Skip to main content

Applications of nanomaterials with enzyme-like activity for the detection of phytochemicals and hazardous substances in plant samples

Abstract

Plants such as herbs, vegetables, fruits, and cereals are closely related to human life. Developing effective testing methods to ensure their safety and quantify their active components are of significant importance. Recently, nanomaterials with enzyme-like activity (known as nanozymes) have been widely developed in various assays, including colorimetric, fluorescence, chemiluminescence, and electrochemical analysis. This review presents the latest advances in analyzing phytochemicals and hazardous substances in plant samples based on nanozymes, including some active ingredients, organophosphorus pesticides, heavy metal ions, and mycotoxins. Additionally, the current shortcomings and challenges of the actual sample analysis were discussed.

Background

Plants, such as Chinese herbal medicines (CHMs) and edible plants, play crucial roles in human life [1]. The bioactive components in plants, namely phytochemicals, are important for human health and disease prevention. They usually have antioxidant capacity that protects cells from oxidative stress, as well as anti-inflammatory, antibacterial, and anti-tumor activities to prevent diseases like cancer, inflammatory bowel disease, and metabolic syndrome [2,3,4]. To date, various analytical techniques have already been used to analyze phytochemicals, including high-performance liquid chromatography (HPLC), mass spectrometry (MS), capillary electrophoresis (CE), gas chromatography (GC) [5, 6], etc. Although these methods are accurate and sensitive, sample handling is sophisticated with long analysis time and high cost. Thus, the development of novel, rapid, and selective approaches for the analysis of phytochemicals in plants and their extracts is of great importance for quick and on-site detection.

Furthermore, plants are frequently exposed to a variety of chemicals that are harmful to the human body, impacting their usability. For instance, inappropriate discharge of industrial wastewater leads to the accumulation of heavy metal ions in the soil, which will inevitably be absorbed by plants during planting. Excessive intake of heavy metal ions is prone to cause neurological disorders, kidney and liver damage, cardiovascular disease, and cancer [7, 8]. Therefore, the Chinese Pharmacopoeia (2020 edition) has set limits for heavy metals in Chinese medicinal materials and tablets of plant species: arsenic (2 mg/kg), cadmium (1 mg/kg), copper (20 mg/kg) lead (5 mg/kg), and mercury (0.2 mg/kg) [9]. In addition, organophosphorus pesticides (OPs) are widely used to protect plants from pests during cultivation, giving rise to the presence of excessive pesticide residues [10]. Once entering the human body, they irreversibly inhibit cholinesterase activity, posing a hazard to the cardiovascular, nervous, and respiratory systems [11]. In response, the Ministry of Agriculture and Rural Affairs of China has established the maximum residue limits (MRLs) of OPs in food. For example, apples’ MRLs of dichlorvos, glyphosate, and chlorpyrifos are 0.1, 0.5, and 1 mg/kg, respectively (GB 2763–2021) [12]. Furthermore, it should also be noted that plants may become contaminated by certain toxins during storage, such as mycotoxins. These toxins are highly carcinogenic and difficult to be completely removed during processing because of their thermal stability [13]. The current analytical methods for detecting heavy metal ions, OPs, and mycotoxins, including HPLC, GC–MS, atomic absorption spectrometry (AAS), inductively coupled plasma mass spectrometry (ICP-MS), and ultra-performance liquid chromatography-tandem mass spectrometry (UPLC-MS/MS), require complex pretreatment and specialized operations [8, 10, 14]. Therefore, there is a necessity to develop simple, efficient, and sensitive methods to detect these hazardous substances in plants for on-site and rapid inspection.

Nanozymes are a type of nanomaterials that exhibit enzyme-like activities, including metals (Au, Ag, Pt, Pd, etc.), metal oxides (CeO2, Fe3O4, Mn3O4, CuO, etc.), carbon-based compounds (carbon nanotubes, graphitic carbon nitride, carbon dots, etc.), and other nanomaterials (e.g., metal–organic frameworks (MOFs), covalent organic frameworks (COFs), metal sulfides, etc.) [15,16,17]. The enzyme-like activity of currently reported nanozymes can be mainly divided into two categories, oxidoreductase such as peroxidase (POD), oxidase (OXD), laccase (LAC), and superoxide dismutase (SOD), and hydrolase such as nuclease, esterase, phosphatase, and protease. Table 1 summarizes the functions of commonly reported nanozymes. Nanozymes have been widely used in the field of biomedicine, environmental monitoring, and food safety for their remarkable advantages of high stability, low cost, controllable activity, and easy storage [18, 19]. Most significantly, the nanozyme-based sensor offers the advantage of a shorter detection time that can fulfill the requirements of real-time detection [20,21,22]. For instance, Wang et al. [23] developed a manganese-based nanozyme that enabled rapid quantitative analysis of glutathione within 1 min. Xu et al. [24] synthesized copper-cobalt bimetallic nanozymes and combined with a smartphone and hydrogel kit to achieve real-time monitoring of perfluorooctane sulfonate (PFOS) in lake water. The approach offers a simpler instrument and quicker build-up compared to traditional methods like HPLC. Herein, this review aims to summarize recent advancements in applying nanomaterials with enzyme-like activity to detect phytochemicals and hazardous substances in plants (Fig. 1). Firstly, the application of nanozymes for detecting active phytochemicals was introduced, including gallic acid, tannic acid, ascorbic acid, rutin, atropine, quercetin, astragaloside-IV, and licorice. Secondly, advancements in the utilization of nanozymes for detecting hazardous substances in plants were presented, such as organophosphorus pesticides, heavy metal ions, and mycotoxins. Finally, the challenges and prospects in nanozyme-based detection of plant samples were discussed. This paper may provide useful information for readers to understand the design, performance, and application of nanozymes, to develop efficient, rapid, highly sensitive, and selective methods for detecting target components in actual plant samples.

Table 1 Summary of functions of currently reported nanozymes
Fig. 1
figure 1

Review of nanozymes-based detection of phytochemicals and hazardous substances in plants

Detection of phytochemicals

Phytochemicals are biologically active secondary metabolites produced by plants for self-protection, including carotenoids, polyphenols, alkaloids, saponins, and others [25], which are obtainable from a variety of sources, such as herbs, vegetables, fruits, and teas [26]. Most of them exhibit potent antioxidant activity and contribute to reducing the risks of heart disease, cancer, diabetes, and other diseases [25, 27]. However, plants are intricate systems containing a multitude of substances, making it challenging to achieve specific analysis and identification of the target phytochemicals. Table 2 summarizes some of the studies on the detection of phytochemicals in plants by nanozyme-based methods.

Table 2 Summary of detection of phytochemicals in plants based on nanozymes

Direct detection

Gallic acid (GA) and tannic acid (TA) are a class of natural phenolic compounds widely found in fruits and teas with various biological activities, such as antioxidant, anticancer, anti-mutagenesis, and antiviral [28,29,30,31]. The nanomaterials with POD-like and OXD-like activities can catalyze the oxidation of the substrate 3, 3’, 5, 5’-tetramethylbenzidine (TMB) to generate blue oxidized TMB (ox-TMB) in the presence of H2O2 and O2, respectively. GA and TA can inhibit the oxidation of TMB due to their antioxidant property, realizing colorimetric detection of them. Besides, instead of complicated pretreatment, these active ingredients can be directly detected in the real samples through a simple water extraction. Perovskite is a type of transition metal oxide, some of which possess splendid catalytic activity. Chen et al. [29] developed a simple colorimetric method to detect GA based on the POD-like activity of LaFeO3 microspheres, which is a typical perovskite, with a linear range of 0.67–40.8 µM and a limit of detection (LOD) of 0.4 µM. In addition, the established method was used in the determination of GA in diet tea, green tea, and pharyngitis tablets with good recoveries of 95.65–102.10% and RSD (n = 3) less than 4.00%. The activity of nanozymes plays a vital role in detecting phytochemicals, which can affect the detection sensitivity. Combining carbon-based materials with perovskite can enhance their catalytic performance. Liu et al. [32] synthesized the heterojunctions composed of strontium titanate (SrTiO3) and reduced graphene oxide (rGO), which facilitate photo-generated charge transfer under ultraviolet irradiation, resulting in an excellent POD-like activity. It is noted that the affinity for TMB of SrTiO3-rGO composites is 19 times higher than that of natural horseradish peroxidase (HRP). Meanwhile, the colorimetric quantitative detection of TA shows a lower LOD of 0.056 µM, which has been successfully applied to detect TA in green tea and Oolong tea.

Atomic doping is also one of the methods to enhance enzyme-like activity. Furthermore, oxygen vacancies (OVs) are a kind of metal oxide defects, which are formed by the detachment of oxygen from the lattice of metal oxides in a specific external environment (e.g., high temperature). OVs can provide rich active sites and high surface energy to improve the catalytic activity of nanozymes [33]. Zhou et al. [28] prepared a raspberry-like nitrogen-doped Mn3O4 nanospheres (N-Mn3O4 NSps) with OVs, which exhibited enhanced OXD-like activity (Fig. 2). The senor based on N-Mn3O4 NSps showed excellent reproducibility, stability, and interference resistance for detecting GA with a linear range of 5–30 µM and a lower LOD of 0.028 µM, which is feasible for the detection of GA both in green tea and black tea with the RSD (n = 5) within 3.27%. Furthermore, a platform based on the smartphone was implemented for GA detection with a LOD of 0.047 µM.

Fig. 2
figure 2

Schematic diagram of detecting GA based on N-Mn3O4 NSs. Reprinted with permission from [28]

In addition, the electrochemical assay has also been exploited for the quantification of GA. Based on gold nanoparticles, Chen et al. [34] developed a peptide-modified dual mimetic enzyme sensor for the detection of GA. The construction mechanism relied on the active center of the methanobactin (Mb) structure that can capture Cu(II), resulting in the coordinated complex Mb(CuII) with polyphenol oxidase (PPO)- and POD-like activities. After the addition of GA, the sensor with surface-modified gold nanoparticles and Mb(CuII) exhibited a high oxidation peak with a peak potential of 0.79 ± 0.05 V. Subsequently, the developed method was employed to detect GA in three real samples, including grapes, oranges, and black tea, with recoveries of 96.76–100.95% and RSD (n = 3) less than 5%.

To explore additional response mechanisms is an effective approach to improve the detection selectivity. The 2,3,6,7,10,11-Hexahydroxytriphenylene (HHTP) is a highly conjugated triol ester that can coordinate with a metal-based node to form two-dimensional porous expansion frameworks known as metal catecholates (M-CATs). Inspired by the structure of M-CATs, Wu et al. [31] prepared a Fe-HHTP amorphous nanomaterial with POD-like activity through an one-step self-assembly strategy. The colorimetric method based on Fe-HHTP can rapidly detect TA within the linear range of 0.5–100 µM with a LOD of 0.5 µM, which was successfully used to measure TA content in tea and red wine samples. Remarkably, the inhibition of TA on the color reaction was resulted not only from its antioxidant ability but also from the formation of a Fe3+-TA complex. However, GA and AA still exhibited certain interference on the detection of TA. To further address this issue, Wu et al. [35] developed a colorimetric/photothermal dual-mode analysis method for TA detection based on the light-enhanced POD-like activity and high photothermal property of CuS hollow nanocages (CuS HNCs) (Fig. 3). TA inhibited the oxidation of TMB and effectively captured the thermal holes generated by CuS HNCs under NIR irradiation, which suppressed the reaction system’s photothermal effect. The established method exhibited better selectivity and higher interference resistance from GA and AA.

Fig. 3
figure 3

Schematic mechanism for dual-mode detection of TA based on the CuS HNC probe. Reprinted with permission from [35]

However, most of the above methods failed to identify GA or TA with absolute specificity because of the interference of other antioxidants in the samples. Alternatively, similar methods were applied to detect total antioxidant capacity (TAC) in the actual samples. He et al. [36] designed a Pd–Pt-Ru nanozyme with good POD-like activity, which was used to detect ascorbic acid (AA) and H2O2 in the ranges of 2–12 µM and 5–40 mM with the LOD values of 1.13 µM and 2.79 mM, respectively. The method was applied in the evaluation of TAC of drinks (iced tea and green tea), foods (orange, lemon, and tomato), and herbs (Cornus officinalis, Cynanchum otophyllum, Dioscorea bulbifera, and Eriobotryae Folium). The results demonstrate that orange and C. officinalis have a higher TAC. Based on the OXD-like activity of cobalt oxyhydroxide (CoOOH) nanorings, Zhang et al. [37] developed a colorimetric sensor with smartphone assistance for the detection of antioxidants in green tea (Fig. 4). The detection mechanism is the decomposition of CoOOH nanorings into Co2+ after the addition of antioxidants, resulting in a decrease of catalytic activity. The established method exhibited high sensitivity with a LOD of 0.025 µM. Moreover, a smartphone can be used as a readout, and the content of total antioxidants in green tea was measured to be 2.55 µM, which is close to the result of Folin’s method. Meanwhile, Murilo et al. [38] synthesized the manganese dioxide/graphene quantum dot (MnO2/GQD) composites with excellent OXD-like activity, which was applied to detect the total antioxidants in fresh lemon juice, black tea, and mango juice, with recovery values of 95–105%. It is noteworthy that the system can differentiate different antioxidants by treating the obtained data through principal components analysis (PCA).

Fig. 4
figure 4

Schematic diagram of detecting antioxidants based on CoOOH nanoring. Reprinted with permission from [37]

Detection after sample pretreatment

Some active ingredients, such as quercetin and rutin, are insoluble in water, so long-time alcohol extractions are needed. During the preparation of real samples, Davoodi-Rad et al. [39] dried the vegetable samples at 60 °C for 4 h, then ground them into powder. A portion of the powder was mixed with methanol and stirred for 24 h, finally followed by filtration, washing, and dilution. Additionally, several plants, like Datura stramonium, contain diverse components. Therefore, the detection of specific components in them requires complex extraction processes. The preparation of Datura samples required drying and degreasing, which was first extracted with methanol and filtered, and then rotary evaporated to remove the solvent. Following ultrasonication, the samples were further extracted twice with dichloromethane (DCM). This process involved several separation steps, pH adjustments, and drying steps [40].

Quercetin, which is a type of naturally polyphenolic flavonoid compound, is one of the active ingredients in many frequently used CHMs and natural products, such as Ginkgo biloba, licorice, and onions. Quercetin has various properties, including antioxidant, anti-cancer, hypoglycemic, and liver-protective [41]. Cao et al. [42] synthesized the CeO2/Co3O4@N-doped hollow carbon microspheres (CeO2/Co3O4@NCH) through a self-template method, which exhibited excellent POD-like activity due to its larger surface area, pore-like structure, and OVs. Based on the reduction property of quercetin, a facile, fast, and cheap sensor was established to detect it with a linear range of 7–22 µM and a LOD of 1.19 µM. In addition, the sensor was applied to analyze quercetin in Yinxingye Dispersible Tablets, showing satisfactory recoveries. Moreover, some nanozymes based on LAC-like activity have also been developed for the quantitative analysis of quercetin. LAC is a copper-containing polyphenol oxidase that catalyzes polyphenols and polyamines to produce colored ortho-quinone [39, 43]. Davoodi-Rad et al. [39] synthesized the Cu-TA nanosheets (Cu-TA NSs) with LAC-like activity to detect quercetin in treated vegetable samples. Firstly, Cu-TA NSs can oxidize quercetin to generate ortho-quinone. Additionally, the addition of the surfactant cetyltrimethylammonium bromide (CTAB) reacted with quercetin by supramolecular interaction, further promoting the oxidation of quercetin. The developed method showed good selectivity to detect quercetin with a lower LOD of 0.064 µM. Then, it was used to detect the quercetin content in red onion, green pepper, and drill samples, and the results are in consistent with that of HPLC analysis.

Nanozyme-based detections for total polyphenols have also been developed. For instance, Rashtbari et al. [44] synthesized molybdenum trioxide nanoparticles through Argon cold plasma surface modification (Ar-MoO3NPs), which exhibited enhanced POD-like activity. The prepared Ar-MoO3NPs can oxidize non-fluorescent terephthalic acid into high-fluorescence emission compounds in the presence of H2O2. In contrast, polyphenols can cause aggregation of Ar-MoO3NPs and act as free radical scavengers, leading to the quenching of fluorescence. Therefore, a fluorescence method was developed to detect polyphenols with high specificity, which was successfully used to detect total polyphenolics in apple, orange, and grape samples.

Rutin, which has antioxidant, anticancer, vasoprotective, and neuroprotective properties [45], is a polyphenolic flavonoid compound and can be hydrolyzed to produce quercetin. Davoodi-Rad et al. [46] developed a colorimetric method for detecting rutin based on the LAC-like activity of Cu-guanosine nanorods (Cu-Guo NRs), which can oxidize rutin, resulting in a color change from light green to dark yellow. The established strategy has a broad linear range of 0.77–54.46 µM and a LOD of 0.114 µM. Then, it was successfully used to detect rutin in propolis dry extract and rutin-containing dietary supplement tablets, with contents of 9.42% and 18.38 mg per tablet, respectively. Covalent triazine framework (CTF) is a special class of COFs with a triazine ring in its structure. Based on the advantage of chemiluminescent (CL) detection with high sensitivity, Tan et al. [47] prepared a CTF-1 with OXD-like activity, which can oxidize luminol to produce intense CL in the presence of O2 (Fig. 5). Whereas the intensity of CL decreased with the increase of rutin concentration, therefore, a very sensitive CL method can be established for detecting rutin with a LOD of 0.015 µM. Compared with the results of HPLC, the developed CL method is reliable for detecting rutin both in tablets and treated Flos Sophorae Immaturus samples.

Fig. 5
figure 5

Schematic diagram of CL of luminol based on CTF-1. Reprinted with permission from [47]

Atropine is an alkaloid that can be used to dilate the pupil, alleviate spasms, and serve as an antidote to organophosphorus pesticides. Datura is a poisonous plant but contains abundant active chemicals, including phenolics, steroids, acyl sugars, amides, and alkaloids. Therefore, tedious sample pretreatment is necessary to detect atropine content in Datura plants. Mahmoudi et al. [40] synthesized a series of Fe3O4 and bimetal-organic framework Zn/Mg (Fe3O4@MOFs) composites for the detection of atropine extracted from two Datura samples through liquid–liquid extraction. The experimental results show that the Fe3O4@MOF/Dextrin composite exhibited the highest POD-like activity, which was primarily attributed to the cooperative interaction of dispersed Fe ions between Zn and Mg metals in the MOF and dextrin layers. In the presence of the material and H2O2, terephthalic acid was oxidized to 2-hydroxy terephthalic acid, emitting fluorescence at 425 nm. This oxidation process can be inhibited by atropine, allowing the fluorescence detection of atropine with the LOD value of 2.27 μg/L. To further improve the sensitivity of atropine, the group [48] developed a CL method to detect atropine based on the Fe3O4@MOF composite, which can oxidize luminol, producing high-intensity CL (Fig. 6). While atropine can bind with the Fe3O4@MOF composite, leading to a significant reduction of CL intensity. Compared with the previous method, the sensitivity was considerably improved and the LOD was as low as 0.02 μg/L.

Fig. 6
figure 6

Schematic diagram of detecting atropine based on Fe3O4@MOFs. Reprinted with permission from [48]

Specific recognition strategy

Molecularly imprinted polymers (MIPs) are formed by polymerizing monomers in the presence of a template molecule. After removing the template molecule, MIPs can be used to bind template molecule specifically, similar to the interaction between an antibody and an antigen [49, 50]. Therefore, MIP can extract template molecules from complicated samples and shield interference from other substances, improving the assay selectivity [51]. As one of the main active ingredients in Huangqi (Astragalus membranaceus), Astragaloside-IV (AS-IV) has many pharmacological activities, including enhancing immunity, antivirus, anti-stress, antifibrosis, and protecting the heart [52, 53]. The AS-IV is usually detected by HPLC in combination with other methods due to its weak ultraviolet absorption, such as pulsed amperometric detection and evaporative light scattering detection (ELSD) [53, 54]. Chen et al. [51] innovatively combined MIP with CuO nanoparticles (CuO NPs) and polydopamine (PDA) to synthesize MIP@PDA/CuO NPs with POD-like activity for the detection of AS-IV. AS-IV can specifically bind to the surface imprinted cavity to prevent the entry of H2O2 and TMB, inhibiting the catalytic process (Fig. 7). Eventually, the established new colorimetric method for the detection of AS-IV showed high selectivity and the linear range was 0.000341–1.024 mg/mL with a LOD of 0.000991 mg/mL. Additionally, it was applied to detect the content of AS-IV in Huangqi Granules and Ganweikang Tablets, and the results are similar to that measured by HPLC-ELSD.

Fig. 7
figure 7

Schematic diagram of preparing MIP@PDA/CuO NPs (A) and detecting AS-IV (B). Reprinted with permission from [51]

Licorice (Glycyrrhiza uralensis) is a CHM with diverse functions, such as anti-inflammatory and detoxification. Among licorice active ingredients, liquiritin and glycyrrhizic acid are the indicators for authenticating licorice, while licochalcone A and isolicoflavonol are the indicators for evaluating its quality. Thus, the simultaneous detection of these four active substances is of great significance. The sensor array consists of a series of cross-response sensing units rather than a specific receptor, which can detect and discriminate structurally similar components or complex mixtures through pattern recognition [55]. Based on three iron oxide nanozymes (Fe2O3, Fe3O4, and histidine (His)-Fe3O4) with POD-like activity, Yuan et al. [56] constructed a colorimetric sensor array for the detection of four licorice active substances (Fig. 8). Different active ingredients inhibited the catalytic activity of different iron oxides to various degrees, and the developed colorimetric sensor successfully identified and distinguished the four licorice active substances in real licorice samples in the concentration range of 1–200 µM.

Fig. 8
figure 8

Schematic diagram of detecting four licorice active substances based on the colorimetric sensor array. Reprinted with permission from [56]

Although nanozyme-based detection methods have the advantages of simplicity, rapidity, and high sensitivity, there were limited number of methods have been developed for the detection of phytochemicals in natural products using nanozymes. Furthermore, due to the complexity of real samples, there are still difficulties in achieving specific detection, which requires sample pretreatment or combining with special methods such as MIP. Therefore, the design and preparation of nanozymes with high selectivity to solve the problem of complex sample matrix in phytochemical analysis remain in the exploratory stage.

Detection of hazardous substances

Detection of heavy metal ions

Due to their bioaccumulation properties, heavy metal ions can reach very high levels through the diet, thereby harming human health [8]. Common heavy metal elements include arsenic (As), lead (Pb), mercury (Hg), copper (Cu), chromium (Cr), cadmium (Cd), iron (Fe), etc. [57]. As shown in Table 3, there are several studies on the detection of As and Pb in plants using nanozymes.

Table 3 Summary of detection of heavy metal ions in plants based on nanozymes

Arsenic ion

Compared with organic arsenic, inorganic arsenic exhibits higher toxicity. Excessive intake of it can cause skin and respiratory diseases, nerve poisoning, organ failure, and even cancer, which may be resulted from its interaction with enzymes in the human body and excess generation of reactive oxygen species (ROS) [58,59,60,61]. Inorganic arsenic includes arsenic trivalent (As(III)), arsenic pentavalent (As(V)), and elemental arsenic, while As(III) is more toxic than As(V) as it can bind to sulfhydryl groups with higher affinity, inhibiting the activity of various proteins [59, 61]. Wang et al. [59] developed a colorimetric method to detect As(III) (Fig. 9). They synthesized different shapes of Mn3O4 NPs with OXD-like activity, while the octahedral one possessed the strongest As(III) adsorption capacity. Furthermore, arsenic adsorption made Mn4+/Mn3+ reduced to Mn2+, which can catalyze O2 to produce oxygen radicals, further oxidizing TMB with the solution turning to yellow color. Based on As(III)-adsorption enhancing the catalytic activity of octahedral Mn3O4 NPs, a colorimetric method was established and achieved the detection of As(III) in the wheat sample with a LOD of 1.32 μg/L.

Fig. 9
figure 9

Schematic diagram of detecting As (III) based on octahedral Mn3O4 NPs. Reprinted with permission from [59]

As(V) can inhibit the catalytic activity of acid phosphatase (ACP), which can catalyze the hydrolysis of ascorbic acid 2-phosphate (AAP) to produce AA. Therefore, several studies have utilized nanozymes and ACP to design enzyme-cascade reactions for As(V) detection. Xu et al. [58] prepared an ACP and hemin-loaded multifunctional Zn-based metal–organic framework (ACP/hemin@Zn-MOF) for the detection of As(V). Hemin exhibited POD-like activity, which can catalyze the oxidation of o-phenylenediamine (OPD) to form a fluorescent product (564 nm) and weaken its intrinsic fluorescence (452 nm) owing to the inner filter effect. After the addition of AAP, the generated AA will competitively suppress the oxidation of OPD, causing a decrease in the fluorescence intensity at 564 nm and a recovered fluorescence at 452 nm. The inhibitory effect of As(V) on ACP enabled the fluorescence signal to be reversed again, realizing a ratio fluorescence detection of As(V) with a linear range of 3.33–300.00 μg/L and a LOD of 1.05 μg/L. Moreover, the method was successfully applied in the analysis of As(V) and total arsenic in rice samples, with recovery rates ranging from 95 to 105%. Similarly, Wang et al. [60] developed a colorimetric method to detect As(V) utilizing the OXD-like activity of Ce(IV) coordination polymer nanoparticles. With the addition of ACP and AAP, the produced AA can not only restrain the oxidation of TMB but also reduce Ce4+ to Ce2+, inhibiting the enzyme-like activity of the material. Therefore, As(V) can be detected by inhibiting ACP and restoring the TMB color reaction. The method displays a high sensitivity and was used to analyze As(V) in rice samples.

Lead ion

Pb is the second most toxic heavy metal after As with bioaccumulation and persistence [62,63,64,65,66]. Low doses of Pb2+ have an impact on the physical and mental health of infants and young children, causing developmental disorders, brain damage, psychiatric disorders, etc. [67]. Tang et al. [66] synthesized the layered WS2 nanosheets with POD-like activity through a simple ultrasonic stripping method, which was employed to detect Pb2+ in wheat samples. Pb2+ blocked the electron transfer between WS2 and H2O2, and then prevented the oxidation of TMB, resulting in a significant decrease of absorbance at 650 nm. For Pb2+ detection, the linear range of the method was 0.015–0.24 nM with a low LOD of 0.012 nM. Using Pb2+-dependent receptors (e.g. DNAzyme) is a good option to achieve a high selective detection. Si et al. [65] developed an electrochemical method to monitor Pb2+ in vegetable samples based on a porphyrin-functionalized metal–organic framework (porph@MOF) and Pb2+-dependent DNAzyme. As shown in Fig. 10, DNA2 was immobilized on an AuNPs-modified glassy carbon electrode via the Au–S bond. It can be specifically cleaved by Pb2+ to generate a short DNA2 fragment, which was further hybridized with porph@MOF-DNA1 through base pairing. Subsequently, the porph@MOF with POD-like activity oxidated OPD in the presence of H2O2, producing the electrochemical signal. The established method exhibited excellent selectivity and high sensitivity with a LOD of 5 pM. However, although cleavage and hybridization of DNA2 can be finished in one step, the long incubation time (80 min) did not fulfill the requirement of rapid detection. Colorimetric and surface-enhanced Raman spectroscopy (SERS) are commonly regarded as rapid analytical methods [68]. Gold nanoparticles (AuNPs) are widely studied for their enzyme-like and SERS properties [69, 70]. Furthermore, several studies have substantiated that carbon dots (CDs) facilitate improving SERS signals and catalytic activity of AuNPs [71,72,73]. For example, Cui et al. [64] prepared the Fe-doped norepinephrine-based CDs through a one-step microwave digestion method, which were self-assembled with Prussian blue (PB) to obtain Fe, NA-CDs/PB with POD-like activity. They also utilized the reducing property of CDs to synthesize the AuNPs. Based on the inhibition of Pb2+ on the POD-like activity of Fe, NA-CDs/PB and AuNPs, the SERS and colorimetric dual-mode sensor was constructed with the LODs of 0.024 nM and 0.015 nM, respectively. Finally, the sensor was successfully applied to detect Pb2+ in Salvia miltiorrhiza, Astragalus membranaceus, Barley Yellow, and pomegranate peel with good recovery of 90.4–108.9% and RSD of 2.6–4.7%.

Fig. 10
figure 10

Schematic diagram of detecting Pb.2+ based on DNAzyme and porph@MOF. Reprinted with permission from [65]

Other ions

Ingestion of inorganic mercury may cause neurological symptoms (including mental retardation, vision and hearing loss, language disorders, and memory loss), as well as cognitive and motion disorders, etc. [74, 75]. Recently, single-atom nanozymes (SAzymes) with ultra-high atomic utilization, excellent stability, and remarkable catalytic activity have been continuously studied [75,76,77]. Ge et al. [75] developed a novel colorimetric strategy for Hg2+ assay using cysteine and single-atom Cu-C-N nanozymes (SACu-C-N). The prepared SACu-C-N exhibited OXD-like activity, catalyzing the oxidation of TMB to blue ox-TMB. However, the ox-TMB was reduced after the introduction of cysteine. Since Hg2+ possesses a strong affinity for thiol groups of cysteine, it can turn the solution to blue color again. Therefore, a simple, sensitive, and selective colorimetric method was established and applied in the analysis of Hg2+ in cabbage samples.

Cu is an essential trace element and is influential in the metabolic process as a cofactor or structural component of many natural enzymes [78,79,80]. However, excess Cu2+ suppresses the activity of some essential enzymes, causing serious side effects, such as neurodegenerative disease, liver damage, and even cancer [79, 81]. In addition, high levels of Cu also damage photosynthesis and then inhibit plant growth [78]. Xiong et al. [80] synthesized a CTF through a simple and rapid microwave-enhanced high-temperature ionothermal method. Interestingly, CTF possessed weak POD-like activity, but Cu2+ can act as the active catalytic center and a bridge for electronic transfers between the substrate and CTF, resulting in enhanced catalytic activity. The LOD value of the developed method was 1.25 nM, and was applied in the quantification of Cu2+ in Chinese water chestnuts and eggplants, with recoveries of 96.0–105.0%.

Most of the sensors can detect only a single heavy metal, and the simultaneous detection of multiple heavy metal ions is still a challenge. Song et al. [76] designed a time-resolved sensor to detect Cr6+ and Fe3+ based on their difference in enhancing the single-atom Ce–N–C nanozyme’s OXD-like activity. In the presence of Fe3+ and Cr6+ alone, the solution turned blue after 30 s and 60 s, respectively, while the former faded after 5 min. The solution turned blue in 30 s and did not fade when both of them were presented. Therefore, the constructed sensor was feasible for the simultaneous detection of Fe3+ and Cr6+ in actual samples. By utilizing gold nanoclusters (AuNCs) as sensing elements, Li et al. [82] developed a colorimetric sensor array for identifying five heavy metal ions (Hg2+, Pb2+, Cu2+, Cd2+, and Co2+) at a concentration down to 0.5 μM, which was successfully used to recognize multiple heavy metal ions in Lonicera japonica and Chrysanthemum morifolium samples.

Detection of mycotoxins

Mycotoxins are secondary metabolites generated by filamentous fungi and are extensively found in maize, wheat, rice, peanuts, and other cereals, which include aflatoxins, ochratoxins, fumonisins, zearalenone [14, 83], etc. Even at low concentrations, they are nephrotoxic, immunotoxic, teratogenic, mutagenic, and carcinogenic [84, 85]. According to the Chinese Pharmacopoeia, the total aflatoxin content in Chinese medicines should not exceed 10 µg/kg, and zearalenone should not exceed 500 µg/kg [86]. At present, a number of nanozyme-based immunoassays have been reported for detecting mycotoxins (Table 4).

Table 4 Summary of detection of mycotoxins in plants based on nanozymes

Aflatoxin b1

Aflatoxins include aflatoxin B1, B2, G1, and G2 [87]. Among them, aflatoxin B1 (AFB1) is the most toxic one with potent hepatocarcinogens, which was classified as a Group 1 carcinogen as early as 2002 for it can induce formatting DNA adducts, leading to hepatoma [88,89,90]. Apart from this, AFB1 is also associated with malnutrition, growth impairment, and immune inhibition [91,92,93]. Lateral flow immunoassay (LFIA) is a real-time analysis method on paper-based equipment. The principle is mainly based on the competitive binding of the target analyte and the fixed antigen on the detection line to the antibody [94]. Owing to its simplicity, rapidity, and low cost, LFIA has become an attractive immunoassay for AFB1 analysis. However, the limited sensitivity hinders LFIA's practical applications [95,96,97]. As mentioned, nanozymes can catalyze the formation of chromogenic substrates for signal amplification to improve the sensitivity of a detection method. Cai et al. [95] prepared MnO2 nanosheets with excellent OXD-like activity as signal labels conjugated with antibodies to detect AFB1. Using MnO2 catalyzing TMB to produce clear color signals, the method achieved sensitive detection of AFB1 with a LOD of 15 pg/mL and a wide linear range of 0.01–150 ng/mL. Compared to antibodies, aptamers are more stable and flexible in labeling. Therefore, aptamer-mediated LFIA is a promising approach to realize a highly sensitive detection. Zhu et al. [97] designed a PDA-modified nanozyme (CuCo@PDA) with abundant amide groups that can be coupled to AFB1 aptamers via a condensation reaction. Based on the POD-like activity of CuCo@PDA, a reliable and ultrasensitive method combined with a smartphone was established for AFB1 detection with a LOD of 2.2 pg/mL. Moreover, it was successfully applied to detect AFB1 in the peanut, corn, and wheat samples with different contamination levels.

The enzyme-linked immunosorbent assay (ELISA) is a widely used immunoassay. In a typical ELISA assay, antigens (analytes) first bind to antibodies immobilized on a well plate, then forming an antibody-antigen–antibody sandwich with enzyme-labeled antibodies (commonly HRP). After washing steps, HRP catalyzes the added substrate, resulting in a color change [98]. Likewise, the combination of nanozymes can effectively improve the relatively low sensitivity of ELISA [77, 99]. Guo et al. [77] developed a nanozyme-linked immunosorbent assay (NLISA) by utilizing Fe–N–C SAzymes to replace HRP for quantitative detection of AFB1 in peanut samples. Unlike the traditional antibody-antigen–antibody “sandwich” type of detection mechanism, the method immobilized antigens on the well plate, achieving the rapid and sensitive detection of AFB1 with a LOD of 3.3 pg/mL. The coupling of nanozymes with bio-enzymes induces a dual signal amplification through an enzyme cascade to further increase the sensitivity. Lai et al. [99] found that copper hexacyanoferrate nanoparticles (CHNPs) with OXD-like activity can be rapidly produced by simply mixing potassium hexacyanoferrate(III) (K3[Fe(CN)6]) with Cu(II). However, AA produced by the hydrolysis of ALP on ascorbic acid 2-phosphate (AAP) can reduce Fe (III) to Fe (II) and then inhibit the formation of CHNPs. Therefore, employing AuNPs coupled to ALP as enzyme labels in ELISA and integrating with the production process of CHNPs, a highly sensitive colorimetric immunoassay for the determination of AFB1 was constructed with a LOD of 0.73 pg/mL. Finally, the developed method was used to detect AFB1 in spiked and naturally contaminated peanut samples.

The single signal mode is prone to false-negative/positive results caused by differences in operating conditions and environment. In contrast, multi-mode detection can offset interferences, reduce false results through self-correction, and yield more precise outcomes [90, 100]. Huang et al. [101] developed a colorimetric/photothermal dual-mode immunoassay method based on Pt supported on nitrogen-doped carbon (Pt-CN) for monitoring AFB1 in peanut samples. After competitive immunoreactivity of glucose oxidase (GOx)-labeled antigen with AFB1, the GOx loaded on the well plates can catalyze the formation of H2O2 from glucose. Subsequently, the Pt-CN with POD-like activity can oxidize TMB to blue ox-TMB, producing a colorimetric signal. On the other hand, the ox-TMB, as a photothermal agent, can convert light to heat under near-infrared (NIR) irradiation, generating a photothermal signal. The LOD values are 0.22 pg/mL and 0.76 pg/mL for the colorimetric and photothermal assays, respectively. The fluorescence method is regarded as one of the most sensitive of the optical methods. Lu et al. [90] designed a photothermal/colorimetric/fluorescent multimodal NLISA to portably and ultra-sensitively detect AFB1. As shown in Fig. 11, magnetic nanoparticles (MNPs) combined with aptamers were immobilized on the well plates via antibody and AFB1. In the presence of K4[Fe(CN)6] and HCl, MNPs as precursors can form Prussian blue nanoparticles (PBNPs) that possessed both excellent POD-like activity and photothermal effect. Particularly, MNPs also acted as the quencher to decrease the fluorescence of the dye (Cy5), which was restored upon the formation of PBNPs, enabling fluorescence detection of FAB1, with an extremely low LOD of 0.54 fg/mL. The established strategy was feasible for the qualitative and quantitative determination of AFB1 in the actual samples on the spot.

Fig. 11
figure 11

Schematic diagram of multimode detecting AFB1. Reprinted with permission from [90]

Ochratoxin A

Ochratoxins are a series of mycotoxins generated by Penicillium and Aspergillus. Among them, ochratoxin A (OTA) is the most poisonous to human health. It possesses various toxicity in animals and humans, including nephrotoxicity, hepatotoxicity, immunotoxicity, teratogenicity, and carcinogenicity [100, 102, 103]. Recently, studies have shown that OTA is also a potent neurotoxin that is considered to be a causative agent of neurodegenerative diseases, and the brain is one of the main target organs of its damage [104]. As mentioned, dual-mode analysis has the advantage of better sensitivity and more accurate results, which was employed in most of the nanozymes-based detection methods of OTA [100, 105,106,107,108,109]. Li et al. [109] developed a SERS/colorimetric dual-mode method for the detection of OTA using Pd–Pt bimetallic nanocrystals (Pd–Pt NRs) conjugated with aptamers as recognition probes. Since the Pd–Pt NRs with POD-like activity can catalyze the oxidation of TMB to ox-TMB that exhibited a strong SERS signal, the SERS and colorimetric detection of OTA can be achieved with LODs of 0.042 nM and 0.097 nM, respectively. The developed method was applied to detect OTA in grape samples, and the results are in consistent with that of UPLC-MS/MS analysis. Zheng et al. [100] established a colorimetric/fluorescence immunoassay method for detecting OTA based on the OXD-like activity of cerium-based nanoparticles (CPNs(IV)) and the fluorescence properties of CPNs(III). The LODs of colorimetric and fluorescence methods are 0.962 ng/mL and 0.404 ng/mL, with recoveries in corn samples ranging from 99.12–102.60% to 97.60–103.55%, respectively. In addition, to improve the accuracy of visual judgments, colorimetric immunoassays with multiple color changes have also been reported. Gold nanomaterials have garnered attention for the color of their solutions, which largely depends on their shape and size [105, 107, 110]. Zhu et al. [110] synthesized octahedral Cu2O nanoparticles with POD-like activity, which can oxidize TMB to TMB2+ in the presence of H2O2 and HCl. Based on the significant color change resulted from the etching of TMB2+ on gold nano bipyramids (Au NBPs), a multi-colorimetric immunoassay was developed to monitoring of OTA in millet samples with a LOD of 0.47 ng/L (Fig. 12).

Fig. 12
figure 12

Schematic diagram of multi-colorimetric detecting OTA. Reprinted with permission from [110]

Zearalenone

Zearalenone (ZEN) is a kind of mycotoxin with estrogenic activity, which can compete with the natural estrogen, resulting in the reproductive dysfunction of animals [83]. In addition, ZEN may cause other toxic effects involving hepatotoxicity, immunotoxicity, genotoxicity, and carcinogenicity [83, 111,112,113,114111‒114]. Sun et al. [114] developed a colorimetric method for the detection of ZEN based on the inhibition of ZEN aptamer on the POD-like activity of AuNPs. After the addition of ZEN, the aptamer bound to it preferentially with the restoration of AuNPs activity. The LOD value of the method is 10 ng/mL and the recovery in spiked corn is in the range of 92%‒102%. Bimetallic nanoparticles are superior to monometallic nanoparticles in terms of catalytic activity [108, 109]. Liu et al. [113] synthesized encapsulated AuPt nanoparticles hydrogel by ZEN aptamer and complementary DNA as crosslinkers. In the presence of ZEN, it will preferentially combine with the aptamer, destroying the hydrogel structure and then releasing the AuPt nanozymes to complete the catalytic reaction. Therefore, a highly sensitive colorimetric method for the determination of ZEN was established with a LOD of 0.6979 ng/mL, which was applied to detect ZEN in corn and wheat samples. However, metal nanoparticles alone are prone to aggregation, resulting in reduced catalytic sites and lower catalytic activity. Anchoring it to a carrier material is an effective solution to this problem [91]. For example, utilizing Ti3C2Tx nanosheet as a carrier material, Huang et al. [112] prepared a Ti3C2Tx/AuNPs nanocomposite with enhanced POD-like activity, which can be used as an immunoprobe to detect ZEN. Employing Ti3C2Tx/AuNP to catalyze the oxidation of TMB and the strong NIR-driven photothermal effect of ox-TMB, the immunoassay achieved ultrasensitive colorimetric and photothermal dual-mode detection of ZEN with LODs of 0.15 pg/mL and 0.48 pg/mL, respectively. Furthermore, the dual-mode strategy was employed for the analysis of ZEN in three contaminated cereal samples, and the results are in good agreement with UPLC-MS/MS analysis.

Detection of organophosphorus pesticide

Organophosphorus pesticide exposure primarily causes chronic or acute toxicity in humans, plants, and animals through inhibiting cholinesterase activity and leading to acetylcholine accumulation [115]. The hazards of OPs on human beings are generally dominated by acute toxicity, which is manifested by a series of neurotoxic symptoms like sweating, tremors, confusion, speech disorders, and in severe cases, respiratory paralysis and even death [116]. As summarized in Table 5, the strategies for detecting OPs based on nanozymes are categorized as follows: (1) direct influence of OPs on the activity of nanozymes [117,118,119,120,121]; (2) nanozymes with organophosphorus hydrolase (OPH)-like or phosphatase-like activity hydrolyze OPs into products that produce signals [122,123,124,125,126,127,128,129,130,131]; (3) nanozymes combine with enzymes that can be inhibited by OPs such as acetylcholinesterase (AChE), alkaline phosphatase (ALP), and ACP [132,133,134,135,136,137,138,139,140,141]; (4) nanozymes combine with antibodies or aptamers [142, 143]. Since a review of comprehensive and systematic description of detecting OPs has been reported [10], this paper will not delve into too much detail.

Table 5 Summary of detection of organophosphorus pesticide in plants based on nanozymes

Challenges and prospects

Nanozymes can overcome some drawbacks of natural enzymes and exhibit higher catalytic activity. At present, some nanozymes have been designed for the analysis of plant samples, but there are still some limitations. To promote the development of nanozymes and their application in the detection of plant samples, the following challenges and prospects are proposed.

  1. (1)

    The nanozymes currently used for phytochemical detection are mainly nanomaterials with POD-like or OXD-like activity. The similar detection mechanisms (inhibition on the catalytic activity of nanozymes by antioxidant properties) make it challenging to achieve high specificity in detection. Therefore, the design of nanozymes with other catalytic properties or multiple reaction mechanisms to improve the selectivity of nanozymes for the detection of phytochemicals deserves further investigation.

  2. (2)

    The variety of components based on nanozymes detection in real samples is limited. In plant samples, complex compositions may affect the activity of nanozymes, leading to inaccurate results. Accordingly, detecting active ingredients in real samples often requires complex pretreatments. Developing nanozymes with specific adsorption or combination with other techniques (e.g., MIP) shows greater prospects.

  3. (3)

    In the detection of hazardous substances (e.g., heavy metal ions and pesticide residues), most samples are spiked with them rather than detecting their actual contents directly. This may be due to the low level of hazardous substances in the original sample and the lack of sensitivity of the assay. Modification of nanomaterials with small molecules (e.g., fluorescein derivatives and vitamin B6) that possess POD-like activity to enhance their enzyme-like activity or enable multiple enzyme activities are anticipated to improve the sensitivity of the assay.

  4. (4)

    Nanozyme-based methods for the detection of mycotoxins are mostly combined with immune methods or aptamers, making them more complex than direct colorimetric or fluorescent assays. Only one study was found on the direct colorimetric detection of AFB1 in corn and peanut samples, but its sensitivity is relatively low [89]. Hence, it is necessary to develop more sensitive and direct methods for easier analysis.

  5. (5)

    Currently, there is less literature on detecting heavy metal ions in plant samples (especially in CHMs) based on nanozymes. CHMs have been extensively used in disease treatment and healthcare for their unique therapeutic effects [144]. However, they are susceptible to contamination by chemicals in the environment. Therefore, evaluating the safety of CHMs using nanozymes is very important.

  6. (6)

    Although many nanozymes can mimic the catalysis activity of natural enzymes, their specificity is still inadequate and requires further optimization. In addition, some soluble transition metal nanozymes may be highly toxic and contaminate environment by releasing toxic metal ions. Consequently, it is important to develop nanozymes with improved specificity and biocompatibility.

Conclusions

This paper summarizes the applications of nanomaterials with enzyme-like activity in plant samples analysis from 2015 to the present, including the analysis of phytochemicals, organophosphorus pesticides, heavy metal ions, and mycotoxins. Improving the selectivity is a research priority for the detection of phytochemicals, which may be achieved through multimode detection and molecular imprinting. Furthermore, due to the trace levels of contaminates, it is crucial to improve sensitivity for detecting hazardous substances. There are various methods have been reported to achieve this goal, including the cascade reaction of natural enzymes with nanozymes and the enhancement of nanozymes’ catalytic activity through doping with other elements or material modification. In general, designing nanozymes with more enzyme-like activities and improving the specificity and sensitivity in their applications are the focus of future research.

Availability of data and materials

All data generated or analyzed during this review is included in published articles.

Abbreviations

AA:

Ascorbic acid

AAP:

Ascorbic acid 2-phosphate

AAS:

Atomic absorption spectrometry

AChE:

Acetylcholinesterase

ACP:

Acid phosphatase

AFB1:

Aflatoxin B1

ALP:

Alkaline phosphatase

Ar-MoO3NPs:

Argon cold plasma surface modified molybdenum trioxide nanoparticles

As:

Arsenic

As(III):

Arsenic trivalent

As(V):

Arsenic pentavalent

AS-IV:

Astragaloside-IV

Au NBPs:

Gold nano bipyramids

AuNCs:

Gold nanoclusters

AuNPs:

Gold nanoparticles

Cd:

Cadmium

CDs:

Carbon dots

CE:

Capillary electrophoresis

CHMs:

Chinese herbal medicines

CHNPs:

Copper hexacyanoferrate nanoparticles

CL:

Chemiluminescent

COFs:

Covalent organic frameworks

CoOOH:

Cobalt oxyhydroxide

CPNs:

Cerium-based nanoparticles

Cr:

Chromium

CTAB:

Cetyltrimethylammonium bromide

CTF:

Covalent triazine framework

Cu:

Copper

Cu-Guo NRs:

Cu-guanosine nanorods

DCM:

Dichloromethane

ELISA:

Enzyme-linked immunosorbent assay

ELSD:

Evaporative light scattering detection

Fe:

Iron

Fe3O4@MOFs:

Fe3O4 and bimetal-organic framework Zn/Mg

GA:

Gallic acid

GC:

Gas chromatography

GOx:

Glucose oxidase

Hg:

Mercury

HHTP:

2, 3, 6, 7, 10, 11-Hexahydroxytriphenylene

HNCs:

Hollow nanocages

HPLC:

High-performance liquid chromatography

HRP:

Horseradish peroxidase

ICP:

Inductively coupled plasma

K3[Fe(CN)6]:

Potassium hexacyanoferrate(III)

LAC:

Laccase

LFIA:

Lateral flow immunoassay

LOD:

Limit of detection

Mb:

Methanobactin

M-CATs:

Metal-catecholates

MIPs:

Molecularly imprinted polymers

MnO2/GQD:

Manganese dioxide/graphene quantum dot

MNPs:

Magnetic nanoparticles

MOFs:

Metal organic frameworks

MRLs:

Maximum residue limits

MS:

Mass spectrometry

NCH:

N-doped hollow carbon microspheres

NIR:

Near-infrared

NLISA:

Nanozyme-linked immunosorbent assay

N-Mn3O4 NSps:

Nitrogen-doped Mn3O4 nanospheres

NPs:

Nanoparticles

Pd–Pt NRs:

Pd–Pt bimetallic nanocrystals

NSs:

Nanosheets

OPD:

O-Phenylenediamine

OPs:

Organophosphorus pesticides

OTA:

Ochratoxin A

OVs:

Oxygen vacancies

OXD:

Oxidase

ox-TMB:

Oxidized TMB

Pb:

Lead

PB:

Prussian blue

PBNPs:

Prussian blue nanoparticles

PCA:

Principal components analysis

PDA:

Polydopamine

PFOS:

Perfluorooctane sulfonate

POD:

Peroxidase

porph@MOF:

Porphyrin-functionalized metal–organic framework

PPO:

Polyphenol oxidase

Pt-CN:

Pt supported on nitrogen-doped carbon

rGO:

Reduced graphene oxide

ROS:

Reactive oxygen species

SACu-C-N:

Single-atom Cu-C-N

SAzymes:

Single-atom nanozymes

SERS:

Surface-enhanced Raman spectroscopy

SrTiO3 :

Strontium titanate

TA:

Tannic acid

TAC:

Total antioxidant capacity

TMB:

3, 3’, 5, 5’-Tetramethylbenzidine

UPLC-MS/MS:

Ultra-performance liquid chromatography-tandem mass spectrometry

ZEN:

Zearalenone

Zn-MOF:

Zn-based metal–organic framework

References

  1. Yang CJ, Zhao Y, Jiang S, Sun XM, Wang XT, Wang ZB, et al. A breakthrough in phytochemical profiling: ultra-sensitive surface-enhanced Raman spectroscopy platform for detecting bioactive components in medicinal and edible plants. Microchim Acta. 2024;191:286.

    Article  CAS  Google Scholar 

  2. Rodríguez-Negrete EV, Morales-González Á, Madrigal-Santillán EO, Sánchez-Reyes K, Álvarez-González I, Madrigal-Bujaidar E, et al. Phytochemicals and their usefulness in the maintenance of health. Plants. 2024;13:523.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Barbieri R, Coppo E, Marchese A, Daglia M, Sobarzo-Sánchez E, Nabavi SF, Nabavi SM. Phytochemicals for human disease: An update on plant-derived compounds antibacterial activity. Microbiol Res. 2017;196:44–68.

    Article  CAS  PubMed  Google Scholar 

  4. Imam H, Wu H, Luo T, Arshad M, Song JY, Xu DX, et al. Phytochemicals and inflammatory bowel disease: a review. Crit Rev Food Sci Nutr. 2022;60:1321–45.

    Google Scholar 

  5. Chen X, Yang Z, Xu Y, Liu Z, Liu YF, Dai YT, et al. Progress and prediction of multicomponent quantification in complex systems with practical LC-UV methods. J Pharm Anal. 2023;13:142–55.

    Article  PubMed  Google Scholar 

  6. Gotti R. Capillary electrophoresis of phytochemical substances in herbal drugs and medicinal plants. J Pharm Biomed Anal. 2011;55:775–801.

    Article  CAS  PubMed  Google Scholar 

  7. Pan ZW, Gong TY, Liang P. Heavy metal exposure and cardiovascular disease. Circ Res. 2024;134:1160–78.

    Article  CAS  PubMed  Google Scholar 

  8. Si LX, Wu Q, Jin YL, Wang Z. Research progress in the detection of trace heavy metal ions in food samples. Front Chem. 2024;12:1423666.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Chinese Pharmacopoeia Commission. Guiding principles for the formulation of the limit of harmful residues in traditional Chinese medicine 9302. In: Pharmacopoeia of People’s Republic of China Part 4. Beijing: China Medical Science Press; 2020. p. 520–2.

    Google Scholar 

  10. Zhao FN, Wang L, Li MY, Wang M, Liu GY, Ping JF. Nanozyme-based biosensor for organophosphorus pesticide monitoring: functional design, biosensing strategy, and detection application. TrAC Trends Anal Chem. 2023;165: 117152.

    Article  CAS  Google Scholar 

  11. Bedair H, Rady HA, Hussien AM, Pandey M, Apollon W, Alkafaas SS, Ghosh S. Pesticide detection in vegetable crops using enzyme inhibition methods: a comprehensive review. Food Anal Methods. 2022;15:1979–2000.

    Article  Google Scholar 

  12. National Health Commission of the People’s Republic of China. National food safety standard-Maximum residue limits for pesticides in food. GB 2763‒2021; 2021.

  13. Zhang XL, Wu D, Zhou XX, Yu YX, Liu JC, Hu N, et al. Recent progress on the construction of nanozymes-based biosensors and their applications to food safety assay. Trends Anal Chem. 2019;121: 115668.

    Article  CAS  Google Scholar 

  14. Agriopoulou S, Stamatelopoulou E, Varzakas T. Advances in analysis and detection of major mycotoxins in foods. Foods. 2020;9:518.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Wei H, Wang EK. Nanomaterials with enzyme-like characteristics (nanozymes): next-generation artificial enzymes. Chem Soc Rev. 2013;42:6060–93.

    Article  CAS  PubMed  Google Scholar 

  16. Wang ZR, Zhang RF, Yan XY, Fan KL. Structure and activity of nanozymes: inspirations for de novo design of nanozymes. Mater Today. 2020;41:81–119.

    Article  CAS  Google Scholar 

  17. Huang X, Zhang ST, Tang YJ, Zhang XY, Bai Y, Pang H. Advances in metal-organic framework-based nanozymes and their applications. Coord Chem Rev. 2021;449: 214216.

    Article  CAS  Google Scholar 

  18. Yang WP, Yang X, Zhu LJ, Chu HS, Li XY, Xu WT. Nanozymes: Activity origin, catalytic mechanism, and biological application. Coord Chem Rev. 2021;448: 214170.

    Article  CAS  Google Scholar 

  19. Zhang DH, Kukkar D, Kaur H, Kim KH. Recent advances in the synthesis and applications of single-atom nanozymes in food safety monitoring. Adv Colloid Interface Sci. 2023;319: 102968.

    Article  CAS  PubMed  Google Scholar 

  20. Wang KD, Meng XQ, Yan XY, Fan KL. Nanozyme-based point-of-care testing: revolutionizing environmental pollutant detection with high efficiency and low cost. Nano Today. 2024;54: 102145.

    Article  CAS  Google Scholar 

  21. Ling ZZ, Yang JY, Zhang YY, Zeng DP, Wang Y, Tian YX, et al. Applications of advanced materials in the pretreatment and rapid detection of small molecules in foods: a review. Trends Food Sci Technol. 2023;141: 104175.

    Article  CAS  Google Scholar 

  22. Chen HY, Zhang L, Hu Y, Zhou CS, Lan W, Fu HY, She YB. Nanomaterials as optical sensors for application in rapid detection of food contaminants, quality and authenticity. Sens Actuators B Chem. 2021;329: 129135.

    Article  CAS  Google Scholar 

  23. Wang JL, Chai TQ, Chen LX, Chen GY, Chen H, Yang FQ. Manganese coordination polymer nanoparticles with excellent oxidase-like activity for the rapidly and selectively colorimetric detection of glutathione. Microchem J. 2024;199: 110207.

    Article  CAS  Google Scholar 

  24. Xu X, Ma MY, Zhou XY, Zhao X, Feng DM, Zhang L. Portable hydrogel kits made with bimetallic nanozymes for point-of-care testing of perfluorooctanesulfonate. ACS Appl Mater Interfaces. 2024;16:15959–69.

    Article  CAS  PubMed  Google Scholar 

  25. Yang Y. Phytochemicals and health. In: Zhang L, editor. Nutritional toxicology. Springer Nature Singapore: Singapore; 2022. p. 309–54.

    Chapter  Google Scholar 

  26. Kumar A, Nirmal P, Kumar M, Jose A, Tomer V, Emel OZ, et al. Major phytochemicals: recent advances in health benefits and extraction method. Molecules. 2023;28:887.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Sharma BR, Kumar V, Gat Y, Kumar N, Parashar A, Pinakin DJ. Microbial maceration: a sustainable approach for phytochemical extraction. 3 Biotech. 2018;8:401.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Zhou T, Chen DQ, Li HR, Ge DH, Chen XJ. Enhanced oxidase mimic activity of raspberry-like N-doped Mn3O4 with oxygen vacancies for efficient colorimetric detection of gallic acid coupled with smartphone. Food Chem. 2024;447: 138919.

    Article  CAS  PubMed  Google Scholar 

  29. Chen LY, Yang J, Chen W, Sun SG, Tang H, Li YC. Perovskite mesoporous LaFeO3 with peroxidase-like activity for colorimetric detection of gallic acid. Sens Actuators B Chem. 2020;321: 128642.

    Article  CAS  Google Scholar 

  30. Xie XY, Chen XF, Wang YH, Zhang MS, Fan YX, Yang XP. High-loading Cu single-atom nanozymes supported by carbon nitride with peroxidase-like activity for the colorimetric detection of tannic acid. Talanta. 2023;257: 124387.

    Article  CAS  PubMed  Google Scholar 

  31. Wu CH, Qin ZY, Liu YX, Qin XG, Liu G, Wei XL, Zhang HZ. Amorphous iron-catecholates featuring efficient peroxidase-like activity for quick colorimetric detection of tannic acid. LWT. 2024;197: 115896.

    Article  CAS  Google Scholar 

  32. Liu YG, Ye HL, Ying MH, Lin X, Jia X, Pan HB. In-situ growth of SrTiO3 nanosheets on graphene oxide for colorimetric detection of tannins in tea and behavior of active oxygen radicals in the nanozymatic process. Colloids Surf A. 2023;675: 132109.

    Article  CAS  Google Scholar 

  33. Zhang JK, Yang Y, Qin FM, Hu TT, Zhao XS, Zhao SC, et al. Catalyzing generation and stabilization of oxygen vacancies on CeO2−x nanorods by Pt nanoclusters as nanozymes for catalytic therapy. Adv Healthcare Mater. 2023;12:2302056.

    Article  CAS  Google Scholar 

  34. Chen LL, Song JQ, Wang L, Li XT, Hao X, Zhang HP, Fan TJ. Fabrication of a dual mimetic enzyme sensor based on gold nanoparticles modified with Cu(II)-coordinated methanobactin for gallic acid detection. J Food Meas Charact. 2024;18:3142–59.

    Article  Google Scholar 

  35. Wu SY, Zhang P, Jiang ZW, Zhang WD, Gong X, Wang Y. Enhanced peroxidase-like activity of CuS hollow nanocages by plasmon-induced hot carriers and photothermal effect for the dual-mode detection of tannic acid. ACS Appl Mater Interfaces. 2022;14:40191–9.

    Article  CAS  PubMed  Google Scholar 

  36. He J, Yang L, Zhang Y, Li RH, Wu JJ, Cao QQ, et al. Pd-Pt-Ru nanozyme with peroxidase-like activity for the detection of total antioxidant capacity. Anal Methods. 2022;15:8–16.

    Article  PubMed  Google Scholar 

  37. Zhang JJ, Li YF, Gong X, Wang Y, Fu WS. Colorimetric detection of total antioxidants in green tea with oxidase-mimetic CoOOH nanorings. Colloids Surf B. 2022;218: 112711.

    Article  CAS  Google Scholar 

  38. Facure MHM, Andre RS, Mercante LA, Correa DS. Colorimetric detection of antioxidants in food samples using MnO2/graphene quantum dot composites with oxidase-like activity. ACS Appl Nano Mater. 2022;5:15211–9.

    Article  CAS  Google Scholar 

  39. Davoodi-Rad K, Shokrollahi A, Shahdost-Fard F, Azadkish K, Madani-Nejad E. A smartphone-based colorimetric assay using Cu-tannic acid nanosheets (Cu-TA NShs) as a laccase-mimicking nanozyme for visual detection of quercetin in vegetables. Microchim Acta. 2024;191:168.

    Article  CAS  Google Scholar 

  40. Mahmoudi S, Chaichi MJ, Shamsipur M, Nazari OL, Samadi Maybodi AR. Modification of bimetal Zn/Mg MOF with nanoparticles Fe3O4 and Fe3O4@SiO2, investigation of the peroxidase-like activity of these compounds by calorimetry and fluorimetry methods. Heliyon. 2023;9: e12866.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ye KX, Xu SF, Zhou QQ, Wang ST, Xu ZG, Liu ZM. Advances in molecular imprinting technology for the extraction and detection of quercetin in plants. Polymers. 2023;15:2107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Cao XY, Zhao S, Liu XW, Zhu XX, Gao Y, Liu QY. CeO2/Co3O4@N-doped hollow carbon microspheres with improved peroxidase-like activity for the determination of quercetin. Anal Bioanal Chem. 2022;414:4767–75.

    Article  CAS  PubMed  Google Scholar 

  43. Wang JH, Huang RL, Qi W, Su RX, He ZM. Construction of biomimetic nanozyme with high laccase- and catecholase-like activity for oxidation and detection of phenolic compounds. J Hazard Mater. 2022;429: 128404.

    Article  CAS  PubMed  Google Scholar 

  44. Rashtbari S, Dehghan G, Amini M, Khorram S, Khataee A. A sensitive colori/fluorimetric nanoprobe for detection of polyphenols using peroxidase-mimic plasma-modified MoO3 nanoparticles. Chemosphere. 2022;295: 133747.

    Article  CAS  PubMed  Google Scholar 

  45. Feng GJ, Yang Y, Zeng JT, Zhu J, Liu JJ, Wu L, et al. Highly sensitive electrochemical determination of rutin based on the synergistic effect of 3D porous carbon and cobalt tungstate nanosheets. J Pharm Anal. 2022;12:453–9.

    Article  PubMed  Google Scholar 

  46. Davoodi-Rad K, Shokrollahi A, Shahdost-Fard F, Azadkish K. Copper-guanosine nanorods (Cu-Guo NRs) as a laccase mimicking nanozyme for colorimetric detection of rutin. Biosensors. 2023;13:374.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Tan HN, Zhao YX, Xu XT, Sun Y, Li YH, Du JX. A covalent triazine framework as an oxidase mimetic in the luminol chemiluminescence system: application to the determination of the antioxidant rutin. Microchim Acta. 2019;187:42.

    Article  Google Scholar 

  48. Mahmoudi S, Chaichi MJ, Shamsipur M, Nazari OL, Samadi-Maybodi A. Fe3O4 and bimetal-organic framework Zn/Mg composite peroxide-like catalyze luminol chemiluminescence for specific measurement of atropine in Datura plant. Luminescence. 2023;38:1711–9.

    Article  CAS  PubMed  Google Scholar 

  49. Belbruno JJ. Molecularly imprinted polymers. Chem Rev. 2018;119:94–119.

    Article  PubMed  Google Scholar 

  50. Zhang ZJ, Li YQ, Zhang XH, Liu JW. Molecularly imprinted nanozymes with faster catalytic activity and better specificity†. Nanoscale. 2019. https://doi.org/10.1039/C8NR09816F.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Chen GY, Chen LX, Gao J, Chen CY, Guan JL, Cao ZM, et al. A novel molecularly imprinted sensor based on CuO nanoparticles with peroxidase-like activity for the selective determination of astragaloside-IV. Biosensors. 2023;13:959.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Chai XY, Gu YQ, Lv L, Chen C, Feng F, Cao Y, et al. Screening of immune cell activators from Astragali Radix using a comprehensive two-dimensional NK-92MI cell membrane chromatography/C18 column/time-of-flight mass spectrometry system. J Pharm Anal. 2022;12:725–32.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Kwon HJ, Park YD. Determination of astragalin and astragaloside content in Radix Astragali using high-performance liquid chromatography coupled with pulsed amperometric detection. J Chromatogr A. 2012;1232:212–7.

    Article  CAS  PubMed  Google Scholar 

  54. Li JD, Li P, Wu D, Wang ZX, Liu Y, Wang XC, Zhou JW. Effects of different processing methods on the content of astragaloside in danggui buxue decoction detection by high performance liquid chromatography-evaporative light scattering detection. Mater Express. 2022;12:1004–11.

    Article  Google Scholar 

  55. Yang MH, Zhang M, Jia MY. Optical sensor arrays for the detection and discrimination of natural products. Nat Prod Rep. 2023;40:628–45.

    Article  CAS  PubMed  Google Scholar 

  56. Yuan XH, Cheng SC, Chen LY, Cheng ZY, Liu J, Zhang H, et al. Iron oxides based nanozyme sensor arrays for the detection of active substances in licorice. Talanta. 2023;258: 124407.

    Article  CAS  PubMed  Google Scholar 

  57. Zhang L, Bi XY, Liu XH, He Y, Li LB, You TY. Advances in the application of metal-organic framework nanozymes in colorimetric sensing of heavy metal ions. Nanoscale. 2023;15:12853–67.

    Article  CAS  PubMed  Google Scholar 

  58. Xu XC, Luo ZJ, Ye K, Zou XB, Niu XH, Pan JM. One-pot construction of acid phosphatase and hemin loaded multifunctional metal-organic framework nanosheets for ratiometric fluorescent arsenate sensing. J Hazard Mater. 2020;412: 124407.

    Article  PubMed  Google Scholar 

  59. Wang JJ, Tao H, Lu TT, Wu YG. Adsorption enhanced the oxidase-mimicking catalytic activity of octahedral-shape Mn3O4 nanoparticles as a novel colorimetric chemosensor for ultrasensitive and selective detection of arsenic. J Colloid Interface Sci. 2020;584:114–24.

    Article  PubMed  Google Scholar 

  60. Wang LJ, Yang JL, Yan Y, Zhang YS, Xu XC. A smartphone-integrated colorimetric quantitative analysis platform based on oxidase-like Ce(IV)-ATP-Tris CPNs/CNF test strip for detection of inorganic arsenic in rice. Anal Chim Acta. 2022;1227: 340308.

    Article  CAS  PubMed  Google Scholar 

  61. Saifullah, Dahlawi S, Naeem A, Iqbal M, Farooq MA, Bibi S, Rengel Z. Opportunities and challenges in the use of mineral nutrition for minimizing arsenic toxicity and accumulation in rice: A critical review. Chemosphere. 2018;194:171‒88.

  62. Zulfiqar U, Farooq M, Hussain S, Maqsood M, Hussain M, Ishfaq M, et al. Lead toxicity in plants: impacts and remediation. J Environ Manage. 2019;250: 109557.

    Article  CAS  PubMed  Google Scholar 

  63. Li XX, Lan X, Liu W, Cui XW, Cui ZJ. Toxicity, migration and transformation characteristics of lead in soil-plant system: effect of lead species. J Hazard Mater. 2020;395: 122676.

    Article  CAS  PubMed  Google Scholar 

  64. Cui YF, Li QL, Yang DZ, Yang YL. Colorimetric-SERS dual-mode sensing of Pb(II) ions in traditional Chinese medicine samples based on carbon dots-capped gold nanoparticles as nanozyme. Spectrochim Acta A Mol Biomol Spectrosc. 2024;313: 124100.

    Article  CAS  PubMed  Google Scholar 

  65. Zhang XN, Huang XY, Xu YW, Wang X, Guo ZM, Huang XW, et al. Single-step electrochemical sensing of ppt-level lead in leaf vegetables based on peroxidase-mimicking metal-organic framework. Biosens Bioelectron. 2020;168: 112544.

    Article  CAS  PubMed  Google Scholar 

  66. Tang Y, Hu Y, Yang YX, Liu BY, Wu YG. A facile colorimetric sensor for ultrasensitive and selective detection of lead(II) in environmental and biological samples based on intrinsic peroxidase-mimic activity of WS2 nanosheets. Anal Chim Acta. 2020;1106:115–25.

    Article  CAS  PubMed  Google Scholar 

  67. Yu Y, Zhang Y, Li WH, Wang ZW, Zhang J. DNA nanocage confined DNAzyme for detection of lead ions coupled with CRISPR-Cas12a system. Chem Eng J. 2024;480: 148177.

    Article  CAS  Google Scholar 

  68. Dasary SSR, Jones YK, Barnes SL, Ray PC, Singh AK. Alizarin dye based ultrasensitive plasmonic SERS probe for trace level cadmium detection in drinking water. Sens Actuators B Chem. 2016;224:65–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Li QL, Han QQ, Yang DZ, Li KX, Wang YJ, Chen D, et al. Methylmercury-sensitized “turn on” SERS-active peroxidase-like activity of carbon dots/Au NPs nanozyme for selective detection of ochratoxin A in coffee. Food Chem. 2024;434: 137440.

    Article  CAS  PubMed  Google Scholar 

  70. Yaseen T, Pu HB, Sun DW. Fabrication of silver-coated gold nanoparticles to simultaneously detect multi-class insecticide residues in peach with SERS technique. Talanta. 2019;196:537–45.

    Article  CAS  PubMed  Google Scholar 

  71. León Anchustegui VA, Zhu JH, He LY, Bi Y, Dong YY, Liu JH, Wang SH. Coencapsulation of carbon dots and gold nanoparticles over escherichia coli for bacterium assay by surface-enhanced Raman scattering. ACS Appl Bio Mater. 2021;4:597–604.

    Article  Google Scholar 

  72. Li H, Jiang CN, He X, Li CN, Jiang ZL. Aptamer SERS and RRS determination of trace lead ions using nitrogen-doped carbon dot to catalyze the new nano-gold reaction. Spectrochim Acta A Mol Biomol Spectrosc. 2023;303: 123146.

    Article  CAS  PubMed  Google Scholar 

  73. Wang HL, Zhang ZH, Chen CQ, Liang AH, Jiang ZL. Fullerene carbon dot catalytic amplification-aptamer assay platform for ultratrace As3+ utilizing SERS/RRS/Abs trifunctional Au nanoprobes. J Hazard Mater. 2021;403: 123633.

    Article  CAS  PubMed  Google Scholar 

  74. Budnik LT, Casteleyn L. Mercury pollution in modern times and its socio-medical consequences. Sci Total Environ. 2019;654:720–34.

    Article  CAS  PubMed  Google Scholar 

  75. Ge J, Yuan YT, Yang H, Deng RJ, Li ZH, Yang Y. Smartphone-assisted colorimetric sensor based on single-atom Cu-C-N nanozyme for mercury (II) ions detection. Mater Today Chem. 2024;37: 102037.

    Article  CAS  Google Scholar 

  76. Song GC, Zhang Q, Liang S, Yao Y, Feng ML, Majid ZNB, et al. Oxidation activity modulation of a single atom Ce-N-C nanozyme enabling a time-resolved sensor to detect Fe3+ and Cr6+. J Mater Chem C. 2022;10:15656–63.

    Article  CAS  Google Scholar 

  77. Guo Q, Huang XR, Huang YJ, Zhang ZW, Li PW, Yu L. Fe-N-C single-atom nanozyme-linked immunosorbent assay for quantitative detection of aflatoxin B1. J Food Compos Anal. 2024;125: 105795.

    Article  CAS  Google Scholar 

  78. Zhang Y, Yuan X, Guo XY, Xu H, Zhang DX, Wu ZY, Zhang J. All-in-one zinc-doped Prussian blue nanozyme for efficient capture, separation, and detection of copper Ion (Cu2+) in complicated matrixes. Small. 2024;20:2306961.

    Article  CAS  Google Scholar 

  79. Liu Y, Ding D, Zhen YL, Guo R. Amino acid-mediated ‘turn-off/turn-on’ nanozyme activity of gold nanoclusters for sensitive and selective detection of copper ions and histidine. Biosens Bioelectron. 2017;92:140–6.

    Article  CAS  PubMed  Google Scholar 

  80. Xiong YH, Su LJ, He XC, Duan ZH, Zhang Z, Chen ZL, et al. Colorimetric determination of copper ions based on regulation of the enzyme-mimicking activity of covalent triazine frameworks. Sens Actuators B Chem. 2017;253:384–91.

    Article  CAS  Google Scholar 

  81. Lee S, Barin G, Ackerman CM, Muchenditsi A, Xu J, Reimer JA, et al. Copper capture in a thioether-functionalized porous polymer applied to the detection of Wilson’s disease. J Am Chem Soc. 2016;138:7603–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Li YY, Mu ZD, Yuan YH, Zhou J, Bai LJ, Qing M. An enzymatic activity regulation-based clusterzyme sensor array for high-throughput identification of heavy metal ions. J Hazard Mater. 2023;454: 131501.

    Article  CAS  PubMed  Google Scholar 

  83. Jing SY, Liu CM, Zheng J, Dong ZJ, Guo N. Toxicity of zearalenone and its nutritional intervention by natural products. Food Funct. 2022;13:10374–400.

    Article  CAS  PubMed  Google Scholar 

  84. Su ZH, Du T, Liang XF, Wang XZ, Zhao LF, Sun J, et al. Nanozymes for foodborne microbial contaminants detection: mechanisms, recent advances, and challenges. Food Control. 2022;141: 109165.

    Article  CAS  Google Scholar 

  85. Xing KY, Shan S, Liu DF, Lai WH. Recent advances of lateral flow immunoassay for mycotoxins detection. TrAC Trends Anal Chem. 2020;133: 116087.

    Article  CAS  Google Scholar 

  86. Chinese Pharmacopoeia Commission. Medicinal materials and cut crude drugs In: Pharmacopoeia of People’s Republic of China Part 1. Beijing: China Medical Science Press; 2020.

    Google Scholar 

  87. Lai WQ, Wei QH, Xu MD, Zhuang JY, Tang DP. Enzyme-controlled dissolution of MnO2 nanoflakes with enzyme cascade amplification for colorimetric immunoassay. Biosens Bioelectron. 2017;89:645–51.

    Article  CAS  PubMed  Google Scholar 

  88. Marchese S, Polo A, Ariano A, Velotto S, Costantini S, Severino L. Aflatoxin B1 and M1: biological properties and their involvement in cancer development. Toxins. 2018;10:214.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Zhang SY, Li H, Xia QH, Yang DZ, Yang YL. Zirconium-porphyrin-MOF-based oxidase-like nanozyme with oxygen vacancy for aflatoxin B1 colorimetric sensing. J Food Sci. 2024;89:3618–28.

    Article  CAS  PubMed  Google Scholar 

  90. Lu D, Jiang H, Zhang GY, Luo Q, Zhao Q, Shi XB. An in situ generated Prussian blue nanoparticle-mediated multimode nanozyme-linked immunosorbent assay for the detection of aflatoxin B1. ACS Appl Mater Interfaces. 2021;13:25738–47.

    Article  CAS  PubMed  Google Scholar 

  91. Zhao YK, Wang XF, Pan SX, Hong F, Lu P, Hu XB, et al. Bimetallic nanozyme-bioenzyme hybrid material-mediated ultrasensitive and automatic immunoassay for the detection of aflatoxin B1 in food. Biosens Bioelectron. 2024;248: 115992.

    Article  CAS  PubMed  Google Scholar 

  92. Rushing BR, Selim MI. Aflatoxin B1: a review on metabolism, toxicity, occurrence in food, occupational exposure, and detoxification methods. Food Chem Toxicol. 2019;124:81–100.

    Article  CAS  PubMed  Google Scholar 

  93. Chen PF, Li SL, Jiang CY, Wang ZP, Ma XY. A surface-enhanced Raman scattering aptasensor for output-signal detection of aflatoxin B1 based on peroxidase-like Cu2O@Au hybrid nanozyme. Food Biosci. 2023;54: 102885.

    Article  CAS  Google Scholar 

  94. Zhou SY, Xu LG, Kuang H, Xiao J, Xu CL. Immunoassays for rapid mycotoxin detection: state of the art. Analyst. 2020;145:7088–102.

    Article  CAS  PubMed  Google Scholar 

  95. Cai XF, Liang MJ, Ma F, Zhang ZW, Tang XQ, Jiang J, et al. Nanozyme-strip based on MnO2 nanosheets as a catalytic label for multi-scale detection of aflatoxin B1 with an ultrabroad working range. Food Chem. 2022;377: 131965.

    Article  CAS  PubMed  Google Scholar 

  96. Liang MJ, Cai XF, Gao YY, Yan HL, Fu JY, Tang XQ, et al. A versatile nanozyme integrated colorimetric and photothermal lateral flow immunoassay for highly sensitive and reliable Aspergillus flavus detection. Biosens Bioelectron. 2022;213: 114435.

    Article  CAS  PubMed  Google Scholar 

  97. Zhu X, Tang J, Ouyang XL, Liao Y, Feng HP, Yu JF, et al. A versatile CuCo@PDA nanozyme-based aptamer-mediated lateral flow assay for highly sensitive, on-site and dual-readout detection of aflatoxin B1. J Hazard Mater. 2024;465: 133178.

    Article  CAS  PubMed  Google Scholar 

  98. Zhao Q, Lu D, Zhang GY, Zhang D, Shi XB. Recent improvements in enzyme-linked immunosorbent assays based on nanomaterials. Talanta. 2021;223: 121722.

    Article  CAS  PubMed  Google Scholar 

  99. Lai WQ, Guo JQ, Wang YQ, Lin YX, Ye SA, Zhuang JY, Tang DP. Enzyme-controllable just-in-time production system of copper hexacyanoferrate nanoparticles with oxidase-mimicking activity for highly sensitive colorimetric immunoassay. Talanta. 2022;247: 123546.

    Article  CAS  PubMed  Google Scholar 

  100. Zheng XL, Sun LL, Zhao YN, Yang HL, Zhu YH, Zhang JX, et al. A fluorescence and colorimetric dual-mode immunoassay for detection of ochratoxin A based on cerium nanoparticles. Microchem J. 2024;201: 110419.

    Article  CAS  Google Scholar 

  101. Huang SY, Lai WQ, Liu BQ, Xu MD, Zhuang JY, Tang DP, Lin YX. Colorimetric and photothermal dual-mode immunoassay of aflatoxin B1 based on peroxidase-like activity of Pt supported on nitrogen-doped carbon. Spectrochim Acta A Mol Biomol Spectrosc. 2023;284: 121782.

    Article  CAS  PubMed  Google Scholar 

  102. Zhu HS, Quan Z, Hou HY, Cai Y, Liu WP, Liu YJ. A colorimetric immunoassay based on cobalt hydroxide nanocages as oxidase mimics for detection of ochratoxin A. Anal Chim Acta. 2020;1132:101–9.

    Article  CAS  PubMed  Google Scholar 

  103. Tang JD, Tian B, Tao XQ. A colorimetric aptasensor for detecting ochratoxin A based on label-free aptamer and gold nanozyme. Anal Sci. 2023;39:1623–6.

    Article  CAS  PubMed  Google Scholar 

  104. Obafemi BA, Adedara IA, Rocha JBT. Neurotoxicity of ochratoxin A: molecular mechanisms and neurotherapeutic strategies. Toxicology. 2023;497–498: 153630.

    Article  PubMed  Google Scholar 

  105. Chen MT, Huang XM, Chen YX, Cao YR, Zhang SS, Lei HT, et al. Shape-specific MOF-derived Cu@Fe-NC with morphology-driven catalytic activity: Mimicking peroxidase for the fluorescent- colorimetric immunosignage of ochratoxin. J Hazard Mater. 2023;443: 130233.

    Article  CAS  PubMed  Google Scholar 

  106. Chen MT, Liu ZX, Guan YY, Chen YX, Liu WP, Liu YJ. Zeolitic imidazolate frameworks-derived hollow Co/N-doped CNTs as oxidase-mimic for colorimetric-fluorescence immunoassay of ochratoxin A. Sens Actuators B Chem. 2022;359: 131609.

    Article  CAS  Google Scholar 

  107. Zhu HS, Cai Y, Qileng A, Quan Z, Zeng W, He KY, Liu YJ. Template-assisted Cu2O@Fe(OH)3 yolk-shell nanocages as biomimetic peroxidase: a multi-colorimetry and ratiometric fluorescence separated-type immunosensor for the detection of ochratoxin A. J Hazard Mater. 2021;411: 125090.

    Article  CAS  PubMed  Google Scholar 

  108. Ke CX, Wu Y, Song ZC, Zheng ME, Zhu HD, Guo HL, et al. A novel competitive fluorescence colorimetric dual-mode immunosensor for detecting ochratoxin A based on the synergistically enhanced peroxidase-like activity of AuAg NCs-SPCN nanocomposite. Food Chem. 2024;437: 137930.

    Article  CAS  PubMed  Google Scholar 

  109. Li M, Wang H, Yu XD, Jia XD, Zhu C, Liu JH, et al. A sensitive and simple competitive nanozyme-linked apta-sorbent assay for the dual-mode detection of ochratoxin A. Analyst. 2022;147:2215–22.

    Article  CAS  PubMed  Google Scholar 

  110. Zhu HS, Liu CH, Liu XX, Quan Z, Liu WP, Liu YJ. A multi-colorimetric immunosensor for visual detection of ochratoxin A by mimetic enzyme etching of gold nanobipyramids. Microchim Acta. 2021;188:62.

    Article  CAS  Google Scholar 

  111. Qiao WL, He BS, Yang J, Ren WJ, Zhao RY, Zhang YR, et al. Pt@AuNF nanozyme and horseradish peroxidase-based lateral flow immunoassay dual enzymes signal amplification strategy for sensitive detection of zearalenone. Int J Biol Macromol. 2024;254: 127746.

    Article  CAS  PubMed  Google Scholar 

  112. Huang N, Sheng W, Jin Z, Bai DM, Sun MY, Ren LS, et al. Colorimetric and photothermal dual-mode immunosensor based on Ti3C2Tx/AuNPs nanocomposite with enhanced peroxidase-like activity for ultrasensitive detection of zearalenone in cereals. Microchim Acta. 2023;190:479.

    Article  CAS  Google Scholar 

  113. Liu QW, Zhou LL, Xin SY, Yang QL, Wu W, Hou XD. Poly (ionic liquid) cross-linked hydrogel encapsulated with AuPt nanozymes for the smartphone-based colorimetric detection of zearalenone. Food Chem X. 2024;22: 101471.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Sun SM, Zhao R, Feng SM, Xie YL. Colorimetric zearalenone assay based on the use of an aptamer and of gold nanoparticles with peroxidase-like activity. Microchim Acta. 2018;185:535.

    Article  Google Scholar 

  115. Sidhu GK, Singh S, Kumar V, Dhanjal DS, Datta S, Singh J. Toxicity, monitoring and biodegradation of organophosphate pesticides: a review. Crit Rev Environ Sci Technol. 2019;49:1135–87.

    Article  CAS  Google Scholar 

  116. Richardson JR, Fitsanakis V, Westerink RHS, Kanthasamy AG. Neurotoxicity of pesticides. Acta Neuropathol. 2019;138:343–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Deng GQ, Chen HY, Shi Q, Ren LX, Liang K, Long WJ, et al. Colorimetric assay based on peroxidase-like activity of dodecyl trimethylammonium bromide-tetramethyl zinc (4-pyridinyl) porphyrin for detection of organophosphorus pesticides. Microchim Acta. 2022;189:375.

    Article  CAS  Google Scholar 

  118. Zhan XQ, Tang Y, Liu YY, Tao H, Wu YG. A novel colorimetric strategy for rapid detection of dimethoate residue in vegetables based on enhancing oxidase-mimicking catalytic activity of cube-shape Ag2O particles. Sens Actuators B Chem. 2022;361: 131720.

    Article  CAS  Google Scholar 

  119. Wang GX, Liu J, Dong HW, Geng LJ, Sun JS, Liu JJ, et al. A dual-mode biosensor featuring single-atom Fe nanozyme for multi-pesticide detection in vegetables. Food Chem. 2023;437: 137882.

    Article  PubMed  Google Scholar 

  120. Tai SM, Qian ZJ, Ren HX, Barimah AO, Peng CF, Wei XL. Highly selective and sensitive colorimetric detection for glyphosate based on β-CD@DNA-CuNCs enzyme mimics. Anal Chim Acta. 2022;1222: 339992.

    Article  CAS  PubMed  Google Scholar 

  121. Song DH, Tian T, Wang L, Zou YT, Zhao LZ, Xiao J, et al. Multi-signal sensor array based on a fluorescent nanozyme for broad-spectrum screening of pesticides. Chem Eng J. 2024;482: 148784.

    Article  CAS  Google Scholar 

  122. Sun YZ, Wei JC, Zou J, Cheng ZH, Huang ZM, Gu LQ, et al. Electrochemical detection of methyl-paraoxon based on bifunctional cerium oxide nanozyme with catalytic activity and signal amplification effect. J Pharm Anal. 2020;11:653–60.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Zhao FN, Li MY, Wang L, Wang M. A colorimetric sensor enabled with heterogeneous nanozymes with phosphatase-like activity for the residue analysis of methyl parathion. Foods. 2023;12:2980.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Gai PP, Pu L, Wang C, Zhu DQ, Li F. CeO2@NC nanozyme with robust dephosphorylation ability of phosphotriester: a simple colorimetric assay for rapid and selective detection of paraoxon. Biosens Bioelectron. 2022;220: 114841.

    Article  PubMed  Google Scholar 

  125. Wei JC, Yang Y, Dong JY, Wang SP, Li P. Fluorometric determination of pesticides and organophosphates using nanoceria as a phosphatase mimic and an inner filter effect on carbon nanodots. Microchim Acta. 2019;186:66.

    Article  Google Scholar 

  126. Wang T, Wang JN, Yang Y, Su P, Yang Y. Co3O4/reduced graphene oxide nanocomposites as effective phosphotriesterase mimetics for degradation and detection of paraoxon. Ind Eng Chem Res. 2017;56:9762–9.

    Article  CAS  Google Scholar 

  127. Luo M, Chen L, Wei JC, Cui XP, Cheng ZH, Wang T, et al. A two-step strategy for simultaneous dual-mode detection of methyl-paraoxon and Ni (II). Ecotoxicol Environ Saf. 2022;239: 113668.

    Article  CAS  PubMed  Google Scholar 

  128. Wu XC, Wei JH, Wu CY, Lv GP, Wu LN. ZrO2/CeO2/polyacrylic acid nanocomposites with alkaline phosphatase-like activity for sensing. Spectrochim Acta A Mol Biomol Spectrosc. 2021;263: 120165.

    Article  CAS  PubMed  Google Scholar 

  129. Wei JC, Yang LL, Luo M, Wang YT, Li P. Nanozyme-assisted technique for dual mode detection of organophosphorus pesticide. Ecotoxicol Environ Saf. 2019;179:17–23.

    Article  CAS  PubMed  Google Scholar 

  130. Wei JC, Xue Y, Dong JY, Wang SP, Hu H, Gao H, et al. A new fluorescent technique for pesticide detection by using metal coordination polymer and nanozyme. Chin Med. 2020;15:22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Yang YY, Hao SJ, Lei XM, Chen JN, Fang GZ, Liu JF, et al. Design of metalloenzyme mimics based on self-assembled peptides for organophosphorus pesticides detection. J Hazard Mater. 2022;428: 128262.

    Article  CAS  PubMed  Google Scholar 

  132. Cai Y, Zhu HS, Zhou WC, Qiu ZY, Chen CC, Qileng AR, et al. Capsulation of AuNCs with AIE effect into metal-organic framework for the marriage of a fluorescence and colorimetric biosensor to detect organophosphorus pesticides. Anal Chem. 2021;93:7275–82.

    Article  CAS  PubMed  Google Scholar 

  133. Feng YY, Hu P, Wang M, Sun XB, Pan W, Wang JP. Introducing Mn into ZIF-8 nanozyme for enhancing its catalytic activities and adding specific recognizer for detection of organophosphorus pesticides. Microchim Acta. 2023;190:437.

    Article  CAS  Google Scholar 

  134. Chang GR, Li SR, Wang YQ, Ran QX, Tan Q, Gou S, et al. Cu-C3N4 nanoenzyme-based freezing-dried bioactive capsule integrated with 3D-printed smartphone platform for visual detection of organophosphorus pesticides paraoxon in scallion. Sens Actuators B Chem. 2023;398: 134584.

    Article  Google Scholar 

  135. Zhu HJ, Liu BX, Wang MZ, Pan JM, Xu LZ, Hu PW, Niu XH. Amorphous Fe-containing phosphotungstates featuring efficient peroxidase-like activity at neutral pH: toward portable swabs for pesticide detection with tandem catalytic amplification. Anal Chem. 2023;95:4776–85.

    Article  CAS  PubMed  Google Scholar 

  136. Wang YM, Li M, Wang ZR, Xu J, Zhao JJ, Gao ZD, Song YY. Photothermal effect-enhanced peroxidase-like performance for sensitive detection of organophosphorus pesticides on a visual test strip. Chem Eng J. 2023;476: 146329.

    Article  CAS  Google Scholar 

  137. Yi YH, Zhou X, Liao DY, Hou JL, Liu HD, Zhu GB. High peroxidase-mimicking metal-organic frameworks decorated with platinum nanozymes for the colorimetric detection of acetylcholine chloride and organophosphorus pesticides via enzyme cascade reaction. Inorg Chem. 2023;62:13929–36.

    Article  CAS  PubMed  Google Scholar 

  138. Wang JN, Wang XY, Wang M, Bian QH, Zhong JC. Novel Ce-based coordination polymer nanoparticles with excellent oxidase mimic activity applied for colorimetric assay to organophosphorus pesticides. Food Chem. 2022;397: 133810.

    Article  CAS  PubMed  Google Scholar 

  139. Yang CL, Yu LH, Pang YH, Shen XF. A colorimetric sensing platform with smartphone for organophosphorus pesticides detection based on PANI-MnO2 nanozyme. Anal Chim Acta. 2023;1286: 342045.

    Article  PubMed  Google Scholar 

  140. Liu FN, Li Z, Wei HY, Xu P, Kang G, Zhu SC, et al. Coordinatively unsaturated cobalt single-atom nanozymes for visual pesticides detection by smartphone-based platform. Nano Res. 2023;17:2298–307.

    Article  Google Scholar 

  141. Ge J, Yang LK, Li ZH, Wan Y, Mao DS, Deng RJ, et al. A colorimetric smartphone-based platform for pesticides detection using Fe-N/C single-atom nanozyme as oxidase mimetics. J Hazard Mater. 2022;436: 129199.

    Article  CAS  PubMed  Google Scholar 

  142. Ouyang H, Tu XM, Fu ZF, Wang WW, Fu SF, Zhu CZ, et al. Colorimetric and chemiluminescent dual-readout immunochromatographic assay for detection of pesticide residues utilizing g-C3N4/BiFeO3 nanocomposites. Biosens Bioelectron. 2018;106:43–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Luo S, Gao JQ, Yuan HW, Yang J, Fan YH, Wang L, et al. Mn single-atom nanozymes with superior loading capability and superb superoxide dismutase-like activity for bioassay. Anal Chem. 2023;95:9366–72.

    Article  CAS  PubMed  Google Scholar 

  144. Zhang Y, Luo D, Zhou SK, Yang L, Yao WF, Cheng FF, et al. Analytical and biomedical applications of nanomaterials in Chinese herbal medicines research. Trends Anal Chem. 2022;156: 116690.

    Article  CAS  Google Scholar 

  145. Geng LG, Sun XD, Wang LD, Liu FP, Hu SQ, Zhao SL, Ye FG. Analyte-induced laccase-mimicking activity inhibition and conductivity enhancement of electroactive nanozymes for ratiometric electrochemical detection of thiram. J Hazard Mater. 2023;463: 132936.

    Article  PubMed  Google Scholar 

  146. Zhang CY, Peng LJ, Chen GY, Zhang H, Yang FQ. Investigation on the peroxidase-like activity of vitamin B6 and its applications in colorimetric detection of hydrogen peroxide and total antioxidant capacity evaluation. Molecules. 2022;27:4262.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Wang SN, Liu PC, Qin YM, Chen ZJ, Shen JC. Rapid synthesis of protein conjugated gold nanoclusters and their application in tea polyphenol sensing. Sens Actuators B Chem. 2015;223:178–85.

    Article  Google Scholar 

  148. Xu YX, Li PP, Hu XJ, Chen HY, Tang Y, Zhu Y, et al. Polyoxometalate nanostructures decorated with CuO nanoparticles for sensing ascorbic acid and Fe2+ ions. ACS Appl Nano Mater. 2021;4:8302–13.

    Article  CAS  Google Scholar 

  149. Sun S, Chen CY, Fu XY, Zhang YD, Wu XY, Hao JK, et al. Poly-β-cyclodextrin strengthen Pr6O11 porous oxidase mimic for dual-channel visual recognition of bioactive cysteine and Fe2+. Anal Bioanal Chem. 2024;416:1951–9.

    Article  CAS  PubMed  Google Scholar 

  150. Cai XF, Liang MJ, Ma F, Mohamed SR, Goda AA, Dawood DH, et al. A direct competitive nanozyme-linked immunosorbent assay based on MnO2 nanosheets as a catalytic label for the determination of fumonisin B1. Anal Methods. 2021;13:5542–8.

    Article  CAS  PubMed  Google Scholar 

  151. Fan YX, Li D, Xie XY, Zhang Y, Jiang L, Huang B, Yang XP. Flower-like L-Cys-FeNiNPs nanozyme aptasensor for sensitive colorimetric detection of aflatoxin B1. Microchem J. 2024;197: 109842.

    Article  CAS  Google Scholar 

  152. Zhang XB, Wang FY, Li ZR, Hu B, Zheng QY, Piao YZ, et al. Dual-mode electrochemical/colorimetric microfluidic sensor integrated tetrahedral DNA nanostructures with Au/Ni-Co LDH NCs nanozyme for ultrasensitive detection of aflatoxin B1. Sens Actuators B Chem. 2023;393: 134322.

    Article  CAS  Google Scholar 

  153. Wu L, Zhou M, Wang YS, Liu JM. Nanozyme and aptamer-based immunosorbent assay for aflatoxin B1. J Hazard Mater. 2020;399: 123154.

    Article  CAS  PubMed  Google Scholar 

  154. He ZY, Zhang JX, Liu M, Meng YH. Polyvalent aptamer scaffold coordinating light-responsive oxidase-like nanozyme for sensitive detection of zearalenone. Food Chem. 2024;431: 136908.

    Article  CAS  PubMed  Google Scholar 

  155. Zhu J, Xu WX, Yang Y, Kong RM, Wang JM. ssDNA-C3N4 conjugates-based nanozyme sensor array for discriminating mycotoxins. Microchim Acta. 2022;190:6.

    Article  Google Scholar 

  156. Xu X, Ma MY, Gao JX, Sun TX, Guo YH, Feng DM, Zhang L. Multifunctional Ni-NPC single-atom nanozyme for removal and smartphone-assisted visualization monitoring of carbamate pesticides. Inorg Chem. 2024;63:1225–35.

    Article  CAS  PubMed  Google Scholar 

  157. Wang Y, Yin L, Qu GX, Leung CH, Han L, Lu LH. Highly active single-atom nanozymes with high-loading iridium for sensitive detection of pesticides. Anal Chem. 2023;95:11960–8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Key Research and Development Program of China (No. 2022YFC2105700), Macau Science and Technology Development Fund (0024/2021/A1).

Author information

Authors and Affiliations

Authors

Contributions

LX drafted and revised the manuscript; FQY, MLL, and PL conceived and designed the review, and revised the manuscript; JJD, HZ, and DW performed literature searches and reviewed the information. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Peng Li or Feng-Qing Yang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that there are no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, L., Luo, ML., Dai, JJ. et al. Applications of nanomaterials with enzyme-like activity for the detection of phytochemicals and hazardous substances in plant samples. Chin Med 19, 140 (2024). https://doi.org/10.1186/s13020-024-01014-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13020-024-01014-9

Keywords