Open access peer-reviewed chapter

Current and Emerging Technologies for Rapid Detection of Pathogens

Written By

Lingwen Zeng, Lihua Wang and Jiao Hu

Submitted: 30 May 2017 Reviewed: 15 December 2017 Published: 21 March 2018

DOI: 10.5772/intechopen.73178

From the Edited Volume

Biosensing Technologies for the Detection of Pathogens - A Prospective Way for Rapid Analysis

Edited by Toonika Rinken and Kairi Kivirand

Chapter metrics overview

2,620 Chapter Downloads

View Full Metrics

Abstract

Foodborne diseases, caused by pathogenic bacteria, have become an important social issue in the field of food safety. It presents a widespread and growing threat to human health in both developed and developing countries. As such, techniques for the detection of foodborne pathogens and waterborne pathogens are urgently needed to prevent the occurrence of human foodborne infections. Although traditional culture-based bacterial isolation and identification are the “gold standard” methods with high preciseness, their drawbacks in time-consuming are inadequate for rapid detection of pathogen to reduce foodborne disease occurrence. Fortunately, with the development of biotechnologies and nanotechnologies, various kinds of new technologies for rapid detection of pathogens have been developed so far, such as nucleic acid-based methods, antibody-based methods, and aptamer-based assays. In this chapter, we summarized the principles and the application of some recent rapid detection technologies for pathogenic bacteria. Moreover, the advantages and disadvantages of the established and emerging rapid detection methods are addressed here.

Keywords

  • pathogen
  • rapid detection
  • nucleic acid
  • antibody
  • aptamer

1. Introduction

Foodborne pathogens, which are widely responsible for many foodborne diseases, constitute a serious threat to human health. In recent years, foodborne and waterborne pathogenic microorganisms have caused numerous epidemic diseases in the world [1]. Salmonella, Shiga bacillus, Escherichia coli O157:H7, Bacillus cereus, Staphylococcus aureus, and Listeria monocytogenes are the primary pathogens that are responsible for most foodborne disease [2, 3, 4, 5, 6]. Centers for Disease Control and Prevention (CDC) reported that approximately 73,000 cases of foodborne disease occur annually. In 2013, a total of 19,056 are infected with foodborne pathogen, of which 4200 are hospitalizations, and 80 are deaths in the United States [7]. The foodborne diseases are even high prevalence in many developing countries. Worldwide, there are 600 million foodborne illnesses with 420,000 deaths in 2010, which is estimated by the World Health Organization (WHO) Foodborne Disease Burden Epidemiology Reference Group (FERG). A great proportion of these cases are due to the contamination of raw or undercooked foods and drinking water [4, 6, 8, 9]. Hence, it is urgent to detect foodborne pathogens in order to control foodborne pathogen spread and reduce foodborne disease occurrence.

Currently, culture-based bacterial isolation and identification are the “gold standard” methods for laboratory detection of foodborne pathogens [10]. However, they suffer from time consumption, which requires 2–3 days for initial culture and enrichment, and more than 1 week for confirming the target pathogenic bacteria [11, 12]. Moreover, it requires expensive instruments and professional technicians and remains problematic due to the lack of phenotypic characteristics to distinguish between generic pathogens, which may largely restrict its application. It is evident that culture and colony-counting methods are inadequate for rapid detection of foodborne pathogens, especially for reduce foodborne disease occurrence. The frequent outbreak of foodborne diseases and the economic and social implications indicate that analytical methodologies that can rapidly detect and identify pathogens are urgently needed. As such, many researchers devote themselves to developing more advanced detection methods that can identify pathogens accurately and rapidly in a timely manner in the food industry [13, 14, 15, 16, 17, 18, 19, 20].

In this chapter, we summarize the recent trends, developments, advantages, and disadvantages (listed in Table 1) about rapid detection of pathogens based on nucleic acid, antibodies, and aptamers and then give a perspective on the future directions of rapid analysis of pathogens.

Method Advantages Disadvantages Sensitivity Ref.
Real-time PCRa —Amplification can be monitored at real time
—Confirmation of specific amplification by melting curve
—Accurate quantification
—Difficulty in multiplex assay
—Need skilled person and support
—False-positive results
10 CFU/mL [25]
Multiplex PCRa —Highly efficient (detection of several pathogens at a time)
—Systematic (suitable for detection of groups of pathogens)
—Difficulty in distinguishing live and dead cells
—Requires post-PCR processing of products (electrophoresis)
—Need skilled person and support
—Costs more than culture-based methods and ELISA
1 CFU/mL [26]
Antibody-based method (ELISAb and LFIAc) —More rapid than culture-based methods (1~2 h vs. 5~7 d)
—Can be automated to reduce assay time and manual labor input
—Able to handle large numbers of samples
—Convenient and suitable for the on-site testing
—Difficulty to differentiate damaged or stressed cells
—Need for pre-enrichment
—High cross-reactivity with close antigens in bacteria
60 CFU/mL [35]
Aptamer-based method (optical and electrochemical methods) —Inexpensive, stable, and can be chemically synthesized than antibody
—Time-saving (2 h vs. 5~7 d of culture-based methods)
—Automated to reduce manual labor input
—High throughput
—Multiplex assays
—High false positive
—Difficulty in detecting damaged or stressed cells
—Need for pre-enrichment
—Possibility of cross contamination
1.5 CFU/mL [78]

Table 1.

Advantages and disadvantages of detection methods.

a PCR, polymerase chain reaction; b ELISA, enzyme-linked immunosorbent assay; c LFIA, lateral flow immunoassays.

Advertisement

2. Methodologies for pathogen detection

2.1. Nucleic acid-based assays

Culture- and colony-based methods are the standard methods for the detection of pathogens. They rely on the ability of microorganisms to multiply to visible colonies [21]. The major drawbacks of these microbiological methods are their labor intensiveness and time consumption as it usually takes 2–3 days for initial results and up to 7–10 days for confirmation. In comparison, nucleic acid-based assays can greatly shorten the testing time.

2.1.1. Real-time PCR

Real-time PCR technology is a reliable method in identification and quantitative detection of bacteria due to its accuracy, rapidity, specificity, and low detection limit. In addition, it is a promising alternative approach to estimating the number of bacteria [22, 23]. For example, Gyawali et al. [22] presented a specific and sensitive real-time PCR method to detect Ancylostoma caninum ova in wastewater matrices. This method exhibited high sensitivity with the ability to detect Ancylostoma caninum DNA up to dilution of 10−4 (equivalent to 500 fg) consistently. Moreover, the precise copy number of a specific nucleic acid sequence can be quantified with the real-time PCR technique based on a calibration curve created with known concentrations of DNA [24, 25]. Gokduman et al. [25] established a recombinant plasmid-based quantitative real-time PCR assay for Salmonella enterica serotypes with the detection limit of 10 CFU/mL. Obviously, the real-time quantitative PCR has already been a promising quantitative method for the quantitative detection of bacteria, due to its lower cost than that of culture-based method.

2.1.2. Multiplex PCR

Multiplex PCR, also known as multiple primer PCR, which is a PCR reaction system with two or more primers, can amplify a plurality of nucleic acid fragments in a system. Compared to other methods, multiplex PCR is very useful as it allows the simultaneous detection of several pathogenic bacteria by introducing different primers to amplify DNA regions coding for specific genes of each bacterial strain targeted [26]. Methods for multiplexing PCR have considerably improved over the last years, thereby decreasing genotyping costs and increasing throughput. Examples of multiplex PCR technique for the simultaneous detection pathogens include multiplex PCR assay for rapid and simultaneous detection of Escherichia coli O157:H7, Salmonella, Staphylococcus aureus, Listeria monocytogenes, and Vibrio parahaemolyticus [27]; simultaneous detection of bacteria of the genus Listeria, Listeria monocytogenes, and major serotypes and epidemic clones of Listeria monocytogenes [28]; and simultaneous detection of Escherichia coli O157:H7 and Listeria monocytogenes [29]. Multiplex real-time PCR-based assay can rapidly detect 25 clinically important pathogens directly from whole blood in <6 h [30].

2.2. Antibody-based assays

Antibodies are a unique natural family of immune system-related glycoproteins known as immunoglobulins, produced by differentiated B cells in response to the attendant of an immunogen during an immune response. Because of the specific interactions and the extremely high equilibrium association constants (1010/M and greater) attainable between an antibody and its corresponding antigen, antibodies are employed as an excellent biorecognition element for the highly sensitive and selective immunoassays [31]. Their utilization in biosensors brings new tools for analysis in the biochemical, clinical, and environmental fields. Without exception, antibody-based assays such as enzyme-linked immunosorbent assay (ELISA), lateral flow immunoassay (LFIA), and so on are very popular for the detection of pathogens.

2.2.1. Enzyme-linked immunosorbent assay (ELISA)

ELISA-based approaches are the most prevalent antibody-based assay for pathogen detection [32]. Compared with the culture-based methods, this immunological approach has been used to detect pathogens in poultry production (poultry feed, feces, litter, carcass rinsing, and water samples) and has provided a better sensitivity and shorter time frame [33]. Recently, improvements by combination with other advanced nanomaterials such as novel enzyme-based signal probes have been made in the basic ELISA method for pathogen detection. For example, by using silica nanoparticles (NPs) carrying poly(acrylic acid) brushes as a “catalase (CAT) container” to increase enzyme loading, Chen et al. [34] presented an improved plasmonic ELISA (pELISA) method for detection of Listeria monocytogenes at ultralow concentrations with the sandwich format. The limit of detection (LOD) obtained by this method (80 CFU/mL) was two and five orders of magnitude lower than that of conventional CAT-based pELISA and horseradish peroxidase (HRP)-based conventional ELISA, respectively. To further simplify the preparation of enzyme-labeled antibody, Lin’s group innovatively prepared an all-in-one organic–inorganic nanoflower, which integrated biorecognition unit (concanavalin A or antibody), signal amplification unit (glucose oxidase or HRP), and carrier unit within a one-pot reaction. And then, it was used for a portable sensitive ELISA detection of Escherichia coli O157:H7. Under the optimal conditions, the detection sensitivity can reach as low as 10 CFU/mL for the case of concanavalin A-glucose oxidase [17] and 60 CFU/mL for the case of antibody-HRP [35].

2.2.2. Lateral flow immunoassay (LFIA)

LFIA-based methods are a form of immunoassay, which emerge for the first time at the end of the 1960s and consist of a chromatographic system and immunochemical reaction [36, 37, 38]. The principle of LFIA is based on antibody–antigen specific interaction. After the sample is applied to the sample pad, it migrates along the test strip via capillary action, and a signal response is obtained about 5–10 min later [39, 40]. Due to its simplicity, rapidity, low cost, portability, and facile interpretation without external reagent or external instrumentation, LFIA has held great potential for foodborne pathogen detection [15, 16, 41]. In addition, the LFIA can realize visual detection and quantitative detection by employing different labels, such as colloid gold, fluorescent materials, and magnetic beads [40, 42, 43]. Descriptions on some of the labels that are applied for pathogenic bacteria detection are presented in the following sections.

2.2.2.1. Colloid gold as label

Colloid gold is the most widely used label of LFIA due to its intense color and direct visualization [44], and it has been widely used for the detection of foodborne pathogens [45, 46, 47, 48]. Jung et al. [45] used a colloid gold-based LFIA to detect Escherichia coli O157:H7 in enriched samples, and the LOD was 1.8 × 105 CFU/mL without enrichment and 1.8 CFU/mL after enrichment. Preechakasedkit et al. [49] also developed a colloid gold immunochromatographic strip for the detection of Salmonella typhimurium with a minimum detection limit of 1.14 × 105 CFU/mL. Park et al. [15] presented a detection method of Escherichia coli O157:H7 and Salmonella typhimurium with a pressed paper-based dipstick by employing colloid gold as labels. The detection limit of Escherichia coli O157:H7 was around 105 CFU/mL, while that of Salmonella typhimurium was around 106 CFU/mL. In the case of the work of Song et al. [41], Shigella boydii, and Escherichia coli O157:H7 can be detected simultaneously in bread, milk, and jelly samples using colloid gold-based LFIA, and the detection limit of 106 CFU/mL for both Shigella boydii and Escherichia coli O157:H7 was achieved.

2.2.2.2. Quantum dots as label

As the low sensitivity of colloid gold, fluorescent materials have gained more and more interest due to their higher sensitivity than colloid gold in the field of lateral flow assay [50]. Furthermore, the fluorescent materials enable lateral flow assay to detect the target quantitatively. Compared with colloid gold, which can only provide qualitative or semiquantitative results, quantitative detection can offer more information [42, 51, 52]. In particular, quantum dots show unique fluorescence properties, such as high and stable fluorescence signal [53, 54, 55]. During the last decade, quantum dot-based lateral flow assays have been applied to the detection of foodborne pathogen [56, 57, 58]. Bruno [56] utilized quantum dot-conjugated antibody as the signal reporter of the lateral flow assay to detect Escherichia coli O157:H7. With the assay, the detection limit of Escherichia coli O157:H7 is calculated to be 600 cells per test, while that of colloid gold-based lateral flow assay is 6000 cells per test, indicating the higher sensitivity of quantum dots than colloid gold as labels of lateral flow assay. Chen et al. [58] also developed a competitive format lateral flow assay with quantum dots for the detection of Staphylococcus aureus in food. The detection limit is 3 CFU/mL.

2.2.2.3. Magnetic beads as label

Magnetic beads are another type of label, which can realize quantitative detection of targets by measuring the magnetic signal [40, 57, 59]. Due to the fact that they are strongly colored and can enrich and separate targets from complex matrix, magnetic beads are new attractive materials to construct a lateral flow assay, which will probably replace traditional labels. Especially, magnetic beads can simultaneously provide visual signal and magnetic signal. Several researches have recently focused on the use of magnetic bead-based lateral flow assay to detect pathogenic bacteria [60, 61, 62]. Wang et al. [60] employed antibody-coated magnetic beads with the diameter of 300 nm as signal reporter of lateral flow assay for Bacillus anthracis spore detection. A detection limit of 6 × 104 spores/g of milk powder, 2 × 105 spores/g of starch, and 5 × 105 spores/g of baking soda was obtained, respectively. Suaifan et al. [63] described a magnetic bead-based lateral flow assay, which can specifically and simultaneously detect Escherichia coli O157:H7 proteases in complex food matrices. The limits of detection were 12 CFU/mL in broth and 30–300 CFU/mL in food matrices. Xia et al. [64] developed a gold magnetic bifunctional nanobead-based lateral flow assay for the detection of Salmonella choleraesuis. Results indicated that the assay was specific and rapid with the detection limit of 5 × 105 CFU/mL, which was much more sensitive than that of colloid gold-based LFIA (5 × 106 CFU/mL), suggesting that magnetic beads were indeed superior to colloid gold.

2.3. Aptamer-based assays

Besides antibodies, other biomolecules have been investigated to selectively capture and enrich pathogens from cultures, among which aptamer is the most prevalent one [65]. Aptamers, as short single-stranded nucleic acids (DNA or RNA), can bind with high affinity and specificity to a wide range of target molecules, such as ions, small organic molecules, and proteins [66, 67, 68]. The affinities of aptamers for their targets are comparable to, or even higher than most monoclonal antibodies. More importantly, compared with antibodies, they also exhibit a number of advantages. First of all, aptamers can be routinely produced by chemical synthesis, avoiding the use of animals required for antibody production. Furthermore, they are generally more chemically stable, and their binding properties are easier to manipulate. To this end, a number of aptasensors based on optics and electrochemistry have been recently reported for pathogenic microorganism typing and detection.

2.3.1. Optical strategies

Surface-enhanced Raman scattering (SERS) possesses several attractive properties, such as ultrahigh sensitivity, high speed, comparatively low cost, and multiplexing ability and portability [69, 70, 71], which enable SERS to be widely used for sensitive detection of chemical and biological agents [72, 73]. Since Holt and Cotton first reported the SERS spectrum of bacteria, the identification and detection of microorganism by SERS have attracted high interest recently due to the spectroscopic fingerprint and nondestructive data acquisition in aqueous environment [74]. To date, there have been many SERS biosensors developed, especially based on a “magnetic separation” approach, which focus on bacterial pathogen detection. Wang et al. [75] reported a magnetically assisted SERS biosensor for single-cell detection of Staphylococcus aureus on the basis of aptamer recognition. The biosensor consists of two basic elements including a SERS substrate (Ag-coated magnetic nanoparticles) and a novel SERS tag (Au nanorod-5,5-dithiobis-(2-nitrobenzoic acid) (AuNR-DTNB)@Ag-DTNB core-shell plasmonic NPs or DTNB-labeled inside-and-outside plasmonic NPs (DioPNPs)). Based on these, the LOD of 10 cells/mL can be achieved for Staphylococcus aureus detection. Similarly, through combined gold NPs (GNPs) modified with Raman molecules and Fe3O4 magnetic GNPs immobilized with aptamer, Zhang et al. [76] successfully fabricated GNP-enhanced SERS aptasensor for the simultaneous detection of Salmonella typhimurium and Staphylococcus aureus. In comparison with these label-based SERS methods, label-free methods do not require a secondary label dye and can directly obtain the intrinsic fingerprint of bacteria, which relies on the mutual interaction of bacteria cell with the SERS substrate [77]. With this regard, a lot of label-free methods have been developed for the detection of pathogens. For example, Gao’s group [78] successfully achieved intuitive label-free SERS detection of bacteria using aptamer-based in situ Ag NP synthesis. The biosensor as prepared can recognize bacteria quickly and directly by SERS with the formation of well-defined bacteria-aptamer@Ag NPs. The detection limit is down to 1.5 CFU/mL.

As another typical spectroscopic method, fluorescence resonance energy transfer (FRET, a homogeneous signal transduction technique), has been gradually employed for the determination of pathogenic bacteria. Yu et al. [79] presented a universal and facile one-step strategy for sensitive and selective detection of pathogenic bacteria using a dual-molecular affinity-based FRET platform based on the recognition of bacterial cell walls by antibiotic and aptamer molecules, respectively. Within 30 min, the FRET signal shows a linear variation with the concentration of Staphylococcus aureus in the range from 20 to 108 CFU/mL with a detection limit of 10 CFU/mL. Moreover, Duan’s group [80] further achieved simultaneous detection of Vibrio parahaemolyticus, Staphylococcus aureus, and Salmonella typhimurium through using multicolor dyes as donors and carbon NPs as a sole acceptor in FRET.

2.3.2. Electrochemical strategies

Compared with optical-based biosensors, electrochemical methods, in general, show the potential for construction of fast, simple, low-cost, sensitive, and high-throughput biosensors that can be miniaturized [81, 82, 83, 84]. To date, electrochemical aptasensors are widely used for identification and quantification of pathogens. For example, Labib et al. [85] developed an impedimetric sensor via assembling their selected highly specific DNA aptamers onto a gold NP-modified screen-printed carbon electrode for the highly sensitive detection of live Salmonella typhimurium. This aptasensor is very simple and highly selective. It can successfully detect Salmonella typhimurium down to 600 CFU/mL (equivalent to 18 live cells in 30 μL of assay volume). Moreover, to further improve the sensitivity, Abbaspour et al. [86] innovatively combined the magnetic beads’ fast separation with the Ag NPs’ signal amplification. They successfully fabricated an electrochemical dual-aptamer-based sandwich detection method for Staphylococcus aureus. The aptasensor as prepared shows an extended dynamic range from 10 to 1 × 106 CFU/mL with a low detection limit of 1.0 CFU/mL (S/N = 3). Despite much progress has been made, these methods always require probe labeling and aptamer immobilization, which may affect the binding affinities between bacteria and their aptamers. With this respect, Ding’s group [87] constructed a label-free potentiometric aptasensor for rapid, sensitive, and selective detection of Listeria monocytogenes. In this strategy, the target-binding event prevents the aptamer from electrostatically interacting with protamine, which can be sensitively detected using a polycation-sensitive membrane electrode.

2.4. Conclusion

Culture-based foodborne pathogen detection methods, although sensitive enough, are often too time-consuming to reduce foodborne disease occurrence. Therefore, a large number of innovative methods have been developed to overcome this performance limitation. These rapid detection methods can be classified into nucleic acid-based methods, antibody-based methods, and aptamer-based methods. All these rapid methods for foodborne pathogen detection are superior to culture-based methods. However, some of them still require improvement in sensitivity, selectivity, simplicity, or accuracy to be of any practical use. Nucleic acid-based methods, as a replacement method for culture-based methods, have high sensitivity and require a shorter time than conventional culture-based techniques for foodborne pathogen detection. Most of them still require highly trained personnel and expensive instruments, which limit their use in a practical environment. The development of antibody-based methods helped improve the time required to yield results. The specific binding of antibody to its antigen results in its high specificity and sensitivity of antibody-based methods, and they work well in food matrices without being interfered by other DNAs, proteins, or nontarget cells. Aptamer-based methods are similar to antibody-based methods, which also exhibit high sensitivity and selectivity. However, they still need to be improved for food matrix detection. Increasing detection accuracy and decreasing detection time are the eternal themes in rapid detection. In the future, new nanomaterials and rational biosensing strategies would be developed to approach the goal.

References

  1. 1. Tian F, Lyu J, Shi JY, Tan F, Yang M. A polymeric microfluidic device integrated with nanoporous alumina membranes for simultaneous detection of multiple foodborne pathogens. Sensors & Actuators, B: Chemical. 2016;225:312-318
  2. 2. Gould LH, Mungai EA, Johnson SD, Richardson LC, Williams IT, Griffin PM, Cole DJ, Hall AJ. Surveillance for foodborne disease outbreaks-United States, 2009–2010. Morbidity and Mortality Weekly Report. 2013;62
  3. 3. Gould LH, Nisler AL, Herman KM, Cole DJ, Williams IT, Mahon BE, Griffin PM, Hall AJ. Surveillance for foodborne disease outbreaks-United States, 2008. JAMA, Journal of the American Medical Association. 2011;306:2212-2214
  4. 4. Law JWF, Mutalib NSA, Chan KG, Lee LH. Rapid methods for the detection of foodborne bacterial pathogens: Principles, applications, advantages and limitations. Frontiers in Microbiology. 2015;5
  5. 5. Chao GX, Zhou XH, Jiao XN, Qian XQ, Xu L. Prevalence and antimicrobial resistance of foodborne pathogens isolated from food products in China. Foodborne Pathogens and Disease. 2007;4:277-284
  6. 6. Henao-Herreño LX, López-Tamayo AM, Ramos-Bonilla JP, Haas CN, Husserl J. Risk of illness with Salmonella due to consumption of raw unwashed vegetables irrigated with water from the Bogota River. Risk Analysis. 2017;37:733-743
  7. 7. Crim SM, Iwamoto M, Huang JY, Griffin PM, Gilliss D, Cronquist AB, Cartter M, Tobin-D'Angelo M, Blythe D, Smith K, et al. Incidence and trends of infection with pathogens transmitted commonly through food-Foodborne Diseases Active Surveillance Network, 10 US Sites, 2006–2013. Morbidity and Mortality Weekly Report. 2014;63:328-332
  8. 8. Vinothkumar K, Bhardwaj AK, Ramamurthy T, Niyogi SK. Triplex PCR assay for the rapid identification of 3 major Vibrio species, Vibrio cholerae, Vibrio parahaemolyticus, and Vibrio fluvialis. Diagnostic Microbiology and Infectious Disease. 2013;76:526-528
  9. 9. Oliver SP, Jayarao BM, Almeida RA. Foodborne pathogens in milk and the dairy farm environment: Food safety and public health implications. Foodborne Pathogens and Disease. 2005;2:115-129
  10. 10. Andrews JR, Ryan ET. Diagnostics for invasive Salmonella infections: Current challenges and future directions. Vaccine. 2015;33:C8-C15
  11. 11. Zhao X, Lin CW, Wang J, Oh DH. Advances in rapid detection methods for foodborne pathogens. Journal of Microbiology and Biotechnology. 2014;24:297-312
  12. 12. Kawasaki S, Fratamico PM, Horikoshi N, Okada Y, Takeshita K, Sameshima T, Kawamoto S. 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 Pathogens and Disease. 2009;6:81-89
  13. 13. Velusamy V, Arshak K, Korostynska O, Oliwa K, Adley C. An overview of foodborne pathogen detection: In the perspective of biosensors. Biotechnology Advances. 2010;28:232-254
  14. 14. Alizadeh N, Memar MY, Moaddab SR, Kafil HS. Aptamer-assisted novel technologies for detecting bacterial pathogens. Biomedicine & Pharmacotherapy. 2017;93:737-745
  15. 15. Park J, Shin JH, Park JK. Pressed paper-based dipstick for detection of foodborne pathogens with multistep reactions. Analytical Chemistry. 2016;88:3781-3788
  16. 16. Shan S, Lai W, Xiong Y, Wei H, Xu H. Novel strategies to enhance lateral flow immunoassay sensitivity for detecting foodborne pathogens. Journal of Agricultural and Food Chemistry. 2015;63:745-753
  17. 17. Wei T, Du D, Zhu MJ, Lin Y, Dai Z. An improved ultrasensitive enzyme-linked immunosorbent assay using hydrangea-like antibody–enzyme–inorganic three-in-one nanocomposites. ACS Applied Materials & Interfaces. 2016;8:6329-6335
  18. 18. Mustafa F, Hassan R, Andreescu S. Multifunctional nanotechnology-enabled sensors for rapid capture and detection of pathogens. Sensors. 2017;17:2121
  19. 19. Park SH, Aydin M, Khatiwara A, Dolan MC, Gilmore DF, Bouldin JL, Ahn S, Ricke SC. Current and emerging technologies for rapid detection and characterization of Salmonella in poultry and poultry products. Food Microbiology. 2014;38:250-262
  20. 20. Wu W, Zeng L. Current and emerging innovations for detection of food-borne Salmonella. Current topics in Salmonella and Salmonellosis. Associate Prof. Mihai Mares (Ed.), INTECH. DOI: 10.5772/67264
  21. 21. De BE, Beumer RR. Methodology for detection and typing of foodborne microorganisms. International Journal of Food Microbiology. 1999;50:119-130
  22. 22. Gyawali P, Sidhu JPS, Ahmed W, Jagals P, Toze S. Rapid concentration and sensitive detection of Hookworm ova from wastewater matrices using a real-time PCR method. Experimental Parasitology. 2015;159:5-12
  23. 23. Mackay IM. Real-time PCR in the microbiology laboratory. Clinical Microbiology and Infection. 2004;10:190-212
  24. 24. Malorny B, Lofstrom C, Wagner M, Kramer N, Hoorfar J. Enumeration of Salmonella bacteria in food and feed samples by real-time PCR for quantitative microbial risk assessment. Applied and Environmental Microbiology. 2008;74:1299-1304
  25. 25. Gokduman K, Avsaroglu MD, Cakiris A, Ustek D, Gurakan GC. Recombinant plasmid-based quantitative real-time PCR analysis of Salmonella enterica serotypes and its application to milk samples. Journal of Microbiological Methods. 2016;122:50-58
  26. 26. Touron A, Berthe T, Pawlak B, Petit F. Detection of Salmonella in environmental water and sediment by a nested-multiplex polymerase chain reaction assay. Research in Microbiology. 2005;156:541-553
  27. 27. Kim JS, Lee GG, Park JS, Jung YH, Kwak HS, Kim SB, Nam YS, Kwon ST. A novel multiplex PCR assay for rapid and simultaneous detection of five pathogenic bacteria: Escherichia coli O157:H7, Salmonella, Staphylococcus aureus, Listeria monocytogenes, and Vibrio parahaemolyticus. Journal of Food Protection. 2007;70:1656-1662
  28. 28. Chen Y, Zhang W, Knabel SJ. Multi-virulence-locus sequence typing identifies single nucleotide polymorphisms which differentiate epidemic clones and outbreak strains of Listeria monocytogenes. Journal of Clinical Microbiology. 2007;45:835-846
  29. 29. Mukhopadhyay A, Mukhopadhyay UK. Novel multiplex PCR approaches for the simultaneous detection of human pathogens: Escherichia coli O157:H7 and Listeria monocytogenes. Journal of Microbiological Methods. 2007;68:193-200
  30. 30. Lehmann LE, Hunfeld KP, Emrich T, Haberhausen G, Wissing H, Hoeft A, Stuber F. A multiplex real-time PCR assay for rapid detection and differentiation of 25 bacterial and fungal pathogens from whole blood samples. Medical Microbiology and Immunology. 2008;197:313-324
  31. 31. Farka Z, Juriik T, Kovaar D, Trnkova L, Sklaadal P. Nanoparticle-based immunochemical biosensors and assays: Recent advances and challenges. Chemical Reviews. 2017;117:9973-10042
  32. 32. Mandal TK, Parvin N. Rapid detection of bacteria by carbon quantum dots. Journal of Biomedical Nanotechnology. 2011;7:846-848
  33. 33. Maciorowski KG, Herrera P, Jones FT, Pillai SD, Ricke SC. Cultural and immunological detection methods for Salmonella spp. in animal feeds–A review. Veterinary Research Communications. 2006;30:127-137
  34. 34. Chen R, Huang XL, Xu HY, Xiong YH, Li YB. Plasmonic enzyme-linked immunosorbent assay using nanospherical brushes as a catalase container for colorimetric detection of ultralow concentrations of Listeria monocytogenes. ACS Applied Materials & Interfaces. 2015;7:28632-28639
  35. 35. Ye R, Zhu C, Song Y, Lu Q, Ge X, Yang X, Zhu MJ, Du D, Li H, Lin Y. Bioinspired synthesis of all-in-one organic-inorganic hybrid nanoflowers combined with a hand-held pH meter for on-site detection of food pathogen. Small. 2016;12:3094-3100
  36. 36. Bahadır EB, Sezgintürk MK. Lateral flow assays: Principles, designs and labels. TrAC Trends in Analytical Chemistry. 2016;82:286-306
  37. 37. Dzantiev BB, Byzova NA, Urusov AE, Zherdev AV. Immunochromatographic methods in food analysis. TrAC Trends in Analytical Chemistry. 2014;55:81-93
  38. 38. Singh J, Sharma S, Nara S. Evaluation of gold nanoparticle based lateral flow assays for diagnosis of Enterobacteriaceae members in food and water. Food Chemistry. 2015;170:470-483
  39. 39. Sajid M, Kawde AN, Daud M. Designs, formats and applications of lateral flow assay: A literature review. Journal of Saudi Chemical Society. 2015;19:689-705
  40. 40. Huang X, Aguilar ZP, Xu H, Lai W, Xiong Y. Membrane-based lateral flow immunochromatographic strip with nanoparticles as reporters for detection: A review. Biosensors & Bioelectronics. 2016;75:166-180
  41. 41. Song CM, Liu C, Wu SY, Li HL, Guo HQ, Yang B, Qiu S, Li JW, Liu L, Zeng HJ, et al. Development of a lateral flow colloidal gold immunoassay strip for the simultaneous detection of Shigella boydii and Escherichia coli O157:H7 in bread, milk and jelly samples. Food Control. 2016;59:345-351
  42. 42. Pyo D, Yoo J. New trends in fluorescence immunochromatography. Journal of Immunoassay and Immunochemistry. 2012;33:203-222
  43. 43. Ju Q, Noor MO, Krull UJ. Paper-based biodetection using luminescent nanoparticles. Analyst. 2016;141:2838-2860
  44. 44. Cordeiro M, Carlos FF, Pedrosa P, Lopez A, Baptista PV. Gold nanoparticles for diagnostics: Advances towards points of care. Diagnostics. 2016;6:43
  45. 45. Jung BY, Jung SC, Kweon CH. Development of a rapid immunochromatographic strip for detection of Escherichia coli O157. Journal of Food Protection. 2005;68:2140-2143
  46. 46. Bautista DA, Elankumaran S, Arking JA, Heckert RA. Evaluation of an immunochromatography strip assay for the detection of Salmonella sp from poultry. Journal of Veterinary Diagnostic Investigation. 2002;14:427-430
  47. 47. Huang SH, Wei HC, Lee YC. One-step immunochromatographic assay for the detection of Staphylococcus aureus. Food Control. 2007;18:893-897
  48. 48. Shim WB, Choi JG, Kim JY, Yang ZY, Lee KH, Kim MG, Ha SD, Kim KS, Kim KY, Kim CH, et al. Production of monoclonal antibody against Listeria monocytogenes and its application to immunochromatography strip test. Journal of Microbiology and Biotechnology. 2007;17:1152-1161
  49. 49. Preechakasedkit P, Pinwattana K, Dungchai W, Siangproh W, Chaicumpa W, Tongtawe P, Chailapakul O. Development of a one-step immunochromatographic strip test using gold nanoparticles for the rapid detection of Salmonella typhi in human serum. Biosensors & Bioelectronics. 2012;31:562-566
  50. 50. Li XP, Lu DL, Sheng ZH, Chen K, Guo XB, Jin ML, Han HY. A fast and sensitive immunoassay of avian influenza virus based on label-free quantum dot probe and lateral flow test strip. Talanta. 2012;100:1-6
  51. 51. Yang Q, Gong X, Song T, Yang J, Zhu S, Li Y, Cui Y, Zhang B, Chang J. Quantum dot-based immunochromatography test strip for rapid, quantitative and sensitive detection of alpha fetoprotein. Biosensors & Bioelectronics. 2011;30:145-150
  52. 52. Li X, Li W, Yang Q, Gong X, Guo W, Dong C, Liu J, Xuan L, Chang J. Rapid and quantitative detection of prostate specific antigen with a quantum dot nanobeads-based immunochromatography test strip. ACS Applied Materials & Interfaces. 2014;6:6406-6414
  53. 53. Medintz IL, Uyeda HT, Goldman ER, Mattoussi H. Quantum dot bioconjugates for imaging, labelling and sensing. Nature Materials. 2005;4:435-446
  54. 54. Alivisatos AP. Semiconductor clusters, nanocrystals, and quantum dots. Science. 1996;271:933-937
  55. 55. Resch-Genger U, Grabolle M, Cavaliere-Jaricot S, Nitschke R, Nann T. Quantum dots versus organic dyes as fluorescent labels. Nature Methods. 2008;5:763-775
  56. 56. Bruno J. Application of DNA aptamers and quantum dots to lateral flow test strips for detection of foodborne pathogens with improved sensitivity versus colloidal gold. Pathogens. 2014;3:341-355
  57. 57. Quesada-González D, Merkoçi A. Nanoparticle-based lateral flow biosensors. Biosensors & Bioelectronics. 2015;73:47-63
  58. 58. Chen XX, Gan M, Xu H, Chen F, Ming X, Xu HY, Wei H, Xu F, Liu CW. Development of a rapid and sensitive quantum dot-based immunochromatographic strip by double labeling PCR products for detection of Staphylococcus aureus in food. Food Control. 2014;46:225-232
  59. 59. Liu CY, Jia QJ, Yang CH, Qiao RR, Jing LH, Wang LB, Xu CL, Gao MY. Lateral flow immunochromatographic assay for sensitive pesticide detection by using Fe3O4 nanoparticle aggregates as color reagents. Analytical Chemistry. 2011;83:6778-6784
  60. 60. Wang DB, Tian B, Zhang ZP, Wang XY, Fleming J, Bi LJ, Yang RF, Zhang XE. Detection of Bacillus anthracis spores by super-paramagnetic lateral-flow immunoassays based on “road closure”. Biosensors & Bioelectronics. 2015;67:608-614
  61. 61. Qiao ZH, Lei CY, Fu YC, Li YB. Rapid and sensitive detection of E-coli O157:H7 based on antimicrobial peptide functionalized magnetic nanoparticles and urease-catalyzed signal amplification. Analytical Methods. 2017;9:5204-5210
  62. 62. Ren W, Cho H, Zhou Z, Irudayaraj J. Ultrasensitive detection of microbial cells using magnetic focus enhanced lateral flow sensors. Chemical Communications. 2016;52:4930-4933
  63. 63. Suaifan G, Alhogail S, Zourob M. Paper-based magnetic nanoparticle-peptide probe for rapid and quantitative colorimetric detection of Escherichia coli O157:H7. Biosensors & Bioelectronics. 2017;92:702-708
  64. 64. Xia SQ, Yu ZB, Liu DF, Xu CL, Lai WH. Developing a novel immunochromatographic test strip with gold magnetic bifunctional nanobeads (GMBN) for efficient detection of Salmonella choleraesuis in milk. Food Control. 2016;59:507-512
  65. 65. Jyoti A, Vajpayee P, Singh G, Patel CB, Gupta KC, Shanker R. Identification of environmental reservoirs of nontyphoidal Salmonellosis: Aptamer-assisted bioconcentration and subsequent detection of Salmonella typhimurium by quantitative polymerase chain reaction. Environmental Science & Technology. 2011;45:8996-9002
  66. 66. Lorsch JR, Szostak JW. In-vitro selection of RNA aptamers specific for cyanocobalamin. Biochemistry. 1994;33:973-982
  67. 67. Wrzesinski J, Ciesiolka J. Characterization of structure and metal ions specificity of Co2+-binding RNA aptamers. Biochemistry. 2005;44:6257-6268
  68. 68. Bock LC, Griffin LC, Latham JA, Vermaas EH, Toole JJ. Selection of single-stranded-DNA molecules that bind and inhibit human thrombin. Nature. 1992;355:564-566
  69. 69. Gao FL, Lei JP, Ju HX. Label-free surface-enhanced Raman spectroscopy for sensitive DNA detection by DNA-mediated silver nanoparticle growth. Analytical Chemistry. 2013;85:11788-11793
  70. 70. Jiang XH, Yang M, Meng YJ, Jiang W, Zhan JH. Cysteamine-modified silver nanoparticle aggregates for quantitative SERS sensing of pentachlorophenol with a portable Raman spectrometer. ACS Applied Materials & Interfaces. 2013;5:6902-6908
  71. 71. Shi ML, Zheng J, Tan YJ, Tan GX, Li JS, Li YH, Li X, Zhou ZG, Yang RH. Ultrasensitive detection of single nucleotide polymorphism in human mitochondrial DNA utilizing ion-mediated cascade surface-enhanced Raman spectroscopy amplification. Analytical Chemistry. 2015;87:2734-2740
  72. 72. Gao FL, Du LL, Tang DQ, Lu Y, Zhang YZ, Zhang LX. A cascade signal amplification strategy for surface enhanced Raman spectroscopy detection of thrombin based on DNAzyme assistant DNA recycling and rolling circle amplification. Biosensors & Bioelectronics. 2015;66:423-430
  73. 73. Xu LJ, Lei ZC, Li JX, Zong C, Yang CJ, Ren B. Label-free surface-enhanced Raman spectroscopy detection of DNA with single-base sensitivity. Journal of the American Chemical Society. 2015;137:5149-5154
  74. 74. Liu Y, Zhou HB, Hu ZW, Yu GX, Yang DT, Zhao JS. Label and label-free based surface-enhanced Raman scattering for pathogen bacteria detection: a review. Biosensors & Bioelectronics. 2017;94:131-140
  75. 75. Wang JF, Wu XZ, Wang CW, Shao NS, Dong PT, Xiao R, Wang SQ. Magnetically assisted surface-enhanced Raman spectroscopy for the detection of Staphylococcus aureus based on aptamer recognition. ACS Applied Materials & Interfaces. 2015;7:20919-20929
  76. 76. Zhang H, Ma X, Liu Y, Duan N, Wu S, Wang Z, Xu B. Gold nanoparticles enhanced SERS aptasensor for the simultaneous detection of Salmonella typhimurium and Staphylococcus aureus. Biosensors & Bioelectronics. 2015;74:872-877
  77. 77. Pahlow S, Meisel S, Cialla-May D, Weber K, Rosch P, Popp J. Isolation and identification of bacteria by means of Raman spectroscopy. Advanced Drug Delivery Reviews. 2015;89:105-120
  78. 78. Gao WC, Li B, Yao RZ, Li ZP, Wang XW, Dong XL, Qu H, Li QX, Li N, Chi H, et al. Intuitive label-free SERS detection of bacteria using aptamer-based in situ silver nanoparticles synthesis. Analytical Chemistry. 2017;89:9836-9842
  79. 79. Yu MQ, Wang H, Fu F, Li LY, Li J, Li G, Song Y, Swihart MT, Song EQ. Dual-recognition förster resonance energy transfer based platform for one-step sensitive detection of pathogenic bacteria using fluorescent vancomycin-gold nanoclusters and aptamer-gold nanoparticles. Analytical Chemistry. 2017;89:4085-4090
  80. 80. Duan N, Gong WH, Wang ZP, Wu SJ. An aptasensor based on fluorescence resonance energy transfer for multiplexed pathogenic bacteria determination. Analytical Methods. 2016;8:1390-1395
  81. 81. Duan N, Ding XY, He LX, Wu SJ, Wei YX, Wang ZP. Selection, identification and application of a DNA aptamer against Listeria monocytogenes. Food Control. 2013;33:239-243
  82. 82. Lian Y, He FJ, Wang H, Tong FF. A new aptamer/graphene interdigitated gold electrode piezoelectric sensor for rapid and specific detection of Staphylococcus aureus. Biosensors & Bioelectronics. 2015;65:314-319
  83. 83. Jiang DN, Liu F, Zhang LQ, Liu LL, Liu C, Pu XY. An electrochemical strategy with molecular beacon and hemin/G-quadruplex for the detection of Clostridium perfringens DNA on screen-printed electrodes. RSC Advances. 2014;4:57064-57070
  84. 84. Wu SJ, Duan N, Shi Z, Fang CC, Wang ZP. Simultaneous aptasensor for multiplex pathogenic bacteria detection based on multicolor upconversion nanoparticles labels. Analytical Chemistry. 2014;86:3100-3107
  85. 85. Labib M, Zamay AS, Koloyskaya OS, Reshetneva IT, Zamay GS, Kibbee RJ, Sattar SA, Zamay TN, Berezovski MV. Aptamer-based viability impedimetric sensor for bacteria. Analytical Chemistry. 2012;84:8966-8969
  86. 86. Abbaspour A, Norouz-Sarvestani F, Noon A, Soltani N. Aptamer-conjugated silver nanoparticles for electrochemical dual-aptamer-based sandwich detection of Staphylococcus aureus. Biosensors & Bioelectronics. 2015;68:149-155
  87. 87. Ding JW, Lei JH, Ma X, Gong J, Qin W. Potentiometric aptasensing of Listeria monocytogenes using protamine as an indicator. Analytical Chemistry. 2014;86:9412-9416

Written By

Lingwen Zeng, Lihua Wang and Jiao Hu

Submitted: 30 May 2017 Reviewed: 15 December 2017 Published: 21 March 2018