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 + 2278 word(s) 2278 2022-02-22 03:02:13 |
2 format + 1 word(s) 2279 2022-03-11 02:00:13 |

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.
Wang, L. Regulation Mechanism of ssDNA Aptamer in Nanozymes. Encyclopedia. Available online: https://encyclopedia.pub/entry/20437 (accessed on 06 July 2024).
Wang L. Regulation Mechanism of ssDNA Aptamer in Nanozymes. Encyclopedia. Available at: https://encyclopedia.pub/entry/20437. Accessed July 06, 2024.
Wang, Lijun. "Regulation Mechanism of ssDNA Aptamer in Nanozymes" Encyclopedia, https://encyclopedia.pub/entry/20437 (accessed July 06, 2024).
Wang, L. (2022, March 10). Regulation Mechanism of ssDNA Aptamer in Nanozymes. In Encyclopedia. https://encyclopedia.pub/entry/20437
Wang, Lijun. "Regulation Mechanism of ssDNA Aptamer in Nanozymes." Encyclopedia. Web. 10 March, 2022.
Regulation Mechanism of ssDNA Aptamer in Nanozymes
Edit

Food safety issues are a worldwide concern. Pathogens, toxins, pesticides, veterinary drugs, heavy metals, and illegal additives are frequently reported to contaminate food and pose a serious threat to human health. Conventional detection methods have difficulties fulfilling the requirements for food development in a modern society. Therefore, novel rapid detection methods are urgently needed for on-site and rapid screening of massive food samples. Due to the extraordinary properties of nanozymes and aptamers, biosensors composed of both of them provide considerable advantages in analytical performances, including sensitivity, specificity, repeatability, and accuracy. They are considered a promising complementary detection method on top of conventional ones for the rapid and accurate detection of food contaminants.

nanozyme ssDNA aptamer regulation mechanism biosensor food safety

1. Introduction

Food safety has become a major global concern because of its impacts on public health as well as the international food trade. Pathogens, toxins, pesticides, veterinary drugs, heavy metals, and illegal additives are common contaminants in foods [1]. To ensure food safety, various conventional methods have been developed to determine food contaminants, including traditional plate culture methods, chromatographic methods, enzyme-linked immunosorbent assay (ELISA), and inductively coupled plasma mass spectrometry (ICP-MS). Although these methods demonstrate high accuracy and reliability for detecting food contaminants, they are complicated, time-consuming, and expensive, and they require special equipment and specially trained personnel, which are difficult to meet the requirements for to conduct on-site and rapid screening of massive food samples and are difficult to apply in many places, such as developing countries. Thus, it is urgent to develop rapid, sensitive, low-cost screening methods as complementary ways of ensuring food safety.
Recently, the fast-growing area of nanomaterials has brought a revolution in the field of analysis and detection, especially with the discovery of nanozymes [2][3]. Nanozymes usually refer to nanomaterials with enzyme-like activities, including the hydrolase and oxidoreductase families. The hydrolase family of nanozymes consists of nuclease- [4], esterase- [5], phosphatase- [6], and protease-like activity [7], while oxidoreductase activity includes peroxidase-, oxidase-, catalase-, and superoxide-dismutase-like activity [8]. Compared with natural enzymes, such as horseradish peroxidase (HRP), nanozymes have several appealing characteristics, such as low cost, high tolerance to harsh conditions (e.g., extreme pH and temperature), facile surface modification, and multifunctionality [9]. These advantages have attracted increasing research interests over the past decade, in turn elucidating the catalytic mechanism of nanozymes and facilitating the rapid development of the field of nanozymes [10][11][12][13]. So far, various nanomaterials, including noble-metal- and transition-metal-based nanomaterials [14][15][16], carbon-based nanomaterials [17][18], metal–organic framework (MOF)-based nanomaterials [19] and their hybrids, have been found to possess peroxidase-like or/and oxidase-like activity, which can be utilized to directly detect H2O2, glucose, ascorbic acid, and so on [12][20][21][22]. Achieving the specific detection of food contaminants usually requires the introduction of molecularly imprinted polymers or biorecognition elements, such as antibodies, antimicrobial peptides, bacteriophages, nucleic acid probes, and aptamers [12][23][24][25]. Among these, aptamers as alternatives to conventional antibodies are one of the most popular biorecognition elements due to their low cost, facile chemical synthesis, flexible chemical modification, and high tolerance to pH and temperature [26][27]. Furthermore, specific aptamers for almost all possible targets can be obtained via systematic evolution of ligands by exponential enrichment (SELEX). Up to now, various aptamers were obtained against a wide range of targets, from small molecules such as heavy metals to whole cells such as foodborne pathogens [28][29]. More importantly, aptamers’ complementary strands and amplification products are nucleotide sequences, which can reversibly regulate nanozyme activity [30]. These advantages endow nanozyme-based aptasensors with more diversity and promise for food safety applications, especially novel aptasensors based on the regulation by single-stranded DNA (ssDNA) of nanozyme activity.

2. Factors Affecting the Regulation by ssDNA of Nanozyme Activity

The influence of ssDNA on nanozyme activity is complicated. Initially, many research works reported that nanozyme activity was inhibited by ssDNA. As research continued, some groups demonstrated that ssDNA actually improved nanozyme activity [31]. The reasons are yet unclear, but the type of nanozyme and DNA, buffer condition, and temperature in the reaction system were found to be different in these contradicting studies. Hence, to achieve an accurate design of ssDNA-controllable nanozyme biosensors, it is requisite to clearly understand the influence of diverse factors, including intrinsic and extrinsic ones, on the regulation of nanozyme activity by ssDNA (Figure 1).
Figure 1. Diverse factors that affect the regulation of ssDNA on nanozyme activity.

2.1. Factors Affecting ssDNA Inhibition of Nanozyme Activity

Since nanozyme activity is mainly dependent on their surface properties in the solution, bioconjugation or adsorption of ssDNA on the surface of nanozymes might change the surface properties and affect their activity. In general, adsorbed ssDNA could block the surface active sites or induce the aggregation of nanozymes, thus decreasing the binding sites on nanozymes, resulting in decreased activity [32][33]. Thus, the amount of ssDNA aptamer is very important to nanozyme activity. With an increase in aptamer concentration from 2 to 10 μM, the peroxidase-like activity of graphitic carbon nitride@Cu2O (g-C3N4@Cu2O) nanocomposites gradually reduced and finally reached a plateau [34], which indicated the saturation of aptamer on the surface of g-C3N4@Cu2O when the amount of aptamer increased to 10 μM. Similar variation trends were observed using Hemin@MOFs nanozyme [35], gold nanoparticles (AuNPs) nanozyme [36], and MnCo2O4 nanozyme [37]. Different nucleotide types demonstrated different inhibition ability towards nanozyme activity. The order of inhibiting effects towards MnCo2O4 nanozyme activity was: A > G > C > T [37]. For Mn3O4 nanozyme, purine nucleotide (A, G) had more active groups to react with octahedral Mn3O4 nanozymes for blocking surface active sites than pyrimidine nucleotide (C, T), demonstrating a stronger inhibitory effect on the oxidase-like activity of Mn3O4 [38]. Theoretically, longer ssDNA would have more binding sites to nanozymes and exhibit a better inhibition effect. In fact, 500 nM of poly A15 was enough to inhibit the oxidase-like activity of 10 μg/mL MnCo2O4. Increasing the number of nucleotides from 15 to 45 did not cause an obvious change in MnCo2O4 nanozyme activity, indicating that a saturation of ssDNA was reached on the MnCo2O4’s surface [37].
Wang et al. [39] found that double-stranded DNA (dsDNA) did not have the ability to inhibit the oxidase-like activity of a mixed-valence state cerium-based metal–organic framework (MVC-MOF). The difference in ability of ssDNA and dsDNA could be ascribed to the selective adsorption of ssDNA by MVC-MOF [40]. ssDNA has a flexible structure, which contributes to strong π–π stacking interaction between the bases of ssDNA and MOF. Contrarily, dsDNA has a stiff and rigid double-helix structure and cannot wrap around nanomaterials effectively [35], causing low affinity between dsDNA and nanomaterials. However, when the amount of dsDNA was far more than that of ssDNA, the inhibition effect of the former was comparable to or exceeded that of the latter [41][42].
The shape of the nanozyme also affects the interaction of ssDNA and nanozymes. For example, compared with other shapes of Mn3O4 nanoparticles (i.e., flower-like, polyhedron, spinel), octahedral Mn3O4 nanozymes had more regular surfaces to interact with ssDNA, resulting in a significant inhibition of oxidase-like activity by ssDNA [38]. Adsorption is the first and key step for realizing DNA regulation of nanozyme activity [43][44]. Different nanomaterials have different kinds of adsorption force to ssDNA [45]. For example, nanoceria adsorbed onto the ssDNA through electrostatic interaction and Lewis acid–base interaction between the phosphate backbone of ssDNA and cerium [33]. The interaction between DNA and AuNPs was tuned by the combination of various noncovalent forces, including electrostatic interactions, hydrophobic forces, and specific bonds between the chemical groups of DNA bases and gold [46]. In addition, the main adsorption force between graphene and ssDNA was π–π stacking interaction [47]. Besides nanozyme type, extrinsic conditions affecting the adsorption forces play an important role in nanozyme activity, including pH and ionic strength. For example, the optimum catalytic condition for MnCo2O4 nanozyme was pH 3.0, while the nanozyme activity can be maximally inhibited by ssDNA at pH 4.0 [37]. Increasing the ionic strength of buffer weakened the inhibition ability of ssDNA toward MnCo2O4 nanozyme and enhanced the dissociation ability of ssDNA from the nanozyme’s surface [37].

2.2. Factors Affecting ssDNA to Improve Nanozyme Activity

With helps from nanotechnology advancement, researchers have found that ssDNA could enhance the catalytic activity of nanomaterials, including noble-metal-based nanomaterials [31], transition-metal-based nanomaterials, carbon-based nanomaterials [48], MOFs, and their hybrids. As mentioned above, the adsorption of ssDNA onto the surface of nanomaterials was a prerequisite for ssDNA to regulate the nanozyme activity. For example, ssDNA on the surface of nanozymes could adsorb more substrate (e.g., 3,3′,5,5′-tetramethylbenzidine (TMB)) around the nanozyme due to electrostatic attraction and π–π stacking interaction between TMB and ssDNA, further accelerating electron transfer from TMB to the target sensing molecules (e.g., H2O2) and enhancing the nanozyme activity [43][49]. Thus, factors affecting both ssDNA–nanomaterial and ssDNA–TMB affinity are important for the enhancement of nanozyme activity by ssDNA. The usual trend of nanozyme activity was to first increase and then decrease with the increase in ssDNA concentration [50][51][52]. However, some data support a different variation trend of nanozyme activity, which is to first increase and then keep stable with the increase of ssDNA concentration [44][48][49][53][54]. The reason for the two different trends was unclear. The researchers speculate that the presence of excessive ssDNA may hinder the contact of the nanozyme and TMB in the former trend since it had a wide range of ssDNA concentration. DNA molecules with different structures were also reported to affect the peroxidase-like activity of nanozymes. The order of enhancement of the ability of catalytic activity was: hybridization chain reaction (HCR) products (long dsDNA) > hairpin DNA > ssDNA > dsDNA [51]. Moreover, different DNA nucleotides had different abilities to enhance the nanozyme activity. The order of enhancement of Fe3O4 nanozyme activity was: C > G > T > A [43]. Cytosine in buffer at pH 4.0 was protonated, which could assist charge neutralization on the surface of Fe3O4 nanoparticles and reduce repulsion among DNA, leading to the adsorption of more DNA. The same variation trend was observed by Wang et al. [48] using protonated graphitic carbon nitride (Pg-C3N4) nanosheets as nanozymes. The peroxidase-like activity of 20 bp homo ssDNA (A20, C20, G20, and T20) modified MoS2 nanozymes followed the trend of G20 ≈ T20 > A20 > C20 > no DNA. The explanation from another perspective was given for the weakest activity of C20-modified MoS2 nanozymes. Protonated cytosine nucleotides in buffer at pH 4.0 increased the electrostatic repulsion between cytosine and positively charged TMB, resulting in a lower affinity of MoS2 nanozymes [49]. Guanine (G) was also reported to significantly enhance the peroxidase-like activity of WS2 nanosheets [55], iron-based MOFs modified with acidized carbon nanotubes (MOF/CNTs) [50], and MIL-53(Fe) [44]. Purine (A, G) modification demonstrated a remarkable enhancement of the peroxidase-like activity of AuNPs, while pyrimidine (T, C) modification enhanced it slightly, which was attributed to the difference in the interaction between TMB and the surface-adsorbed nucleobases [56]. In addition, the effect of DNA length on the enhancement of nanozyme activity was also investigated. The longest DNA (ploy C30) demonstrated the largest enhancement of peroxidase-like activity of Fe3O4 nanoparticles due to the presence of more binding sites [43]. Increasing the number of cytosine nucleotides from 30 to 40, the peroxidase-like activity of Pg-C3N4 nanosheets kept steady, which indicated the saturation of ssDNA on the surface of the Pg-C3N4 nanosheets [48]. A slight inhibition of the catalytic activity of MoS2 nanozymes was found when the number of thymidine nucleotides increased from 30 to 40 [49].
Besides the intrinsic factors mentioned above, the reaction conditions were also very important for DNA to regulate the nanozyme activity. The buffer pH at 4.0 had been found to be the optimum condition for both the catalytic activity of nanozymes and the inhibition of nanozyme activity by ssDNA [43][44][54][55]. TMB was the most-used substrate in those literatures. It carries positive charges at pH 4.0 that is below its pKa, which is beneficial to the reaction between ssDNA and TMB [43]. However, continuously reducing the pH resulted in the protonation of two amino acids of TMB, which made them insusceptible to oxidation [57]. Moreover, buffer type and concentration were also reported to affect the ability of ssDNA to enhance the peroxidase-like activity of nanozymes. In a phosphate buffer, low-concentration DNA moderately increased the peroxidase-like activity of CeO2, while DNA in acetate buffer had no effect on the catalytic activity of CeO2, except for the high-concentration DNA (more than 10 µM), which inhibited the CeO2 nanozyme activity [58]. This was because phosphate could compete with ssDNA to bind to CeO2 nanozymes and enhance their peroxidase-like activity. Low ionic strength increased the peroxidase-like activity of ssDNA-modified WS2 nanosheets [55], while high ionic strength could shield the electrostatic reaction, which reduced the interaction among ssDNA, nanozymes, and TMB [59].
More amazingly, the enhancement or inhibition of the peroxidase-like activity of AuNPs by ssDNA can be controlled by H2O2 concentration. In a 10 mM H2O2 reaction system, DNA inhibited the peroxidase-like activity of AuNPs in the first 10 min and enhanced the catalytic activity of AuNPs in the next 20 min of the reaction time, while at a lower concentration of H2O2 (5 mM), it prolonged the inhibition time to 30 min and enhancement occurred after 30 min [56]. Moreover, by changing the substrate from the positively charged TMB to the negatively charged 2,2′-azino-bis (3-ethyl benzothiazoline-6-sulfonic acid) (ABTS), the enhanced ability of ssDNA disappeared, and the inhibiting ability of ssDNA appeared [43][55].
As previously mentioned, ssDNA can either enhance or inhibit the catalytic activity of nanozymes, and the deciding factor for enhancement or inhibition was to increase or decrease the affinity of substrate (e.g., TMB or ABTS) and nanozymes, respectively, by the introduction of ssDNA. A small change of these factors caused the interaction of ssDNA and nanozymes to be altered, which resulted in variations of the nanozyme activity. Thus, the process of designing aptamer-assisted nanozyme sensing needs to be thoughtful since various factors can affect the regulation of ssDNA on nanozyme activities.

References

  1. Han, Y.; Yang, W.; Luo, X.; He, X.; Zhao, H.; Tang, W.; Yue, T.; Li, Z. Carbon dots based ratiometric fluorescent sensing platform for food safety. Crit. Rev. Food Sci. 2020, 62, 244–260.
  2. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583.
  3. Das, B.; Franco, J.L.; Logan, N.; Balasubramanian, P.; Kim, M.I.; Cao, C. Nanozymes in point-of-care diagnosis: An emerging futuristic approach for biosensing. Nano-Micro Lett. 2021, 13, 193.
  4. Czescik, J.; Zamolo, S.; Darbre, T.; Rigo, R.; Sissi, C.; Pecina, A.; Riccardi, L.; De Vivo, M.; Mancin, F.; Scrimin, P. A gold nanoparticle nanonuclease relying on a Zn(II) mononuclear complex. Angew. Chem. Int. Ed. 2021, 60, 1423–1432.
  5. Nandhakumar, P.; Kim, G.; Park, S.; Kim, S.; Kim, S.; Park, J.K.; Lee, N.S.; Yoon, Y.H.; Yang, H. Metal nanozyme with ester hydrolysis activity in the presence of ammonia-borane and its use in a sensitive immunosensor. Angew. Chem. Int. Ed. 2020, 59, 22419–22422.
  6. Hu, X.; Huang, T.; Liao, H.; Hu, L.; Wang, M. The phosphatase-like activity of zirconium oxide nanoparticles and their application in near-infrared intracellular imaging. J. Mater. Chem. B 2020, 8, 4428–4433.
  7. Li, B.; Chen, D.; Nie, M.; Wang, J.; Li, Y.; Yang, Y. Carbon dots/Cu2O composite with intrinsic high protease-Like activity for hydrolysis of proteins under physiological conditions. Part. Part. Syst. Char. 2018, 35, 1800277.
  8. Zhang, Y.A.; Jin, Y.L.; Cui, H.X.; Yan, X.Y.; Fan, K.L. Nanozyme-based catalytic theranostics. Rsc Adv. 2020, 10, 10–20.
  9. Wang, Q.; Wei, H.; Zhang, Z.; Wang, E.; Dong, S. Nanozyme: An emerging alternative to natural enzyme for biosensing and immunoassay. Trends Anal. Chem. 2018, 105, 218–224.
  10. Yang, W.; Yang, X.; Zhu, L.; Chu, H.; Li, X.; Xu, W. Nanozymes: Activity origin, catalytic mechanism, and biological application. Coordin. Chem. Rev. 2021, 448, 214170.
  11. Wang, W.Z.; Gunasekaran, S. Nanozymes-based biosensors for food quality and safety. Trends Anal. Chem. 2020, 126, 115841.
  12. Wu, J.J.X.; Wang, X.Y.; Wang, Q.; Lou, Z.P.; Li, S.R.; Zhu, Y.Y.; Qin, L.; Wei, H. Nanomaterials with enzyme-like characteristics (nanozymes): Next-generation artificial enzymes (II). Chem. Soc. Rev. 2019, 48, 1004–1076.
  13. Wang, P.; Min, D.; Chen, G.; Li, M.; Tong, L.; Cao, Y. Inorganic Nanozymes: Prospects for Disease Treatments and Detection Applications. Front. Chem. 2021, 9, 773285.
  14. Navyatha, B.; Singh, S.; Nara, S. AuPeroxidase nanozymes promises and applications in biosensing. Biosens. Bioelectron. 2021, 175, 112882.
  15. Ye, M.L.; Zhu, Y.; Lu, Y.; Gan, L.; Zhang, Y.; Zhao, Y.G. Magnetic nanomaterials with unique nanozymes-like characteristics for colorimetric sensors: A review. Talanta 2021, 230, 122299.
  16. Liu, Q.; Zhang, A.; Wang, R.; Zhang, Q.; Cui, D. A review on metal- and metal oxide-based nanozymes: Properties, mechanisms, and applications. Nano-Micro Lett. 2021, 13, 154.
  17. Jin, J.; Li, L.; Zhang, L.; Luan, Z.; Xin, S.; Song, K. Progress in the application of crbon dots-based nanozymes. Front. Chem. 2021, 9, 748044.
  18. Sun, H.; Zhou, Y.; Ren, J.; Qu, X. Carbon nanozymes: Enzymatic properties, catalytic mechanism, and applications. Angew. Chem. Int. Ed. 2018, 57, 9224–9237.
  19. Huang, X.; Zhang, S.; Tang, Y.; Zhang, X.; Bai, Y.; Pang, H. Advances in metal-organic framework-based nanozymes and their applications. Coordin. Chem. Rev. 2021, 449, 214216.
  20. Jin, S.; Wu, C.; Ye, Z.; Ying, Y. Designed inorganic nanomaterials for intrinsic peroxidase mimics: A review. Sens. Actuator B Chem. 2019, 283, 18–34.
  21. Wang, Q.; Liu, S.; Tang, Z. Recent progress in the design of analytical methods based on nanozymes. J. Mater. Chem. B 2021, 9, 8174–8184.
  22. Wei, H.; Wang, E. Fe3O4 magnetic nanoparticles as peroxidase mimetics and their applications in H2O2 and glucose detection. Anal. Chem. 2008, 80, 2250–2254.
  23. Shen, Y.; Xu, L.; Li, Y. Biosensors for rapid detection of Salmonella in food: A review. Compr. Rev. Food Sci. Food Saf. 2021, 20, 149–197.
  24. Tao, X.; Wang, X.; Liu, B.; Liu, J. Conjugation of antibodies and aptamers on nanozymes for developing biosensors. Biosens. Bioelectron. 2020, 168, 112537.
  25. Yan, M.M.; Chen, G.; She, Y.X.; Ma, J.; Hong, S.H.; Shao, Y.; Abd El-Aty, A.M.; Wang, M.; Wang, S.S.; Wang, J. Sensitive and simple competitive biomimetic nanozyme-linked immunosorbent assay for colorimetric and surface-enhanced raman scattering sensing of triazophos. J. Agric. Food Chem. 2019, 67, 9658–9666.
  26. Majdinasab, M.; Hayat, A.; Marty, J.L. Aptamer-based assays and aptasensors for detection of pathogenic bacteria in food samples. Trends Anal. Chem. 2018, 107, 60–77.
  27. Wang, L.J.; Wang, R.H.; Wei, H.; Li, Y.B. Selection of aptamers against pathogenic bacteria and their diagnostics application. World J. Microb. Biot. 2018, 34, 149.
  28. Dong, Y.Y.; Xu, Y.; Yong, W.; Chu, X.G.; Wang, D.N. Aptamer and its potential applications for food safety. Crit. Rev. Food Sci. 2014, 54, 1548–1561.
  29. Song, S.-H.; Gao, Z.-F.; Guo, X.; Chen, G.-H. Aptamer-based detection methodology studies in food safety. Food Anal. Methods 2019, 12, 966–990.
  30. Wang, X.; Xu, Y.; Cheng, N.; Wang, X.; Huang, K.; Luo, Y. Recent advances in nucleic acid modulation for functional nanozyme. Catalysts 2021, 11, 638.
  31. Chang, Y.; Gao, S.; Liu, M.; Liu, J. Designing signal-on sensors by regulating nanozyme activity. Anal. Methods 2020, 12, 4708–4723.
  32. Sharma, T.K.; Ramanathan, R.; Weerathunge, P.; Mohammadtaheri, M.; Daima, H.K.; Shuklaa, R.; Bansala, V. Aptamer-mediated ‘turn-off turn-on’ nanozyme activity of gold nanoparticles for kanamycin detection. Chem. Commun. 2014, 50, 15856–15859.
  33. Pautler, R.; Kelly, E.Y.; Huang, P.J.; Cao, J.; Liu, B.; Liu, J. Attaching DNA to nanoceria: Regulating oxidase activity and fluorescence quenching. ACS Appl. Mater. Interfaces 2013, 5, 6820–6825.
  34. Tarokh, A.; Pebdeni, A.B.; Othman, H.O.; Salehnia, F.; Hosseini, M. Sensitive colorimetric aptasensor based on g-C3N4@Cu2O composites for detection of Salmonella typhimurium in food and water. Microchim. Acta 2021, 188, 87.
  35. Liu, S.; Yang, M.; Guo, W. Programmable and reversible regulation of activities with DNA structures. Chem. Res. Chin. Univ. 2020, 36, 301–306.
  36. Weerathunge, P.; Ramanathan, R.; Torok, V.A.; Hodgson, K.; Xu, Y.; Goodacre, R.; Behera, B.K.; Bansal, V. Ultrasensitive colorimetric detection of murine norovirus using nanozyme aptasensor. Anal. Chem. 2019, 91, 3270–3276.
  37. Huang, L.J.; Chen, K.; Zhang, W.T.; Zhu, W.X.; Liu, X.N.; Wang, J.; Wang, R.; Hu, N.; Suo, Y.R.; Wang, J.L. ssDNA-tailorable oxidase-mimicking activity of spinel MnCo2O4 for sensitive biomolecular detection in food sample. Sens. Actuator B Chem. 2018, 269, 79–87.
  38. Wang, J.; Wang, J.; Zhou, P.; Tao, H.; Wang, X.; Wu, Y. Oligonucleotide-induced regulation of the oxidase-mimicking activity of octahedral Mn3O4 nanoparticles for colorimetric detection of heavy metals. Microchim. Acta 2020, 187, 99.
  39. Wang, C.; Tang, G.; Tan, H. Colorimetric determination of mercury(II) via the inhibition by ssDNA of the oxidase-like activity of a mixed valence state cerium-based metal-organic framework. Microchim. Acta 2018, 185, 475.
  40. Wang, G.; Song, C.; Kong, D.; Ruan, W.; Chang, Z.; Li, Y. Two luminescent metal-organic frameworks for the sensing of nitroaromatic explosives and DNA strands. J. Mater. Chem. A 2014, 2, 2213–2220.
  41. Wang, L.; Liao, T.; Zhou, H.; Huang, Y.; Chen, P.; Yang, X.; Chen, X. Colorimetric method for Salmonella spp. detection based on peroxidase-like activity of Cu(II)-rGO nanoparticles and PCR. Anal. Biochem. 2021, 615, 114068.
  42. Kim, M.I.; Park, K.S.; Park, H.G. Ultrafast colorimetric detection of nucleic acids based on the inhibition of the oxidase activity of cerium oxide nanoparticles. Chem. Commun. 2014, 50, 9577–9580.
  43. Liu, B.; Liu, J. Accelerating peroxidase mimicking nanozymes using DNA. Nanoscale 2015, 7, 13831–13835.
  44. Song, C.; Ding, W.; Liu, H.; Zhao, W.; Yao, Y.; Yao, C. Label-free colorimetric detection of deoxyribonuclease I activity based on the DNA-enhanced peroxidase-like activity of MIL-53(Fe). New J. Chem. 2019, 43, 12776–12784.
  45. Lopez, A.; Zhang, Y.; Liu, J. Tuning DNA adsorption affinity and density on metal oxide and phosphate for improved arsenate detection. J. Coll. Interfaces Sci. 2017, 493, 249–256.
  46. Koo, K.M.; Sina, A.A.I.; Carrascosa, L.G.; Shiddiky, M.J.A.; Trau, M. DNA-bare gold affinity interactions: Mechanism and applications in biosensing. Anal. Methods 2015, 7, 7042–7054.
  47. Liu, M.; Zhao, H.; Chen, S.; Yu, H.; Quan, X. Interface engineering catalytic graphene for smart colorimetric biosensing. ACS Nano 2012, 6, 3142–3151.
  48. Wang, L.; Zhu, F.; Liao, S.; Chen, M.; Zhu, Y.Q.; Liu, Q.; Chen, X. Single-stranded DNA modified protonated graphitic carbon nitride nanosheets: A versatile ratiometric fluorescence platform for multiplex detection of various targets. Talanta 2019, 197, 422–430.
  49. Zhao, L.; Wang, J.; Su, D.; Zhang, Y.; Lu, H.; Yan, X.; Bai, J.; Gao, Y.; Lu, G. The DNA controllable peroxidase mimetic activity of MoS2 nanosheets for constructing robust colorimetric biosensor. Nanoscale 2020, 12, 19420–19428.
  50. Sun, S.H.; Fan, Y.F.; Du, J.Y.; Song, Z.L.; Zhao, H.M. CNT-modified MIL-88 (NH2)-Fe for enhancing DNA-regulated peroxidase-like activity. J. Anal. Test. 2019, 3, 238–245.
  51. Zeng, C.; Lu, N.; Wen, Y.; Liu, G.; Zhang, R.; Zhang, J.; Wang, F.; Liu, X.; Li, Q.; Tang, Z.; et al. Engineering nanozymes using DNA for catalytic regulation. ACS Appl. Mater. Inter. 2019, 11, 1790–1799.
  52. Zhang, L.; Qi, Z.; Zou, Y.; Zhang, J.; Xia, W.; Zhang, R.; He, Z.; Cai, X.; Lin, Y.; Duan, S.; et al. Engineering DNA-nanozyme interfaces for rapid detection of dental bacteria. ACS Appl. Mater. Inter. 2019, 11, 30640–30647.
  53. Wang, Y.; Liu, J.; Adkins, G.B.; Shen, W.; Trinh, M.P.; Duan, L.; Jiang, J.; Zhong, W. Enhancement of the intrinsic peroxidase-like activity of graphitic carbon nitride nanosheets by ssDNAs and its application for detection of exosomes. Anal. Chem. 2017, 89, 12327–12333.
  54. Zhu, X.; Tang, L.; Wang, J.; Peng, B.; Ouyang, X.; Tan, J.; Yu, J.; Feng, H.; Tang, J. Enhanced peroxidase-like activity of boron nitride quantum dots anchored porous CeO2 nanorods by aptamer for highly sensitive colorimetric detection of kanamycin. Sens. Actuator B Chem. 2021, 330, 129318.
  55. Tang, Y.; Hu, Y.; Zhou, P.; Wang, C.; Tao, H.; Wu, Y. Colorimetric detection of kanamycin residue in foods based on the aptamer-enhanced peroxidase-mimicking activity of layered WS2 nanosheets. J. Agric. Food Chem. 2021, 69, 2884–2893.
  56. Hizir, M.S.; Top, M.; Balcioglu, M.; Rana, M.; Robertson, N.M.; Shen, F.; Sheng, J.; Yigit, M.V. Multiplexed activity of perAuxidase: DNA-capped AuNPs act as adjustable peroxidase. Anal. Chem. 2016, 88, 600–605.
  57. Drozd, M.; Pietrzak, M.; Parzuchowski, P.G.; Malinowska, E. Pitfalls and capabilities of various hydrogen donors in evaluation of peroxidase-like activity of gold nanoparticles. Anal. Bioanal. Chem. 2016, 408, 8505–8513.
  58. Zhao, Y.; Li, H.; Lopez, A.; Su, H.; Liu, J. Promotion and inhibition of the oxidase-mimicking activity of nanoceria by phosphate, polyphosphate and DNA. ChemBioChem 2020, 21, 2178–2186.
  59. Zhang, F.; Wang, S.; Liu, J. Gold nanoparticles adsorb DNA and aptamer probes too strongly, and a comparison with graphene oxide for biosensing. Anal. Chem. 2019, 91, 14743–14750.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 446
Revisions: 2 times (View History)
Update Date: 11 Mar 2022
1000/1000
Video Production Service