatalytic and Ultraviolet Treatments Using a Polyguanine-Labeled Secondary Bead Sensor vol.18, pp.5, 2014, https://doi.org/10.3390/s18051497
  67. Multiplex PCR assay and lyophilization for detection of Salmonella spp., Staphylococcus aureus and Bacillus cereus in pork products vol.27, pp.3, 2014, https://doi.org/10.1007/s10068-017-0286-9
  68. Real-time and rapid detection ofSalmonellaTyphimurium using an inexpensive lab-built surface plasmon resonance setup vol.15, pp.7, 2014, https://doi.org/10.1088/1612-202x/aabed8
  69. Surface‐Enhanced Raman Scattering for Rapid Detection and Characterization of Antibiotic‐Resistant Bacteria vol.7, pp.13, 2014, https://doi.org/10.1002/adhm.201701335
  70. Peptide-Based Biosensor Utilizing Fluorescent Gold Nanoclusters for Detection of Listeria monocytogenes vol.1, pp.7, 2014, https://doi.org/10.1021/acsanm.8b00600
  71. Teicoplanin-functionalized magnetic beads for detection of Staphylococcus aureus via inhibition of the luminol chemiluminescence by intracellular catalase vol.185, pp.8, 2014, https://doi.org/10.1007/s00604-018-2921-4
  72. A PCR-Based Method for Distinguishing between Two Common Beehive Bacteria, Paenibacillus larvae and Brevibacillus laterosporus vol.84, pp.22, 2014, https://doi.org/10.1128/aem.01886-18
  73. An automated bacterial concentration and recovery system for pre-enrichment required in rapid Escherichia coli detection vol.8, pp.None, 2014, https://doi.org/10.1038/s41598-018-35970-8
  74. Gold nanoparticle-based colorimetric platform technology as rapid and efficient bacterial pathogens detection method from various sources vol.30, pp.2, 2014, https://doi.org/10.1097/mrm.0000000000000160
  75. Detection of Sarcocystis spp. and Shiga toxin-producing Escherichia coli in Japanese sika deer meat using a loop-mediated isothermal amplification-lateral flow strip vol.81, pp.4, 2014, https://doi.org/10.1292/jvms.18-0372
  76. Non-specific Liquid Fingerprinting in Monitoring the Hygiene and Authenticity of Milk vol.12, pp.1, 2014, https://doi.org/10.1007/s12161-018-1348-1
  77. GLAPD: Whole Genome Based LAMP Primer Design for a Set of Target Genomes vol.10, pp.None, 2019, https://doi.org/10.3389/fmicb.2019.02860
  78. Rapid and Accurate Diagnosis of the Respiratory Disease Pertussis on a Point-of-Care Biochip vol.8, pp.None, 2014, https://doi.org/10.1016/j.eclinm.2019.02.008
  79. Development of a multiplex real‐time PCR for simultaneous detection of BACILLUS CEREUS , LISTERIA MONOCYTOGENES , and STAPHYLOCOCCUS AUREUS in food vol.39, pp.1, 2014, https://doi.org/10.1111/jfs.12558
  80. LAMP-on-a-chip: Revising microfluidic platforms for loop-mediated DNA amplification vol.113, pp.None, 2019, https://doi.org/10.1016/j.trac.2019.01.015
  81. Simultaneous Sensing of Seven Pathogenic Bacteria by Guanidine-Functionalized Upconversion Fluorescent Nanoparticles vol.4, pp.5, 2019, https://doi.org/10.1021/acsomega.9b00775
  82. Exploring DNA quantity and quality from raw materials to botanical extracts vol.5, pp.6, 2019, https://doi.org/10.1016/j.heliyon.2019.e01935
  83. Preparation of Fluorescent Molecularly Imprinted Polymers via Pickering Emulsion Interfaces and the Application for Visual Sensing Analysis of Listeria Monocytogenes vol.11, pp.6, 2019, https://doi.org/10.3390/polym11060984
  84. Functionalized gold nanostructures: promising gene delivery vehicles in cancer treatment vol.9, pp.41, 2014, https://doi.org/10.1039/c9ra03608c
  85. Understanding and Exploiting Phage-Host Interactions vol.11, pp.6, 2014, https://doi.org/10.3390/v11060567
  86. Immunofluorescence Assay Using Monoclonal and Polyclonal Antibodies for Detection of Staphylococcal Enterotoxins A in Milk vol.13, pp.1, 2019, https://doi.org/10.2174/187407070190130137
  87. Principles of Hyperspectral Microscope Imaging Techniques and Their Applications in Food Quality and Safety Detection: A Review vol.18, pp.4, 2014, https://doi.org/10.1111/1541-4337.12432
  88. Developing a multiplex real-time PCR with a new pre-enrichment to simultaneously detect four foodborne bacteria in milk vol.14, pp.10, 2019, https://doi.org/10.2217/fmb-2019-0044
  89. Development of a Rapid Test Method for Salmonella enterica Detection Based on Fluorescence Probe-Based Recombinase Polymerase Amplification vol.12, pp.8, 2014, https://doi.org/10.1007/s12161-019-01526-3
  90. Contrast of Real-Time Fluorescent PCR Methods for Detection of Escherichia coli O157:H7 and of Introducing an Internal Amplification Control vol.7, pp.8, 2014, https://doi.org/10.3390/microorganisms7080230
  91. Ultrafast, low-power, PCB manufacturable, continuous-flow microdevice for DNA amplification vol.411, pp.20, 2014, https://doi.org/10.1007/s00216-019-01911-1
  92. Microfluidic-Based Approaches for Foodborne Pathogen Detection vol.7, pp.10, 2014, https://doi.org/10.3390/microorganisms7100381
  93. Virulence genes in Escherichia coli isolates from commercialized saltwater mussels Mytella guyanensis (Lamarck, 1819) vol.79, pp.4, 2014, https://doi.org/10.1590/1519-6984.185930
  94. Carbon Nanomaterial-Based Electrochemical Biosensors for Foodborne Bacterial Detection vol.49, pp.6, 2014, https://doi.org/10.1080/10408347.2018.1561243
  95. A microfluidic immunosensor for visual detection of foodborne bacteria using immunomagnetic separation, enzymatic catalysis and distance indication vol.186, pp.12, 2019, https://doi.org/10.1007/s00604-019-3883-x
  96. RNA detection with high specificity and sensitivity using nested fluorogenic Mango NASBA vol.25, pp.12, 2014, https://doi.org/10.1261/rna.072629.119
  97. Transcriptomic Analysis of Viable but Non-Culturable Escherichia coli O157:H7 Formation Induced by Low Temperature vol.7, pp.12, 2014, https://doi.org/10.3390/microorganisms7120634
  98. Black phosphorus-Au filter paper-based three-dimensional SERS substrate for rapid detection of foodborne bacteria vol.497, pp.None, 2014, https://doi.org/10.1016/j.apsusc.2019.143825
  99. A multiplex loop-mediated isothermal amplification assay for rapid detection of Bacillus cereus and Staphylococcus aureus vol.13, pp.6, 2014, https://doi.org/10.5582/bst.2019.01267
  100. Construction of a Chimeric Plasmid Vector and Its Investigation for Usage as an Internal Control for Detection of Shiga Toxin-Producing Escherichia coli by Polymerase Chain Reaction and Real-Time PCR vol.13, pp.2, 2014, https://doi.org/10.5812/jjm.98035
  101. Review on Major Food-Borne Zoonotic Bacterial Pathogens vol.2020, pp.None, 2020, https://doi.org/10.1155/2020/4674235
  102. An in situ -Synthesized Gene Chip for the Detection of Food-Borne Pathogens on Fresh-Cut Cantaloupe and Lettuce vol.10, pp.None, 2020, https://doi.org/10.3389/fmicb.2019.03089
  103. Comparison of Antimicrobial Resistance Detected in Environmental and Clinical Isolates from Historical Data for the US vol.2020, pp.None, 2014, https://doi.org/10.1155/2020/4254530
  104. A Recombinase Polymerase Amplification and Lateral Flow Strip Combined Method That Detects Salmonella enterica Serotype Typhimurium With No Worry of Primer-Dependent Artifacts vol.11, pp.None, 2020, https://doi.org/10.3389/fmicb.2020.01015
  105. Formation and Control of the Viable but Non-culturable State of Foodborne Pathogen Escherichia coli O157:H7 vol.11, pp.None, 2020, https://doi.org/10.3389/fmicb.2020.01202
  106. Loop-mediated isothermal amplification-based microfluidic chip for pathogen detection vol.60, pp.2, 2020, https://doi.org/10.1080/10408398.2018.1518897
  107. Two rapid and sensitive methods based on TaqMan qPCR and droplet digital PCR assay for quantitative detection of Bacillus subtilis in rhizosphere vol.128, pp.2, 2014, https://doi.org/10.1111/jam.14481
  108. Predominant Mycotoxins, Pathogenesis, Control Measures, and Detection Methods in Fermented Pastes vol.12, pp.2, 2014, https://doi.org/10.3390/toxins12020078
  109. Immuno- and nucleic acid-based current technique for Salmonella detection in food vol.246, pp.3, 2020, https://doi.org/10.1007/s00217-019-03423-9
  110. Microbiological safety of ready‐to‐eat foods in low‐ and middle‐income countries: A comprehensive 10‐year (2009 to 2018) review vol.19, pp.2, 2014, https://doi.org/10.1111/1541-4337.12533
  111. Bacteriophage Based Biosensors: Trends, Outcomes and Challenges vol.10, pp.3, 2014, https://doi.org/10.3390/nano10030501
  112. Evolving techniques for the detection ofListeria monocytogenes: underlining the electrochemical approach vol.100, pp.4, 2014, https://doi.org/10.1080/03067319.2019.1674502
  113. Rapid screening and quantitative detection of Salmonella using a quantum dot nanobead-based biosensor vol.145, pp.6, 2014, https://doi.org/10.1039/d0an00035c
  114. High-flux simultaneous screening of common foodborne pathogens and their virulent factors vol.43, pp.4, 2020, https://doi.org/10.1007/s00449-019-02267-7
  115. Chitosan Stabilized Silver Nanoparticles for the Electrochemical Detection of Lipopolysaccharide: A Facile Biosensing Approach for Gram-Negative Bacteria vol.11, pp.4, 2014, https://doi.org/10.3390/mi11040413
  116. Methods for detection of viable foodborne pathogens: current state-of-art and future prospects vol.104, pp.10, 2014, https://doi.org/10.1007/s00253-020-10542-x
  117. Design 5.0 µm Gap Aluminium Interdigitated Electrode for Sensitive pH Detection vol.864, pp.None, 2014, https://doi.org/10.1088/1757-899x/864/1/012178
  118. Physicochemical properties and mode of action of a novel bacteriocin BM1122 with broad antibacterial spectrum produced by Lactobacillus crustorum MN047 vol.85, pp.5, 2014, https://doi.org/10.1111/1750-3841.15131
  119. Translating ‘big data’: better understanding of host-pathogen interactions to control bacterial foodborne pathogens in poultry vol.21, pp.1, 2014, https://doi.org/10.1017/s1466252319000124
  120. Advances in Biomimetic Systems for Molecular Recognition and Biosensing vol.5, pp.2, 2020, https://doi.org/10.3390/biomimetics5020020
  121. Detection of toxins involved in foodborne diseases caused by Gram‐positive bacteria vol.19, pp.4, 2014, https://doi.org/10.1111/1541-4337.12571
  122. From hazard analysis to risk control using rapid methods in microbiology: A practical approach for the food industry vol.19, pp.4, 2020, https://doi.org/10.1111/1541-4337.12592
  123. Dielectrophoresis-based microwire biosensor for rapid detection of Escherichia coli K-12 in ground beef vol.132, pp.None, 2020, https://doi.org/10.1016/j.lwt.2020.109230
  124. How to rapidly and sensitively detect for Escherichia coli O157:H7 and Salmonella Typhimurium in cabbage using filtration, DNA concentration, and real-time PCR after short-term enrichment vol.132, pp.None, 2014, https://doi.org/10.1016/j.lwt.2020.109840
  125. Biosensors in Evaluation of Quality of Meat and Meat Products - A Review vol.20, pp.4, 2014, https://doi.org/10.2478/aoas-2020-0057
  126. Electrochemical Immuno- and Aptamer-Based Assays for Bacteria: Pros and Cons over Traditional Detection Schemes vol.20, pp.19, 2020, https://doi.org/10.3390/s20195561
  127. Planar Interdigitated Aptasensor for Flow-Through Detection of Listeria spp. in Hydroponic Lettuce Growth Media vol.20, pp.20, 2020, https://doi.org/10.3390/s20205773
  128. Microbial detection and identification methods: Bench top assays to omics approaches vol.19, pp.6, 2014, https://doi.org/10.1111/1541-4337.12618
  129. Rapid Detection of Diarrheagenic Escherichia coli by a New Multiplex Real-Time Quantitative PCR Assay vol.56, pp.6, 2014, https://doi.org/10.1134/s0003683820060174
  130. Direct detection of Salmonella from poultry samples by DNA isothermal amplification vol.61, pp.6, 2014, https://doi.org/10.1080/00071668.2020.1808188
  131. Discrimination between pathogenic and non-pathogenic E. coli strains by means of Raman microspectroscopy vol.412, pp.30, 2014, https://doi.org/10.1007/s00216-020-02957-2
  132. Rapid quantification of ESCHERICHIA COLI O157 : H7 in lettuce and beef using an on‐chip staining microfluidic device vol.40, pp.6, 2020, https://doi.org/10.1111/jfs.12851
  133. Development and validation of TOF/Q-TOF MS/MS, HPLC method and in vitro bio-strategy for aflatoxin mitigation vol.37, pp.12, 2020, https://doi.org/10.1080/19440049.2020.1815861
  134. Tutorial: design and fabrication of nanoparticle-based lateral-flow immunoassays vol.15, pp.12, 2014, https://doi.org/10.1038/s41596-020-0357-x
  135. Risk factors related to bacterial contamination by Enterobacteriaceae and fecal coliforms and the prevalence of Salmonella spp. in Algerian farms, slaughterhouses and butcheries: a two-year follow-up vol.6, pp.3, 2014, https://doi.org/10.3934/agrfood.2021046
  136. Development of Fluorescence Imaging Technique to Detect Fresh-Cut Food Organic Residue on Processing Equipment Surface vol.11, pp.1, 2014, https://doi.org/10.3390/app11010458
  137. Recent Advances in Molecular Diagnostics of Fungal Plant Pathogens: A Mini Review vol.10, pp.None, 2014, https://doi.org/10.3389/fcimb.2020.600234
  138. Advances in typing and identification of foodborne pathogens vol.37, pp.None, 2021, https://doi.org/10.1016/j.cofs.2020.09.002
  139. Point-of-Need Diagnostics for Foodborne Pathogen Screening vol.26, pp.1, 2014, https://doi.org/10.1177/2472630320962003
  140. An Engineered Reporter Phage for the Fluorometric Detection of Escherichia coli in Ground Beef vol.9, pp.2, 2014, https://doi.org/10.3390/microorganisms9020436
  141. Assessment of human norovirus inhibition in cabbage kimchi by electron beam irradiation using RT‐qPCR combined with immunomagnetic separation vol.86, pp.2, 2021, https://doi.org/10.1111/1750-3841.15562
  142. Ultrasensitive label-free immunochromatographic strip sensor for Salmonella determination based on salt-induced aggregated gold nanoparticles vol.343, pp.None, 2014, https://doi.org/10.1016/j.foodchem.2020.128518
  143. Application of bacteriophage in rapid detection of Escherichia coli in foods vol.39, pp.None, 2021, https://doi.org/10.1016/j.cofs.2020.12.015
  144. Multicolor Coding Up-Conversion Nanoplatform for Rapid Screening of Multiple Foodborne Pathogens vol.13, pp.23, 2014, https://doi.org/10.1021/acsami.1c05522
  145. Advances in differentiation and identification of foodborne bacteria using near infrared spectroscopy vol.13, pp.23, 2021, https://doi.org/10.1039/d1ay00124h
  146. Advance methods for the qualitative and quantitative determination of microorganisms vol.166, pp.None, 2014, https://doi.org/10.1016/j.microc.2021.106188
  147. Magnetic Separation and Centri-Chronoamperometric Detection of Foodborne Bacteria Using Antibiotic-Coated Metallic Nanoparticles vol.11, pp.7, 2014, https://doi.org/10.3390/bios11070205
  148. Isothermal amplification assay for visual on-site detection of Staphylococcus aureus in Chevon vol.35, pp.3, 2021, https://doi.org/10.1080/08905436.2021.1941078
  149. Changes of Viscoelastic Properties of Aptamer-Based Sensing Layers Following Interaction with Listeria innocua vol.21, pp.16, 2014, https://doi.org/10.3390/s21165585
  150. Enzyme-Free Multiplex Detection of Foodborne Pathogens Using Au Nanoparticles-Decorated Multiwalled Carbon Nanotubes vol.1, pp.7, 2021, https://doi.org/10.1021/acsfoodscitech.1c00124
  151. Sero‐prevalence and risk factors of Brucella presence in farm bulk milk from open and zero grazing cattle production systems in Rwanda vol.7, pp.5, 2021, https://doi.org/10.1002/vms3.562
  152. Rapid and Simultaneous Detection of Five, Viable, Foodborne Pathogenic Bacteria by Photoinduced PMAxx-Coupled Multiplex PCR in Fresh Juice vol.18, pp.9, 2014, https://doi.org/10.1089/fpd.2020.2909
  153. Snipe: highly sensitive pathogen detection from metagenomic sequencing data vol.22, pp.5, 2014, https://doi.org/10.1093/bib/bbab064
  154. Bio-Specific Au/Fe3+ Porous Spongy Nanoclusters for Sensitive SERS Detection of Escherichia coli O157:H7 vol.11, pp.10, 2014, https://doi.org/10.3390/bios11100354
  155. Overview of Rapid Detection Methods for Salmonella in Foods: Progress and Challenges vol.10, pp.10, 2014, https://doi.org/10.3390/foods10102402
  156. Label-Free Colorimetric Method for Detection of Vibrio parahaemolyticus by Trimming the G-Quadruplex DNAzyme with CRISPR/Cas12a vol.93, pp.42, 2014, https://doi.org/10.1021/acs.analchem.1c03468
  157. Invited review: Stress resistance of Cronobacter spp. affecting control of its growth during food production vol.104, pp.11, 2014, https://doi.org/10.3168/jds.2021-20591
  158. A portable CRISPR Cas12a based lateral flow platform for sensitive detection of Staphylococcus aureus with double insurance vol.132, pp.None, 2014, https://doi.org/10.1016/j.foodcont.2021.108485
  159. Development of modified enrichment broth for short enrichment and recovery of filter-injured Salmonella Typhimurium vol.362, pp.None, 2022, https://doi.org/10.1016/j.ijfoodmicro.2021.109497
  160. Biomimetic dandelion-like magnetic nanoparticles for capture and detection of S. aureus and L. monocytogenes vol.355, pp.None, 2022, https://doi.org/10.1016/j.snb.2021.131289
  161. Biosensors for simplistic detection of pathogenic bacteria: A review with special focus on field-effect transistors vol.141, pp.None, 2014, https://doi.org/10.1016/j.mssp.2021.106404