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 -- 1972 2023-02-23 16:30:59 |
2 format correct Meta information modification 1972 2023-02-24 03:02:38 | |
3 format correct Meta information modification 1972 2023-02-24 03:07:54 | |
4 format correct + 6 word(s) 1978 2023-02-24 03:10:29 |

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.
Zhao, Z.; Zhang, Z.; Zhang, H.; Liang, Z. Applications of Small Peptides in Mycotoxin Detection. Encyclopedia. Available online: https://encyclopedia.pub/entry/41591 (accessed on 27 July 2024).
Zhao Z, Zhang Z, Zhang H, Liang Z. Applications of Small Peptides in Mycotoxin Detection. Encyclopedia. Available at: https://encyclopedia.pub/entry/41591. Accessed July 27, 2024.
Zhao, Zitong, Zhenzhen Zhang, Haoxiang Zhang, Zhihong Liang. "Applications of Small Peptides in Mycotoxin Detection" Encyclopedia, https://encyclopedia.pub/entry/41591 (accessed July 27, 2024).
Zhao, Z., Zhang, Z., Zhang, H., & Liang, Z. (2023, February 23). Applications of Small Peptides in Mycotoxin Detection. In Encyclopedia. https://encyclopedia.pub/entry/41591
Zhao, Zitong, et al. "Applications of Small Peptides in Mycotoxin Detection." Encyclopedia. Web. 23 February, 2023.
Applications of Small Peptides in Mycotoxin Detection
Edit

Mycotoxins pose significant risks to humans and livestock. In addition, contaminated food- and feedstuffs can only be discarded, leading to increased economic losses and potential ecological pollution. Mycotoxin removal and real-time toxin level monitoring are effective approaches to solve this problem.Small peptides derived from phage display peptide libraries, combinatorial peptide libraries, and rational design approaches can act as coating antigens, competitive antigens, and anti-immune complexes in immunoassays for the detection of mycotoxins. Furthermore, as a potential approach to mycotoxin degradation, small peptides can mimic the natural enzyme catalytic site to construct artificial enzymes containing oxidoreductases, hydrolase, and lyase activities. In summary, with the advantages of mature synthesis protocols, diverse structures, and excellent biocompatibility, also sharing their chemical structure with natural proteins, small peptides are widely used for mycotoxin detection and artificial enzyme construction, which have promising applications in mycotoxin degradation. 

Small peptides mycotoxin detection antigens

1. Introduction

Peptides have emerged as a promising approach to synthetic biomimetics [1]. The excellent properties of synthetic peptides in small-molecule contaminant (SMCs) recognition make them a potential alternative to antibodies and natural receptors in mycotoxin biosensor applications [2]. Small peptides show high-affinity binding to these small analytes and have stability, playing an increasingly important role in the rapid detection of mycotoxins [3]. Among immunologically based mycotoxin detection methods, peptides can serve as competing antigens, coating antigens, and anti-immune complexes (Table 1).

2. Peptides as Competing Antigens

As stated earlier, competitive antigens are essential in competitive immunoassays, but the obtaining of specificity antigens always involves the use of mycotoxins. Peptides can serve as competing antigens and competitively bind with specific monoclonal antibody (mAb), which can avoid the use of poisonous mycotoxins [11][14]. In 1999, Yuan et al. [32] identified two phage-displayed mimotopes (SWGPFPF and SWGPLPF) and used them to detect DON firstly. Subsequently, mimotopes as competitive antigens for the detection of mycotoxins have been increasingly reported, involving OTA, AFB1, ZEN, FB1, and other mycotoxins [33]. Peltomaa et al. [26] reported the selection of a novel dodecapeptide (VTPNDDTFDPFR) from a 12-mer peptide library; then, a biotinylated synthetic derivative of this mimotope (VTPNDDTFDPFRGGGSK-Biotin) was used for the detection of FB1 by a competitive binding inhibition assay. Its 50% inhibitory concentration (IC50) was 37.1 ng/mL, with a detection limit (LOD) of 11.1 ng/mL, and a dynamic range from 17.3 to 79.6 ng/mL. Recently, the authors used the same method to obtain a ZEN mimetic epitope peptide (GWWGPYGEIELL); the peptide was used to create fusion proteins with the bioluminescent Gaussia luciferase (GLuc) that were directly used as tracers for mycotoxin detection in a competitive immunoassay [21]. Meanwhile, using the same small peptide sequence developed a competitive upconversion-linked immunosorbent assay (ULISA) for ZEN monitoring with a LOD of 20 pg mL−1 [22].
Chen et al. [18] constructed a peptide@Tyr-RMC probe by selecting a mimotope peptide from a phage display library and labeling it a Tyr-RMC composite, which was used to develop a competitive enzyme-linked immunosorbent assay (ELISA) for the ultrasensitive detection of ZEN. It has a linear range of 10−6–1.0 ng/mL, and a low detection limit of 10−6 ng/mL. Zhao et al. [11], using AFB1 as a model system, selected a mimotope (YSWHEWYIPQLS) from Ph.D-12 phage display peptide library, and the rapid magnetic-beads-based directed competitive ELISA (MB-dcELISA) was developed by mimotope ME17. The IC50 and LOD of the MB-dcELISA were 0.75 and 0.13 ng/mL, respectively, with a linear range of 0.24–2.21 ng/mL. Zou et al. [9] obtained a mimotope from the commercial Ph.D.-7 phage display peptide library and connected it with biotin (GMVQTIF-GGGSK-biotin). The biotinylated 12-mer peptide was used as a competing antigen to develop a competitive peptide ELISA for OTA detection and showed a wide linear range of 0.005–0.2 ng/mL with the detection limit of 0.001 ng/mL. The IC50 was 0.024 ng/mL, which is approximately five times more sensitive as a competing antigen than the OTA-HRP conjugates used in the conventional ELISA .

3. Peptides as Coating Antigens

Peptides also can serve as competing antigens in immunological-based mycotoxin detection. He et al. [16] selected a phage display dodecapeptide (ESYWATVPWTRH) as a substitute for coating antigens and applied it for the rapid detection of ZEN by dot-immunoassay . The cut-off level for detecting ZEN in cereal samples was 50 mg/kg and the results can be accomplished within 10 min. With the phage display peptide library technology, Zhou et al. [23] reported a phage mimotope-based direct competitive fluorescence immunosorbent assay (P-dcFLISA); the IC50 of P-dcFLISA was 0.301 ng/mL, which was lower than the phage-based indirect competitive enzyme-linked immunosorbent assay (P-icELISA) under the same conditions. The LOD and detection range of P-dcFLISA was 0.023 ng/mL and 0.060–1.531 ng/mL, respectively. However, there is no corresponding amino acid sequence information in the article. Liu et al. [24], utilizing mimotope peptide-bovine serum albumin conjugate as a coating antigen, developed a peptide ELISA for detecting FB1, in which the IC50 and LOD were 6.06 ng/mL and 1.18 ng/mL, respectively. Except for phage display peptides, small synthetic peptides by rational design are also used as coating antigens for mycotoxin detection. Bazin et al. [34] designed an OTA-binding peptide (VYMNRKYYKCCK) derived from an oxidoreductase and developed a peptide-based competitive enzyme-linked immunosorbent assay (peptide-based competitive ELISA) in which the peptide was the coating antigen.
Mimotope-based fusion proteins can also be used as synthetic coating antigens for the detection of mycotoxins [35]. Xu et al. [25], using FB1 as a model hapten, screened two mimotopes (F1: NNAAMYSEMATD, F15: TTLQMRSEMADD) that have affinity to the anti-FB1 antibody from a 12-mer peptide library and developed a new method for the development of a sensitive and environmentally friendly immunoassay for FB1 based on the peptide–MBP (maltose-binding protein) fusion protein. Quantitative immunoassay for FB1 using F1-MBP and F15-MBP showed the LOD was 0.32 and 0.21 ng/mL, respectively, and the IC50 of the assay was 2.15 and 1.26 ng/mL, which was 10 times more sensitive than the conventional FB1-BSA conjugate-based ELISA. Meanwhile, using the same strategy, they also constructed a small peptide–MBP fusion protein-coated antigen that can be used to detect OTA and DON with good sensitivity and stability [7][15]. Recently, to monitor the co-contamination of mycotoxins in agricultural products and foods, Yan et al. [36] developed a mimotope–MBP fusion protein-based multiplex immunochromatographic assay (mICA) that can quickly and simultaneously detect FB1, ZEN, and OTA without the building-up process of mycotoxin conjugates. The LOD of peptide–MBP-based mICA for FB1, ZEN, and OTA were 0.25, 3.0, and 0.5 ng/mL, respectively.

4. Anti-Immune Complex Peptides

As small molecule contaminants (SMCs), mycotoxins usually only have one immunological binding site and are not suitable for detection by conventional sandwich non-competitive immunoassays [37]. Alternatively, the development of non-competitive immunoassays using specific recognizers for the immune complex of anti-SMC antibodies is a feasible strategy. Anti-immune complex peptide (AIcP) is a peptide that specifically binds to immune complexes, neither antibodies nor antigen monomers [38]. Biopanning the phage display random peptides library using antigen–antibody conjugates allows the screening of peptides that can bind specifically to immune complexes and can be used to establish non-competitive immunoassays [39]. Unlike mimotopes, anti-immune complex peptides recognize only complexes of antigens and antibodies and do not bind to antibodies or antigens alone, and because of this dual site recognition pattern, anti-immune complex peptides are often thought to improve the specificity of assays.
González-Sapienza’s team achieved promising results in the detection of pesticide contaminants using anti-immune complex peptides, demonstrating the advantages of the method in the detection of SMCs [31][39][40][41]. Zou et al. [10], using AFB1 and anti-AFB1 nanobody conjugates as the immune complex, screened anti-immune complex peptides from a phage display random linear 8-mer peptide library; the best binding peptide was biotinylated and coupled with horseradish peroxidase-labeled streptavidin (SA-HRP) for developing the magnetic-phage anti-immune complex immunoassay (MPHAIA) detection of AFB1. Phage-peptide p13 (DLLWVPST)-based MPHAIA, showing the lowest SC50 (50% signal saturation concentration) value (0.12 ng/mL) and the highest ODmax/SC50 ratio (12.75), was selected for further study. Under the ultimate condition, the LOD was 0.006 ng/mL, with a linear range of 0.019–0.407 ng/mL. Lassabe et al. [42] used a verotoxin (VTX) of Escherichia coli as a scaffold for multivalent display anti-immunocomplex peptides, and, among these peptides, ICX09m (CLEAPNVEAC) showed 10-fold increased sensitivity and excellent recovery than the competitive ELISA for clomazone detection.
Other studies use this method for residual pesticide and veterinary drug detection, but with few applications in mycotoxin detection. Although the non-competitive assay has advantages, the technique is still in the early stages of development, and the biggest challenge in developing this assay is the screening of anti-immune complex peptides. Therefore, improving the success rate of peptide screening for anti-immune complexes will significantly enhance the development of non-competitive immunoassays for small-molecule compounds, such as mycotoxins and pesticides. The anti-immune complex peptides of mycotoxins, pesticides, and other small molecules are summarized in Table 1.

References

  1. Levin, A.; Hakala, T.A.; Schnaider, L.; Bernardes, G.J.L.; Gazit, E.; Knowles, T.P.J. Biomimetic peptide self-assembly for functional materials. Nat. Rev. Chem. 2020, 4, 615–634.
  2. Hou, S.; Ma, J.; Cheng, Y.; Wang, Z.; Yan, Y. Overview-gold nanoparticles-based sensitive nanosensors in mycotoxins detection. Crit. Rev. Food Sci. Nutr. 2022, 62, 1–16.
  3. Wang, Y.; Zhang, C.; Wang, J.; Knopp, D. Recent progress in rapid determination of mycotoxins based on emerging biorecognition molecules: A review. Toxins 2022, 14, 73.
  4. Bazin, I.; Tria, S.A.; Hayat, A.; Marty, J.L. New biorecognition molecules in biosensors for the detection of toxins. Biosensors & Bioelectronics 2017, 87, 285–298.
  5. Liu, R.; Yu, Z.; He, Q.; Xu, Y. An immunoassay for ochratoxin A without the mycotoxin. Food Control 2007, 18, 872–877.
  6. He, Z.; He, Q.; Xu, Y.; Li, Y.; Liu, X.; Chen, B.; Lei, D.; Sun, C. Ochratoxin A Mimotope from Second-Generation Peptide Library and Its Application in Immunoassay. Anal. Chem. 2013, 85, 10304–10311.
  7. Xu, Y.; He, Z.; He, Q.; Qiu, Y.; Chen, B.; Chen, J.; Liu, X. Use of Cloneable Peptide-MBP Fusion Protein as a Mimetic Coating Antigen in the Standardized Immunoassay for Mycotoxin Ochratoxin. J. Agric. Food. Chem. 2014, 62, 8830–8836.
  8. Rahi, S.; Lanjekar, V.; Ghormade, V. Development of a rapid dot-blot assay for ochratoxin A (OTA) detection using peptide conjugated gold nanoparticles for bio-recognition and detection. Food Control 2022, 136, 108842.
  9. Zou, X.; Chen, C.; Huang, X.; Chen, X.; Wang, L.; Xiong, Y. Phage-free peptide ELISA for ochratoxin A detection based on biotinylated mimotope as a competing antigen. Talanta 2016, 146, 394–400.
  10. Zou, W.; Shi, R.; Wang, G.; Zhao, Z.; Zhao, F.; Yang, Z. Rapid and sensitive noncompetitive immunoassay for detection of aflatoxin B1 based on anti-immune complex peptide. Food Chem. 2022, 393, 133317.
  11. Zhao, F.; Tian, Y.; Shen, Q.; Liu, R.; Shi, R.; Wang, H.; Yang, Z. A novel nanobody and mimotope based immunoassay for rapid analysis of aflatoxin B1. Talanta 2019, 195, 55–61.
  12. Mukhtar, H.; Ma, L.; Pang, Q.; Zhou, Y.; Wang, X.; Xu, T.; Hammock, B.D.; Wang, J. Cyclic peptide: A safe and effective alternative to synthetic aflatoxin B1-competitive antigens. Anal. Bioanal. Chem. 2019, 411, 3881–3890.
  13. Liu, B.; Peng, J.; Wu, Q.; Zhao, Y.; Shang, H.; Wang, S. A novel screening on the specific peptide by molecular simulation and development of the electrochemical immunosensor for aflatoxin B1 in grains. Food Chem. 2022, 372, 131322.
  14. Yan, J.; Shi, Q.; You, K.; Li, Y.; He, Q. Phage displayed mimotope peptide-based immunosensor for green and ultrasensitive detection of mycotoxin deoxynivalenol. J. Pharm. Biomed. Anal. 2019, 168, 94–101.
  15. Xu, Y.; Yang, H.; Huang, Z.; Li, Y.; He, Q.; Tu, Z.; Ji, Y.; Ren, W. A peptide/maltose-binding protein fusion protein used to replace the traditional antigen for immunological detection of deoxynivalenol in food and feed. Food Chem. 2018, 268, 242–248.
  16. He, Q.; Xu, Y.; Zhang, C.; Li, Y.; Huang, Z. Phage-borne peptidomimetics as immunochemical reagent in dot-immunoassay for mycotoxin zearalenone. Food Control 2014, 39, 56–61.
  17. He, Q.; Xu, Y.; Huang, Y.-H.; Liu, R.-R.; Huang, Z.-B.; Li, Y.-P. Phage-displayed peptides that mimic zearalenone and its application in immunoassay. Food Chem. 2011, 126, 1312–1315.
  18. Chen, Y.; Zhang, S.; Hong, Z.; Lin, Y.; Dai, H. A mimotope peptide-based dual-signal readout competitive enzyme-linked immunoassay for non-toxic detection of zearalenone. J. Mater. Chem. B 2019, 7, 6972–6980.
  19. Chen, Y.; Zhang, S.; Huang, Y.; Lv, L.; Dai, H.; Lin, Y. A bio-bar-code photothermal probe triggered multi-signal readout sensing system for nontoxic detection of mycotoxins. Biosens. Bioelectron. 2020, 167, 112501.
  20. Liu, R.; Shi, R.; Zou, W.; Chen, W.; Yin, X.; Zhao, F.; Yang, Z. Highly sensitive phage-magnetic-chemiluminescent enzyme immunoassay for determination of zearalenone. Food Chem. 2020, 325, 126905.
  21. Peltomaa, R.; Fikacek, S.; Benito-Peña, E.; Barderas, R.; Head, T.; Deo, S.; Daunert, S.; Moreno-Bondi, M.C. Bioluminescent detection of zearalenone using recombinant peptidomimetic Gaussia luciferase fusion protein. Microchim. Acta 2020, 187, 547.
  22. Peltomaa, R.; Farka, Z.; Mickert, M.J.; Brandmeier, J.C.; Pastucha, M.; Hlaváček, A.; Martínez-Orts, M.; Canales, Á.; Skládal, P.; Benito-Peña, E.; et al. Competitive upconversion-linked immunoassay using peptide mimetics for the detection of the mycotoxin zearalenone. Biosens. Bioelectron. 2020, 170, 112683.
  23. Zhou, J.; Wang, X.; Li, Y.; Chen, Y.; Liu, Y.; Liu, H.; Liang, C.; Zhu, X.; Qi, Y.; Wang, A. Fluorescence immunoassay based on phage mimotope for nontoxic detection of Zearalenone in maize. J. Food Saf. 2022, 42, e12982.
  24. Liu, X.; Xu, Y.; He, Q.; He, Z.; Xiong, Z. Application of mimotope peptides of fumonisin B1 in Peptide ELISA. J. Agric. Food. Chem. 2013, 61, 4765–4770.
  25. Xu, Y.; Chen, B.; He, Q.; Qiu, Y.; Liu, X.; He, Z.; Xiong, Z. New approach for development of sensitive and environmentally friendly immunoassay for mycotoxin fumonisin B1 based on using peptide-MBP fusion protein as substitute for coating antigen. Anal. Chem. 2014, 86, 8433–8440.
  26. Peltomaa, R.; Benito-Peña, E.; Barderas, R.; Sauer, U.; González Andrade, M.; Moreno-Bondi, M.C. Microarray-based immunoassay with synthetic mimotopes for the detection of fumonisin B1. Anal. Chem. 2017, 89, 6216–6223.
  27. Peltomaa, R.; Agudo-Maestro, I.; Más, V.; Barderas, R.; Benito-Peña, E.; Moreno-Bondi, M.C. Development and comparison of mimotope-based immunoassays for the analysis of fumonisin B1. Anal. Bioanal. Chem. 2019, 411, 6801–6811.
  28. You, T.; Ding, Y.; Chen, H.; Song, G.; Huang, L.; Wang, M.; Hua, X. Development of competitive and noncompetitive immunoassays for clothianidin with high sensitivity and specificity using phage-displayed peptides. J. Hazard. Mater. 2022, 425, 128011.
  29. Fu, H.J.; Chen, Z.J.; Wang, H.; Luo, L.; Wang, Y.; Huang, R.-M.; Xu, Z.L.; Hammock, B. Development of a sensitive non-competitive immunoassay via immunocomplex binding peptide for the determination of ethyl carbamate in wine samples. J. Hazard. Mater. 2021, 406, 124288.
  30. Chen, H.; You, T.; Zong, L.; Mukhametova, L.I.; Zherdev, D.O.; Eremin, S.A.; Ding, Y.; Wang, M.; Hua, X. Competitive and noncompetitive fluorescence polarization immunoassays for the detection of benzothiostrobin using FITC-labeled dendrimer-like peptides. Food Chem. 2021, 360, 130020.
  31. González-Techera, A.; Kim, H.J.; Gee, S.J.; Last, J.A.; Hammock, B.D.; González-Sapienza, G. Polyclonal antibody-based noncompetitive immunoassay for small analytes developed with short peptide loops isolated from phage libraries. Anal. Chem. 2007, 79, 9191–9196.
  32. Yuan, Q.; Pestka, J.J.; Hespenheide, B.M.; Kuhn, L.A.; Linz, J.E.; Hart, L.P. Identification of mimotope peptides which bind to the mycotoxin deoxynivalenol-specific monoclonal antibody. Appl. Environ. Microbiol. 1999, 65, 3279–3286.
  33. Huang, D.T.; Fu, H.J.; Huang, J.J.; Luo, L.; Lei, H.T.; Shen, Y.D.; Chen, Z.J.; Wang, H.; Xu, Z.L. Mimotope-based immunoassays for the rapid analysis of mycotoxin: A review. J. Agric. Food. Chem. 2021, 69, 11743–11752.
  34. Bazin, I.; Andreotti, N.; Hassine, A.I.H.; De Waard, M.; Sabatier, J.M.; Gonzalez, C. Peptide binding to ochratoxin A mycotoxin: A new approach in conception of biosensors. Biosens. Bioelectron. 2013, 40, 240–246.
  35. Li, P.; Deng, S.; Zech Xu, Z. Toxicant substitutes in immunological assays for mycotoxins detection: A mini review. Food Chem. 2021, 344, 128589.
  36. Yan, J.X.; Hu, W.J.; You, K.H.; Ma, Z.E.; Xu, Y.; Li, Y.P.; He, Q.H. Biosynthetic mycotoxin conjugate mimetics-mediated green strategy for multiplex mycotoxin immunochromatographic assay. J. Agric. Food. Chem. 2020, 68, 2193–2200.
  37. Hua, X.; Shi, H.; Wang, M. Phage display peptide library technology and its research progress in immunoassay of pesticide residue. J. Food Saf. Qual. 2014, 5, 3955–3961.
  38. Zhao, F.; Shi, R.; Liu, R.; Tian, Y.; Yang, Z. Application of phage-display developed antibody and antigen substitutes in immunoassays for small molecule contaminants analysis: A mini-review. Food Chem. 2021, 339, 128084.
  39. Rossotti, M.A.; Carlomagno, M.; González-Techera, A.; Hammock, B.D.; Last, J.; González-Sapienza, G. Phage anti-immunocomplex assay for clomazone: Two-site recognition increasing assay specificity and facilitating adaptation into an on-site format. Anal. Chem. 2010, 82, 8838–8843.
  40. González-Techera, A.; Vanrell, L.; Last, J.A.; Hammock, B.D.; González-Sapienza, G. Phage anti-immune complex assay: general strategy for noncompetitive immunodetection of small molecules. Anal. Chem. 2007, 79, 7799–7806.
  41. Kim, H.J.; McCoy, M.; Gee, S.J.; González-Sapienza, G.; Hammock, B.D. Noncompetitive phage anti-immunocomplex real-time polymerase chain reaction for sensitive detection of small molecules. Anal. Chem. 2011, 83, 246–253.
  42. Lassabe, G.; Rossotti, M.; González-Techera, A.; González-Sapienza, G. Shiga-like toxin b subunit of escherichia coli as scaffold for high-avidity display of anti-immunocomplex peptides. Anal. Chem. 2014, 86, 5541–5546.
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: 303
Revisions: 4 times (View History)
Update Date: 24 Feb 2023
1000/1000
Video Production Service