Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2150 2022-06-08 00:21:30 |
2 format corrected. Meta information modification 2150 2022-06-08 06:06:10 | |
3 references added. -39 word(s) 2111 2022-06-08 06:07:27 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Rasetti-Escargueil, C.; Popoff, M. Applications of BoNT Detection. Encyclopedia. Available online: https://encyclopedia.pub/entry/23802 (accessed on 30 August 2024).
Rasetti-Escargueil C, Popoff M. Applications of BoNT Detection. Encyclopedia. Available at: https://encyclopedia.pub/entry/23802. Accessed August 30, 2024.
Rasetti-Escargueil, Christine, Michel Popoff. "Applications of BoNT Detection" Encyclopedia, https://encyclopedia.pub/entry/23802 (accessed August 30, 2024).
Rasetti-Escargueil, C., & Popoff, M. (2022, June 08). Applications of BoNT Detection. In Encyclopedia. https://encyclopedia.pub/entry/23802
Rasetti-Escargueil, Christine and Michel Popoff. "Applications of BoNT Detection." Encyclopedia. Web. 08 June, 2022.
Applications of BoNT Detection
Edit

Botulinum neurotoxins (BoNTs) are produced as protein complexes by bacteria of the genus Clostridium that are Gram-positive, anaerobic and spore forming (Clostridium botulinum, C. butyricum, C. baratii and C. argentinense spp.). BoNTs show a high immunological and genetic diversity. Therefore, fast, precise, and more reliable detection methods are still required to monitor outbreaks and ensure surveillance of botulism. The botulinum toxin field also comprises therapeutic uses, basic research studies and biodefense issues.

botulinum neurotoxins detection botulism in vitro in vivo cell-based assays countermeasures

1. Introduction

Indeed, BoNTs are divided into nine toxinotypes (A, B, C, D, E, F, G, H or F/A, X) based on their neutralization by specific corresponding antisera, and their actions on different substrates. In addition, each toxinotype is subdivided into subtypes based on amino acid variations. Currently, 41 subtypes have been identified [1]. While types A, B, E and F are mainly responsible for human botulism, toxinotypes C and D are associated with animal botulism mainly occurring in birds and cattle. The various BoNTs toxinotypes and subtypes interact with distinct membrane receptors, and cleave different intracellular SNARE proteins (soluble N-ethylmaleimide-sensitive factor attachment receptor), VAMP (vesicle associated membrane protein)/synaptobrevin, synaptosomal-associated protein 25 (SNAP-25), and syntaxin, at different cleavage sites [2][3]. The rapid progress in genomic information has revealed the presence of bont related sequences in non-clostridial strains, such as bont/Wo or bont/I detected in the genome of the Weisenella oryzae, a bacterium of fermented rice; the bont/J (ebont/F or bont/En) found in the genome of a Enterococcus faecalis strain isolated from cow feces; Cp1 from Chryseobacterium piperi from sediment [4][5][6][7]; and the paraclostridial mosquitocidal protein 1 (PMP1) in a Paraclostridium bifermentans strain [8].
Botulism occurs in three major syndromes: food-borne botulism due to consumption of preformed BoNT in contaminated food, wound botulism, and botulism by intestinal colonization (infant botulism and adult intestinal toxemia). C. botulinum can grow in various food matrices and produce toxin as a result of inappropriate storage temperature in association with anaerobic conditions. Quantities as low as 30 ng BoNT can cause botulism. Wound botulism also occurs within injured tissues where the spores can grow and produce toxin in the tissues [9]. C. botulinum spores may result in a toxico-infection by colonization of the intestinal tract and in situ BoNT production. Children under the age of 1 year can develop infant botulism, since they are more susceptible to intestinal colonization by C. botulinum [10][11][12]. In addition, inhalational botulism can result from aerosolization of BoNT in rare cases of laboratory botulism, and iatrogenic botulism can result from the injection of BoNT overdoses after therapeutic or cosmetic use [13][14][15][16].

2. Applications of BoNT Detection

2.1. Clinical Presentation of Botulism

Botulism cases are initially investigated based on the clinical presentation provided by the clinicians, and only confirmed after thorough laboratory investigation of the presence of BoNTs and/or clostridial organisms able to produce BoNTs. Symptoms occur within 6–36 h after toxin ingestion, and it is challenging for clinicians to differentiate symptoms of botulism from certain neurological conditions such as Gillain-Barré syndrome, or cerebrovascular accidents. Very often, botulism presentation consists of gastrointestinal early symptoms, such as constipation, abdominal pain, nausea and vomiting; however, these types of symptoms may also be caused by microorganisms not related to Clostridium botulinum, hence delaying the proper diagnosis based on flaccid paralysis [17].
Rapid administration of the antitoxin is currently the only efficient therapy to treat the disease, which implies a rapid and accurate diagnosis since the laboratory confirmation takes a few days. Medical treatment also includes supportive care, intubation, and mechanical ventilation when necessary. However, once the toxin has entered the neurons, intoxication remains irreversible, and no post-exposure therapy exists [18].

2.2. Laboratory Confirmation

The detection of BoNT, or identification of the producing organism, is a requirement for laboratory confirmation of botulism. However, the current gold standard assay, the mouse bioassay (MBA), is causing ethical concerns and is difficult to standardize. In order to address this challenge, many in vitro or cell-based assays have been developed to detect BoNTs or characterize the BoNT-producing organism using PCR, to reduce animal use or even replace the MBA. Nevertheless, many of the recent assays are not fully validated for the detection of all toxinotypes in clinical specimen or food matrices. BoNT detection is currently performed using the in vivo MBA routinely used to detect BoNTs in serum, feces of patients or in suspected food or environmental samples. BoNT detection in the serum of suspected patients is the most direct way to confirm a diagnosis of botulism. However, very sensitive methods of BoNT detection are required to identify the minute BoNT amounts in sera [9]. The MBA is also able to detect the wide spectrum of BoNT subtypes described in the literature, but it cannot distinguish the individual subtypes [16][18].
Many DNA-based methods have been developed to identify the presence of C. botulinum or spores in food, clinical or environmental samples, and study the phylogeny of outbreak strains to complete epidemiological investigations [10][17]. However, actual detection of preformed BoNT in food suspected to be responsible for a botulism outbreak is required to support the diagnosis of botulism.

2.3. Food Industry: Survey of Food Safety

Food safety regulations have been established by the World Trade Organization (WTO) to ensure the safety of food products worldwide and reduce risks to human health. Risk assessment in parallel with scientifically based food safety criteria have been established to assess the appropriate level of protection. The microbiological criteria are reviewed at regular intervals to account for new information on, for example, new scientific knowledge including newly identified pathogens and infectious doses. Several food safety regulations have been established around the globe to ensure safe food processing and decrease the incidence of food poisoning. Food safety has become a growing concern in Asia; major Asiatic countries have now established regulatory agencies to implement food safety regulations [19]. C. botulinum of Group I are widely spread across the environment, and pose the risk of food-borne botulism if they enter into the food chain, leading to consumption of pre-formed BoNT in contaminated food. The high thermally resistant spores produced by group I C. botulinum strains cause great concern in low-acid canned food. Thus, sensitive detection of BoNT in food samples represents the most reliable approach for source-tracking. Subsequent regulatory decisions, such as batch recalls or seizure, are based on laboratory methods confirming the presence of BoNT as a health hazard [17][18][19][20].
Detection of BoNT in food or water is also an important aspect in biodefense. BoNT is classified as one of the most potentially dangerous bioweapons in category A of the classification of the Centers for Disease Control and Prevention (CDC) of Atlanta (Classification of biological agents according to the Centers for Diseases Control and Prevention. http://www.bt.cdc.gov/agent/agentlistcategory.asp, accessed on 1 January 2022). BoNT could be spread through food, water or aerosols [21][22].

2.4. Pharmaceutical Industry Uses, Needs for Precise Quantification of Active BoNT

BoNT/A and BoNT/B are available as licensed pharmaceutical products to treat cervical dystonia, blepharospasm, spastic conditions, hyperhidrosis, and in an increasing number of other medical indications. BoNT is also used as a cosmetic to reduce facial lines. Safe dosing of BoNT products requires accurate and reliable measurement of potency before batch release onto the market. The reference test for potency testing is the MBA. Indeed, BoNT therapeutic units are mouse lethal doses (LD50). The biological activity of BoNT has to be determined accurately for safety and efficacy of the product. Current requirements of the European Pharmacopea for potency determination stipulate that every production batch of BoNT must be tested using the MBA [23]. In the USA, the Food and Drug Administration (FDA) requires the submission of data on BoNT potency (using in vivo and/or approved in vitro methods), in addition to the safety, purity, sterility and other parameters for these products. The regulations are specified in the Code of Federal Regulations 21.

3. Discussion: Suitability of Each Method for Their Applications

To summarize the findings of this research, many different methods have been developed for the detection of botulinum neurotoxins. They all have potential to replace the MBA to address the three Rs principles: replacement, refinement and reduction in animal research and use [24]. Many in vitro methods show sensitivity levels lower than that of the MBA, as shown in Table 1 below; however, not all in vitro methods are able to distinguish the protein toxin from fully active BoNT.
Table 1. Methods of BoNT detection, sensitivities and durations.
Method Principles Analysis Time BoNT Toxinotype Sensitivity Benefits/Limitations References
Immunological methods: sandwich ELISA, electro-chemiluminescent assay 6–7 h A–F 2–176 pg/mL Rapid detection/detection of active and inactive BoNTs, detection hampered by neurotoxin associated proteins [25][26][27][28][29][30][31]
Immunological methods: lateral flow assay, 30 min A–B 10–50 ng/mL
(10,000–50,000 pg/mL)
Rapid detection/detection of active and inactive BoNTs, detection hampered by neurotoxin associated proteins [25][32][33]
Mass spectrometry 5–8 h A–F 0.1–1 pg/mL pg/mL Rapid detection/detection of active and inactive BoNTs [34][35][36][37][38][39]
Endopeptidase ELISA based or MS based 7–8 h A–G 0.1–1000 pg/mL Rapid detection/detection of cleavage only [39][40][41][42][43][44][45][46][47][48][49][50][51]
Immunosensors and FRET assays 2–5 h A 0.1–20 pg/mL Rapid detection/detection of active and inactive BoNTs [52][53][54][55][56][57]
In vivo mouse bioassay 4 days A–F 1–10 pg/mL Sensitive method detecting functional toxin but ethical concern, variability and duration [23][58][59]
Ex vivo methods
hemidiaphragm test
9–5 h A–F 1–10 pg/mL Sensitive method detecting functional toxin but ethical concern and technically demanding [60][61][62]
Cell-based assays
human neurons from induced pluripotent cells and monitoring of SNAP-25 cleavage by Western blot
3–5 days A–E 0.003 pM–10 pM
(0.55–1500 pg/mL)
Sensitive method detecting functional toxin but technically demanding [63][64][65][66][67][68][69][70][71][72][73][74]
Cell-based assays using differentiated cell lines 3–5 days A–E 5.5 pM–10 nM
(825–150,000 pg/mL)
Sensitive method detecting functional toxin but technically demanding [75][76][77][78][79][80][81][82][83][84][85][86]
Electrical conductance assays 1–3 days A 25,000 pg/mL Method detecting functional toxin but long and technically demanding [80][81]
The immunological methods offer a range of benefits, including rapid analysis time, reproducibility and high sensitivity. However, their implementation is subject to the availability of suitable tools and reagents, such as antibodies. In addition, immunological methods only detect the protein toxin itself, with no indication of its true functionality. The mass spectrometry approach is the most sensitive in vitro method currently available for a rapid detection of active or inactive forms of BoNTs. In addition, ELISA methods are easily transferable, and offer the potential for high-throughput analysis.
Note that these methods require inter-laboratory validation in different food, environmental, and clinical sample matrices before they can be implemented more widely in reference laboratories.
As opposed to ELISA or MS, the endopeptidase ELISA or MS-based methods offer the rapid and sensitive detection of the cleavage activity. Nevertheless, they still require the use of a range of cleavage site-specific antibodies, which is a limitation considering the wide diversity of the BoNTs; in contrast, the fully functional assays do not require a range of specific antibodies. In addition, these surrogate in vitro models also require full validation before they can be used as a replacement bioassay of potency.
Cell-based assays using neuronal cells in culture recapitulate all functional domains of the toxin but they are still at various stages of development. Moreover, the cell-based assays that use primary cells are not sufficiently robust as a result of high batch-to-batch variability of their cultures. The models that employ differentiated mouse and human stem cells offer high sensitivity, and may provide the most useful alternative models for toxin assessment as well as for functional assessment of antibodies. The mouse phenic nerve assay is in excellent agreement with the mouse bioassay for precise assessment of botulinum activity and antitoxin neutralization; however, this method is hardly transferable to routine laboratories [62].
However, it is worth noting that cell culture assays are subject to interference and non-specific toxic effects, as a result of the presence of adjuvants and other ingredients in the matrices. Therefore, the in vivo model currently remains as the method of choice for detecting BoNT in parallel with immune-biochemical assays that correlate well with in vivo potency [51].
Nevertheless, significant progress has also been achieved by establishing functional muscle-nerve co-culture systems developed using hiPSC-derived motor neurons (hMNs) and human immortalized skeletal muscle cells. The recent implementation of optogenetics, combined with live calcium imaging, allows the monitoring of BoNTs’ intoxication on synaptic transmission in human motor endplate models [87]. These types of seminal studies demonstrate the potential to replace the mouse bioassay with well-designed, innovative, cell-based systems. Cell-based assays have a strong potential to replace the MBA for BoNT potency determination in pharmaceutical formulations; they can also help identify suitable inhibitors while reducing the number of animals used. The rapid progress in cell differentiation procedures as well in innovations of detection methods will support the replacement of the in vivo techniques; however, recently, the development of suitable countermeasures will still require in vivo studies in complement with the in vitro or cell-based approaches [59].

References

  1. Peck, M.W.; Smith, T.J.; Anniballi, F.; Austin, J.W.; Bano, L.; Bradshaw, M.; Cuervo, P.; Cheng, L.W.; Derman, Y.; Dorner, B.G.; et al. Historical Perspectives and Guidelines for Botulinum Neurotoxin Subtype Nomenclature. Toxins 2017, 9, 38.
  2. Dong, M.; Masuyer, G.; Stenmark, P. Botulinum and Tetanus Neurotoxins. Annu. Rev. Biochem. 2019, 88, 811–837.
  3. Pirazzini, M.; Rossetto, O.; Eleopra, R.; Montecucco, C. Botulinum Neurotoxins: Biology, Pharmacology, and Toxicology. Pharmacol. Rev. 2017, 69, 200–235.
  4. Mansfield, M.J.; Adams, J.B.; Doxey, A.C. Botulinum neurotoxin homologs in non-Clostridium species. FEBS Lett. 2015, 589, 342–348.
  5. Zhang, S.; Masuyer, G.; Zhang, J.; Shen, Y.; Lundin, D.; Henriksson, L.; Miyashita, S.-I.; Martínez-Carranza, M.; Dong, M.; Stenmark, P. Identification and characterization of a novel botulinum neurotoxin. Nat. Commun. 2017, 8, 14130.
  6. Zornetta, I.; Tehran, D.A.; Arrigoni, G.; Anniballi, F.; Bano, L.; Leka, O.; Zanotti, G.; Binz, T.; Montecucco, C. The first non Clostridial botulinum-like toxin cleaves VAMP within the juxtamembrane domain. Sci. Rep. 2016, 6, 30257.
  7. Kosenina, S.; Masuyer, G.; Zhang, S.; Dong, M.; Stenmark, P. Crystal structure of the catalytic domain of the Weissella oryzae botulinum-like toxin. FEBS Lett. 2019, 593, 1403–1410.
  8. Contreras, E.; Masuyer, G.; Qureshi, N.; Chawla, S.; Dhillon, H.S.; Lee, H.L.; Chen, J.; Stenmark, P.; Gill, S.S. A neurotoxin that specifically targets Anopheles mosquitoes. Nat. Commun. 2019, 10, 2869.
  9. Rasetti-Escargueil, C.; Lemichez, E.; Popoff, M.R. Toxemia in Human Naturally Acquired Botulism. Toxins 2020, 12, 716.
  10. Fenicia, L.; Anniballi, F.; Aureli, P. Intestinal toxemia botulism in Italy, 1984–2005. Eur. J. Clin. Microbiol. Infect. Dis. 2007, 26, 385–394.
  11. Barash, J.R.; Arnon, S.S. Dual toxin-producing strain of Clostridium botulinum type Bf isolated from a California patient with infant botulism. J. Clin. Microbiol. 2004, 42, 1713–1715.
  12. Dabritz, H.A.; Hill, K.K.; Barash, J.R.; Ticknor, L.O.; Helma, C.H.; Dover, N.; Payne, J.R.; Arnon, S.S. Molecular epidemiology of infant botulism in California and elsewhere, 1976–2010. J. Infect. Dis. 2014, 210, 1711–1722.
  13. Sobel, J. Botulism. Clin. Infect. Dis. 2005, 41, 1167–1173.
  14. King, L.A.; Popoff, M.R.; Mazuet, C.; Espie, E.; Vaillant, V.; de Valk, H. Infant botulism in France, 1991–2009. Arch. Pediatr. 2010, 17, 1288–1292.
  15. Ghasemi, M.; Norouzi, R.; Salari, M.; Asadi, B. Iatrogenic botulism after the therapeutic use of botulinum toxin-A: A case report and review of the literature. Clin. Neuropharmacol. 2012, 35, 254–257.
  16. Floresta, G.; Patamia, V.; Gentile, D.; Molteni, F.; Santamato, A.; Rescifina, A.; Vecchio, M. Repurposing of FDA-Approved Drugs for Treating Iatrogenic Botulism: A Paired 3D-QSAR/Docking Approach(dagger). ChemMedChem 2020, 15, 256–262.
  17. Thirunavukkarasu, N.; Johnson, E.; Pillai, S.; Hodge, D.; Stanker, L.; Wentz, T.; Singh, B.; Venkateswaran, K.; McNutt, P.; Adler, M.; et al. Botulinum Neurotoxin Detection Methods for Public Health Response and Surveillance. Front. Bioeng. Biotechnol. 2018, 6, 80.
  18. Bremer, P.T.; Adler, M.; Phung, C.H.; Singh, A.K.; Janda, K.D. Newly Designed Quinolinol Inhibitors Mitigate the Effects of Botulinum Neurotoxin A in Enzymatic, Cell-Based, and ex Vivo Assays. J. Med. Chem. 2017, 60, 338–348.
  19. Iriajen, A.; Rasetti-Escargueil, C. Integrated Approach used by Government Agencies and Industry to Protect the Consumer: Food Safety Legislation. In Clostridium Botulinum: A Spore Forming Organism and a Challenge to Food Safety; Surman-Lee, S., Rasetti-Escargueil, C., Eds.; Nova Publishers: New York, NY, USA, 2012.
  20. Lindstrom, M.; Keto, R.; Markkula, A.; Nevas, M.; Hielm, S.; Korkeala, H. Multiplex PCR assay for detection and identification of Clostridium botulinum types A, B, E, and F in food and fecal material. Appl. Environ. Microbiol. 2001, 67, 5694–5699.
  21. Arnon, S.S.; Schechter, R.; Inglesby, T.V.; Henderson, D.A.; Bartlett, J.G.; Ascher, M.S.; Eitzen, E.; Fine, A.D.; Hauer, J.; Layton, M.; et al. Botulinum toxin as a biological weapon: Medical and public health management. JAMA 2001, 285, 1059–1070.
  22. Cenciarelli, O.; Riley, P.W.; Baka, A. Biosecurity Threat Posed by Botulinum Toxin. Toxins 2019, 11, 681.
  23. Adler, S.; Bicker, G.; Bigalke, H.; Bishop, C.; Blümel, J.; Dressler, D.; Fitzgerald, J.; Gessler, F.; Heuschen, H.; Kegel, B.; et al. The current scientific and legal status of alternative methods to the LD50 test for botulinum neurotoxin potency testing. The report and recommendations of a ZEBET Expert Meeting. Altern. Lab. Anim. 2010, 38, 315–330.
  24. Hobbs, R.J.; Thomas, C.A.; Halliwell, J.; Gwenin, C.D. Rapid Detection of Botulinum Neurotoxins-A Review. Toxins 2019, 11, 418.
  25. Singh, A.K.; Stanker, L.H.; Sharma, S.K. Botulinum neurotoxin: Where are we with detection technologies? Crit. Rev. Microbiol. 2013, 39, 43–56.
  26. Lindstrom, M.; Kiviniemi, K.; Korkeala, H. Hazard and control of group II (non-proteolytic) Clostridium botulinum in modern food processing. Int. J. Food Microbiol. 2006, 108, 92–104.
  27. Worbs, S.; Fiebig, U.; Zeleny, R.; Schimmel, H.; Rummel, A.; Luginbühl, W.; Dorner, B.G. Qualitative and Quantitative Detection of Botulinum Neurotoxins from Complex Matrices: Results of the First International Proficiency Test. Toxins 2015, 7, 4935–4966.
  28. Sharma, S.K.; Ferreira, J.L.; Eblen, B.S.; Whiting, R.C. Detection of type A, B, E, and F Clostridium botulinum neurotoxins in foods by using an amplified enzyme-linked immunosorbent assay with digoxigenin-labeled antibodies. Appl. Environ. Microbiol. 2006, 72, 1231–1238.
  29. Singh, A.; Datta, S.; Sachdeva, A.; Maslanka, S.; Dykes, J.; Skinner, G.; Burr, D.; Whiting, R.C.; Sharma, S.K. Evaluation of an enzyme-linked immunosorbent assay (ELISA) kit for the detection of botulinum neurotoxins A, B, E, and F in selected food matrices. Health Secur. 2015, 13, 37–44.
  30. Cheng, L.W.; Stanker, L.H. Detection of botulinum neurotoxin serotypes A and B using a chemiluminescent versus electrochemiluminescent immunoassay in food and serum. J. Agric. Food Chem. 2013, 61, 755–760.
  31. von Berg, L.; Stern, D.; Weisemann, J.; Rummel, A.; Dorner, M.B.; Dorner, B.G. Optimization of SNAP-25 and VAMP-2 Cleavage by Botulinum Neurotoxin Serotypes A–F Employing Taguchi Design-of-Experiments. Toxin 2019, 11, 588.
  32. Simon, S.; Fiebig, U.; Liu, Y.; Tierney, R.; Dano, J.; Worbs, S.; Endermann, T.; Nevers, M.-C.; Volland, H.; Sesardic, D.; et al. Recommended Immunological Strategies to Screen for Botulinum Neurotoxin-Containing Samples. Toxins 2015, 7, 5011–5034.
  33. Gessler, F.; Pagel-Wieder, S.; Avondet, M.A.; Böhnel, H. Evaluation of lateral flow assays for the detection of botulinum neurotoxin type A and their application in laboratory diagnosis of botulism. Diagn. Microbiol. Infect. Dis. 2007, 57, 243–249.
  34. Kalb, S.R.; Goodnough, M.C.; Malizio, C.J.; Pirkle, J.L.; Barr, J.R. Detection of botulinum neurotoxin A in a spiked milk sample with subtype identification through toxin proteomics. Anal. Chem. 2005, 77, 6140–6146.
  35. Mazuet, C.; Ezan, E.; Volland, H.; Popoff, M.R.; Becher, F. Toxin detection in patients’ sera by mass spectrometry during two outbreaks of type A Botulism in France. J. Clin. Microbiol. 2012, 50, 4091–4094.
  36. Sulaiman, I.M.; Miranda, N.; Simpson, S. MALDI-TOF Mass Spectrometry and 16S rRNA Gene Sequence Analysis for the Identification of Foodborne Clostridium spp. J. AOAC Int. 2021, 104, 1381–1388.
  37. Kull, S.; Pauly, D.; Störmann, B.; Kirchner, S.; Stämmler, M.; Dorner, M.B.; Lasch, P.; Naumann, D.; Dorner, B.G. Multiplex detection of microbial and plant toxins by immunoaffinity enrichment and matrix-assisted laser desorption/ionization mass spectrometry. Anal. Chem. 2010, 82, 2916–2924.
  38. Koike, H.; Kanda, M.; Hayashi, H.; Matsushima, Y.; Yoshikawa, S.; Ohba, Y.; Hayashi, M.; Nagano, C.; Sekimura, K.; Otsuka, K.; et al. Development of an alternative approach for detecting botulinum neurotoxin type A in honey: Analysis of non-toxic peptides with a reference labelled protein via liquid chromatography-tandem mass spectrometry. Food Addit. Contam. Part A Chem. Anal. Control Expo Risk Assess. 2020, 37, 1359–1373.
  39. Morineaux, V.; Mazuet, C.; Hilaire, D.; Enche, J.; Popoff, M.R. Characterization of botulinum neurotoxin type A subtypes by immunocapture enrichment and liquid chromatography-tandem mass spectrometry. Anal. Bioanal. Chem. 2015, 407, 5559–5570.
  40. Kalb, S.R.; Moura, H.; Boyer, A.E.; McWilliams, L.G.; Pirkle, J.L.; Barr, J.R. The use of Endopep-MS for the detection of botulinum toxins A, B, E, and F in serum and stool samples. Anal. Biochem. 2006, 351, 84–92.
  41. Rosen, O.; Feldberg, L.; Gura, S.; Zichel, R. A new peptide substrate for enhanced botulinum neurotoxin type B detection by endopeptidase-liquid chromatography-tandem mass spectrometry/multiple reaction monitoring assay. Anal. Biochem. 2015, 473, 7–10.
  42. Rosen, O.; Feldberg, L.; Gura, S.; Zichel, R. Improved detection of botulinum type E by rational design of a new peptide substrate for endopeptidase-mass spectrometry assay. Anal. Biochem. 2014, 456, 50–52.
  43. Kalb, S.R.; Baudys, J.; Wang, D.; Barr, J.R. Recommended mass spectrometry-based strategies to identify botulinum neurotoxin-containing samples. Toxins 2015, 7, 1765–1778.
  44. Kalb, S.R.; Baudys, J.; Smith, T.J.; Smith, L.A.; Barr, J.R. Characterization of Hemagglutinin Negative Botulinum Progenitor Toxins. Toxins 2017, 9, 193.
  45. Rosen, O.; Feldberg, L.; Yamin, T.S.; Dor, E.; Barnea, A.; Weissberg, A.; Zichel, R. Development of a multiplex Endopep-MS assay for simultaneous detection of botulinum toxins A, B and E. Sci. Rep. 2017, 7, 14859.
  46. Kalb, S.R.; Baudys, J.; Kiernan, K.; Wang, D.; Becher, F.; Barr, J.R. Proposed BoNT/A and /B Peptide Substrates Cannot Detect Multiple Subtypes in the Endopep-MS Assay. J. Anal. Toxicol. 2020, 44, 173–179.
  47. Wang, D.; Baudys, J.; Hoyt, K.M.; Barr, J.R.; Kalb, S.R. Further optimization of peptide substrate enhanced assay performance for BoNT/A detection by MALDI-TOF mass spectrometry. Anal. Bioanal. Chem. 2017, 409, 4779–4786.
  48. Wang, D.; Baudys, J.; Hoyt, K.; Barr, J.R.; Kalb, S.R. Sensitive detection of type G botulinum neurotoxin through Endopep-MS peptide substrate optimization. Anal. Bioanal. Chem. 2019, 411, 5489–5497.
  49. Tevell Aberg, A.; Karlsson, I.; Hedeland, M. Modification and validation of the Endopep-mass spectrometry method for botulinum neurotoxin detection in liver samples with application to samples collected during animal botulism outbreaks. Anal. Bioanal. Chem. 2021, 413, 345–354.
  50. Jones, R.G.A.; Ochiai, M.; Liu, Y.; Ekong, T.; Sesardic, D. Development of improved SNAP25 endopeptidase immuno-assays for botulinum type A and E toxins. J. Immunol. Methods 2008, 329, 92–101.
  51. Liu, Y.Y.; Rigsby, P.; Sesardic, D.; Marks, J.D.; Jones, R.G. A functional dual-coated (FDC) microtiter plate method to replace the botulinum toxin LD50 test. Anal. Biochem. 2012, 425, 28–35.
  52. Lévêque, C.; Ferracci, G.; Maulet, Y.; Grand-Masson, C.; Blanchard, M.-P.; Seagar, M.; El Far, O. A substrate sensor chip to assay the enzymatic activity of Botulinum neurotoxin A. Biosens. Bioelectron. 2013, 49, 276–281.
  53. Ferracci, G.; Marconi, S.; Mazuet, C.; Jover, E.; Blanchard, M.-P.; Seagar, M.; Popoff, M.; Lévêque, C. A label-free biosensor assay for botulinum neurotoxin B in food and human serum. Anal. Biochem. 2011, 410, 281–288.
  54. Cheng, H.P.; Chuang, H.S. Rapid and Sensitive Nano-Immunosensors for Botulinum. ACS Sens. 2019, 4, 1754–1760.
  55. Narayanan, J.; Sharma, M.K.; Ponmariappan, S.; Sarita Shaik, M.; Upadhyay, S. Electrochemical immunosensor for botulinum neurotoxin type-E using covalently ordered graphene nanosheets modified electrodes and gold nanoparticles-enzyme conjugate. Biosens. Bioelectron. 2015, 69, 249–256.
  56. Joshi, S.G. Detection of biologically active botulinum neurotoxin—A in serum using high-throughput FRET-assay. J. Pharmacol. Toxicol. Methods 2012, 65, 8–12.
  57. Guo, J.; Xu, C.; Li, X.; Chen, S. A simple, rapid and sensitive FRET assay for botulinum neurotoxin serotype B detection. PLoS ONE 2014, 9, e114124.
  58. Das, R.G.; Sesardic, D. Alternatives to the LD50 assay for botulinum toxin potency testing: Strategies and progress towards refinement, reduction and replacement. AATEX 14. In Proceedings of the 6th World Congress on Alternatives & Animal Use in the Life Sciences, Tokyo, Janpan, 21–25 August 2007; pp. 581–585.
  59. Pellett, S.; Tepp, W.H.; Johnson, E.A. Critical Analysis of Neuronal Cell and the Mouse Bioassay for Detection of Botulinum. Neurotoxins 2019, 11, 713.
  60. Broide, R.S.; Rubino, J.; Nicholson, G.S.; Ardila, M.C.; Brown, M.S.; Aoki, K.R.; Francis, J. The rat Digit Abduction Score (DAS) assay: A physiological model for assessing botulinum neurotoxin-induced skeletal muscle paralysis. Toxicon 2013, 71, 18–24.
  61. Aoki, K.R. A comparison of the safety margins of botulinum neurotoxin serotypes A, B, and F in mice. Toxicon 2001, 39, 1815–1820.
  62. Rasetti-Escargueil, C.; Jones, R.G.A.; Liu, Y.; Sesardic, D. Measurement of botulinum types A, B and E neurotoxicity using the phrenic nerve-hemidiaphragm: Improved precision with in-bred mice. Toxicon 2009, 53, 503–511.
  63. Whitemarsh, R.C.M.; Strathman, M.J.; Chase, L.G.; Stankewicz, C.; Tepp, W.H.; Johnson, E.A.; Pellett, S. Novel application of human neurons derived from induced pluripotent stem cells for highly sensitive botulinum neurotoxin detection. Toxicol. Sci. 2012, 126, 426–435.
  64. McNutt, P.; Celver, J.; Hamilton, T.; Mesngon, M. Embryonic stem cell-derived neurons are a novel, highly sensitive tissue culture platform for botulinum research. Biochem. Biophys. Res. Commun. 2011, 405, 85–90.
  65. Kiris, E.; Nuss, J.E.; Burnett, J.C.; Kota, K.P.; Koh, D.C.; Wanner, L.M.; Torres-Melendez, E.; Gussio, R.; Tessarollo, L.; Bavari, S. Embryonic stem cell-derived motoneurons provide a highly sensitive cell culture model for botulinum neurotoxin studies, with implications for high-throughput drug discovery. Stem Cell Res. 2011, 6, 195–205.
  66. Pellett, S. Progress in cell based assays for botulinum neurotoxin detection. Curr. Top. Microbiol. Immunol. 2013, 364, 257–285.
  67. Keller, J.E.; Neale, E.A. The role of the synaptic protein snap-25 in the potency of botulinum neurotoxin type A. J. Biol. Chem. 2001, 276, 13476–13482.
  68. Sheridan, R.E.; Smith, T.J.; Adler, M. Primary cell culture for evaluation of botulinum neurotoxin antagonists. Toxicon 2005, 45, 377–382.
  69. Stahl, A.M.; Ruthel, G.; Torres-Melendez, E.; Kenny, T.A.; Panchal, R.G.; Bavari, S. Primary cultures of embryonic chicken neurons for sensitive cell-based assay of botulinum neurotoxin: Implications for therapeutic discovery. J. Biomol. Screen. 2007, 12, 370–377.
  70. Keller, J.E.; Cai, F.; Neale, E.A. Uptake of botulinum neurotoxin into cultured neurons. Biochemistry 2004, 43, 526–532.
  71. Pellett, S.; Du, Z.W.; Pier, C.L.; Tepp, W.H.; Zhang, S.C.; Johnson, E.A. Sensitive and quantitative detection of botulinum neurotoxin in neurons derived from mouse embryonic stem cells. Biochem. Biophys. Res. Commun. 2011, 404, 388–392.
  72. Pellett, S.; Schwartz, M.P.; Tepp, W.H.; Josephson, R.; Scherf, J.M.; Pier, C.L.; Thomson, J.A.; Murphy, W.; Johnson, E.A. Human Induced Pluripotent Stem Cell Derived Neuronal Cells Cultured on Chemically-Defined Hydrogels for Sensitive In Vitro Detection of Botulinum Neurotoxin. Sci. Rep. 2015, 5, 14566.
  73. Kiris, E.; Kota, K.P.; Burnett, J.C.; Soloveva, V.; Kane, C.D.; Bavari, S. Recent developments in cell-based assays and stem cell technologies for botulinum neurotoxin research and drug discovery. Expert Rev. Mol. Diagn. 2014, 14, 153–168.
  74. Schenke, M.; Prause, H.C.; Bergforth, W.; Przykopanski, A.; Rummel, A.; Klawonn, F.; Seeger, B. Human-Relevant Sensitivity of iPSC-Derived Human Motor Neurons to BoNT/A1 and B1. Toxins 2021, 13, 585.
  75. Fernandez-Salas, E.; Wang, J.; Molina, Y.; Nelson, J.B.; Jacky, B.P.; Aoki, K.R. Botulinum neurotoxin serotype A specific cell-based potency assay to replace the mouse bioassay. PLoS ONE 2012, 7, e49516.
  76. Rasetti-Escargueil, C.C.M.; R Preneta-Blanc, R. Fleck and D. Sesardic. Enhanced sensitivity to Botulinum type A neurotoxin of human neuroblastoma SH-SY5Y cells after differentiation into mature neuronal cells. Botulinum J. 2011, 2, 30–48.
  77. Tegenge, M.A.; Bohnel, H.; Gessler, F.; Bicker, G. Neurotransmitter vesicle release from human model neurons (NT2) is sensitive to botulinum toxin A. Cell Mol. Neurobiol. 2012, 32, 1021–1029.
  78. Lee, J.O.; Rosenfield, J.; Tzipori, S.; Park, J.B. M17 human neuroblastoma cell as a cell model for investigation of botulinum neurotoxin A activity and evaluation of BoNT/A specific antibody. Botulinum J. 2008, 1, 135–152.
  79. Purkiss, J.R.; Friis, L.M.; Doward, S.; Quinn, C.P. Clostridium botulinum neurotoxins act with a wide range of potencies on SH-SY5Y human neuroblastoma cells. Neurotoxicology 2001, 22, 447–453.
  80. Hong, W.S.; Pezzi, H.M.; Schuster, A.R.; Berry, S.M.; Sung, K.E.; Beebe, D.J. Development of a Highly Sensitive Cell-Based Assay for Detecting Botulinum Neurotoxin Type A through Neural Culture Media Optimization. J. Biomol. Screen. 2016, 21, 65–73.
  81. Diamant, E.; Torgeman, A.; Epstein, E.; Mechaly, A.; Ben David, A.; Levin, L.; Schwartz, A.; Dor, E.; Girshengorn, M.; Barnea, A.; et al. A cell-based alternative to the mouse potency assay for pharmaceutical type E botulinum antitoxins. ALTEX 2022, 39, 113–122.
  82. Torgeman, A.; Diamant, E.; Levin, L.; Ben David, A.; Epstein, E.; Girshengorn, M.; Mazor, O.; Rosenfeld, R.; Zichel, R. An in vitro cell-based potency assay for pharmaceutical type A botulinum antitoxins. Vaccine 2017, 35, 7213–7216.
  83. Yowler, B.C.; Kensinger, R.D.; Schengrund, C.L. Botulinum neurotoxin A activity is dependent upon the presence of specific gangliosides in neuroblastoma cells expressing synaptotagmin I. J. Biol. Chem. 2002, 277, 32815–32819.
  84. Rust, A.; Doran, C.; Hart, R.; Binz, T.; Stickings, P.; Sesardic, D.; Peden, A.A.; Davletov, B. A Cell Line for Detection of Botulinum Neurotoxin Type B. Front. Pharmacol. 2017, 8, 796.
  85. Neuschafer-Rube, F.; Pathe-Neuschafer-Rube, A.; Puschel, G.P. Discrimination of the Activity of Low-Affinity Wild-Type and High-Affinity Mutant Recombinant BoNT/B by a SIMA Cell-Based Reporter Release Assay. Toxins 2022, 14, 65.
  86. Pathe-Neuschafer-Rube, A.; Neuschafer-Rube, F.; Haas, G.; Langoth-Fehringer, N.; Puschel, G.P. Cell-Based Reporter Release Assay to Determine the Potency of Proteolytic Bacterial Neurotoxins. Toxins 2018, 10, 360.
  87. de Lamotte, J.D.; Polentes, J.; Roussange, F.; Lesueur, L.; Feurgard, P.; Perrier, A.; Nicoleau, C.; Martinat, C. Optogenetically controlled human functional motor endplate for testing botulinum neurotoxins. Stem Cell Res. Ther. 2021, 12, 599.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 424
Revisions: 3 times (View History)
Update Date: 14 Jun 2022
1000/1000
ScholarVision Creations