Tao Chen,
Jinmin Cao,
Xiaofang Bao,
Yu Peng,
Li Liu* and
Wensheng Fu*
Chongqing Key Laboratory of Green Synthesis and Applications, College of Chemistry, Chongqing Normal University, Chongqing 401331, China. E-mail: liliu1208@163.com; fuwensheng@cqnu.edu.cn
First published on 15th December 2021
Artificial nanozymes are designed for pursuing the functions of splendid catalytic efficiency and prominent selectivity of natural enzymes, meanwhile obtaining higher stability than that of natural enzymes. This emerging technology shows widespread application in the crossing field between nanotechnology and biomedicine. In this work, we employed a universal approach to fabricate a Co@N-CNTs hybrid nanocomposite as an oxidase mimic, in which fine Co nanoparticles were wrapped in N-doped carbon nanotubes, stacking on a hollow dodecahedron carbon skeleton. The synergistic effects of nanostructure engineering, N-doping and carbon coating, as well as the derived interfacial effect contribute to the glorious oxidase-like activity, stability and reusability. It can catalytically oxidize the colorless substrate 3,3′,5,5′-tetramethylbenzidine (TMB) to a blue oxidation product (ox-TMB). As a result, a colorimetric technique with excellent selectivity and sensitivity for detecting ascorbic acid (AA) with naked eyes was established, in view of specific inhibitory effects towards oxidation of TMB. Under optimal detection conditions, this method exhibits a good linearity ranging from 0.1 to 160 μM with a low limit of detection (LOD) of 0.076 μM. For practical applications, Co@N-CNTs hybrid catalyst as a mimic oxidase was used for the determination of AA in human serum, which yielded satisfactory results. This work may serve as a new research thought to guide the design of high-performance nanozymes and establish a sensing platform for the detection of AA.
At present, many noble metal nanomaterials have been proved to possess mimic oxide enzyme properties. For example, Naoki Toshima et al. proposed a method of loading gold nanoparticles on platinum nanoclusters forming a complex, which exhibited a conspicuous activity for the catalytic oxidation of glucose.7 Regrettably, the application of these noble metal nanomaterials is limited by their high cost and low abundance. Therefore, the cost-effective and easily accessible transition metal nanomaterials with the oxidase-like activity deserve to be more explored. Transition metal cobalt, due to its unique electronic structure of 3d74s2, has raised great concern in the area of nanoenzymes, especially cobalt particles in nanoscale may act as splendid oxidase-like enzymes.8 However, the nanosized metal particles are thermodynamically unstable when directly exposed to the reaction solution, considering the fact that the high surface energy easily leads to agglomeration and inactivation.3,9 As such, a method of coating metal nanoparticles with other materials is urgently required to make them relatively stable under traditional reaction conditions.
Recently, loading nanozymes onto carbonaceous supports such as carbon nanotubes,10 graphenes,11 and porous carbon12 is deemed as an effective strategy to prevent aggregation and ameliorate the catalytic activity of nanozymes. In particular, carbon nanotubes (CNTs) with porous structures can provide high specific surface area, admirable conductivity, desirable chemical stability and biocompatibility, which have a good ability to regulate the rapid electron transfer kinetics.13 However, the catalytic activity of carbon nanotubes is finite. Studies show that non-metallic dopants such as nitrogen, oxygen, boron, sulfur and phosphorus can observably enhance the oxidase-like catalytic activity of carbon materials.14,15 Among these, more electronegative nitrogen is an effective dopant, which can regulate the charge distribution on the surface of carbon materials and provide a large number of structural defects.16 Meanwhile, the lone-pair electrons on N atoms are conducive to anchoring the transition metal nanoparticles via coordination interaction. For example, Liu's group synthesized Co–N–C composites immobilized on N-doped carbon nanotubes for the catalytic oxidation of ethylbenzene. As the introduction of N heteroatoms and combination of CNTs promote the mass transfer between O2 and the substrate active site, the selectively catalytic oxidation performance of Co-N-Cx/CNTs for ethylbenzene was significantly improved.17
In addition, designing the novel morphology by nanostructure engineering is also a valid strategy to obtain artificial nanoenzymes with prominent catalytic activity, because the unique micro-structure could endow a large specific surface area and contribute to the rapid transfer of catalytic reaction-related species. For instance, Zhuang, et al. synthesized a hollow Co3O4@CuO nanocage, which presents high oxidase-like properties and is applied for dopamine biomolecular detection.18 Moreover, according to Chen's report, a Pt-loaded hollow mesoporous carbon nanosphere is proposed as the peroxidase mimic, which offers a huge hollow cavity and a mesoporous carbon skeleton.3 Its excellent peroxidase-like catalytic activity is ascribed to the unique hollow structure and the interfacial effect between metal Pt nanoparticles and the carbon skeleton; consequently, it could be applied in the colorimetric detection of ascorbic acid. Given the above, it is still a great challenge to construct hybrid nanomaterials with oxidase-mimic performance by combination of nanostructure engineering, element doping and loading of carbon materials.
Ascorbic acid (AA) is an important and special water-soluble vitamin. Its molecular structure is very simple, while the physical and chemical properties are unstable.19 The human body requires AA for certain functions to maintain the normal function of the human body. Due to AA deficiency, the synthesis of collagen will be affected, resulting in a cell connection barrier, meanwhile causing scurvy leading to decreased human immunity and body's emergency capacity.20 Therefore, it is very critical to precisely detect AA in food products and pharmaceutical formulation and for disease diagnosis. To date, various methods such as electrochemical methods,21 fluorescence spectroscopy,22 liquid chromatography,23 and chemiluminescence24 have been exploited for AA detection. Admittedly, these platforms enable the detection of AA with excellent sensitivities; however, there still exist some limitations, for example, they are time-consuming and these detection methods often demand complex detectors or operation. By contrast, the colorimetric assay has obvious advantages including simple operation, rapid response and easy readout with the naked eye, which is urgently needed for AA detection.
In this work, we synthesized a Co@N-CNTs hybrid nanomaterial by combination of nanostructural engineering, element doping and loading of carbon materials, in which Co nanoparticles were loaded on N-doped carbon nanotubes assembled in a hollow rhomboid dodecahedron framework. The hollow microstructure with open transfer channels contributes to the more exposed active sites and improves the molecular diffusion kinetics. Besides, coating N-doped carbon nanotubes (N-CNTs) onto Co nanoparticles endows them with good electron conductivity and prevents aggregation. Meanwhile, the hybrid structure creates a heterojunction at the junction between Co nanoparticles and N-CNTs, leading to an interfacial electric field, which further expedites the electron transfer. In addition, doping N atoms into a carbon skeleton could effectively create positive charge sites on nearby carbon atoms (Cδ+), which is beneficial to adsorb dissolved oxygen onto the surface of the Co@N-CNTs catalyst, and then the adsorbed O2 is transferred to the catalytic active sites (pyridinic N–Co and pyrrolic N–Co sites) to disintegrate into ROS species, promoting the oxidation of the TMB substrate. Therefore, the synergistic catalytic effect endows the Co@N-CNTs hybrid nanomaterial with outstanding oxidase-like activity, which enables to catalyze the oxidation of the TMB substrate to produce blue oxidation products. As a result, a colorimetric detection platform against ascorbic acid with bare eyes was established based on the oxidase-like activity of the Co@N-CNTs mimic enzyme and the inhibitory effect induced by AA. Considering the sensitivity, selectivity and long storage stability, the Co@N-CNTs catalyst shows great potential in the detection of AA in practical applications.
As displayed in Fig. 2a, Co@N-CNTs dodecahedron is successfully acquired on a large scale, revealing the feasibility of the synthetic method. The enlarged SEM image in Fig. 2b exhibits that Co@N-CNTs inherits the dodecahedron morphology of the precursor, but the surface becomes rougher along with slight shrinkage. This observation indirectly confirms the mass loss during pyrolysis. Fig. 2c shows the transmission electron microscopic (TEM) image of the Co@N-CNTs dodecahedron, in which the hollow structure is clearly observed. In addition, its surface seems to be decorated with many nanoparticles (Fig. 2d). Zooming in further would show that these nanoparticles with an average size of ∼10 nm are wrapped in well-organized nanotubes, and the nanotubes are mounted on the hollow dodecahedron framework, as shown in Fig. 2e. Meanwhile, a lot of holes can be observed in the dodecahedron framework, as marked by yellow circles in Fig. 2e, implying that the material owns a considerable surface area. In order to verify this point, the Brunauer–Emmett–Teller (BET) measurement was adopted. As illustrated in Fig. S2,† the N2 adsorption–desorption isotherm of the Co@N-CNTs dodecahedron displays the characteristic behavior of type IV with a distinct hysteresis loop, manifesting the porous structure. In addition, the pore size distribution curve (inset) shows that the size of pores is mainly concentrated at 4.04 nm, illustrating the existence of mesoporous. Additionally, the specific surface area determined from the BET model reaches up to 407.71 m2 g−1 with a pore volume of 0.48 cm3 g−1. Such a mesoporous structure with a huge specific surface area is conducive to the diffusion of reactants and the exposure of active sites, contributing to a desired catalytic activity.
What is more, the HRTEM (Fig. 2f) image shows a typical lattice fringe with a spacing of ∼0.206 nm corresponding to the (111) lattice planes of metallic Co. A few lattice fringes with a spacing of ∼0.338 nm are also detected at the edge of the nanoparticles, which are ascribed to the (002) lattice plane of graphitized carbon. The thickness of the carbon shell is 2–4 nm. Besides, the SAED image (inset) reveals the polycrystalline nature. The X-ray diffraction (XRD) measurement was utilized to further confirm the crystalline feature and chemical component of the Co@N-CNTs dodecahedron. In Fig. 2g, the XRD pattern displays three diffraction peaks located at 44.3°, 51.5°, and 75.9° corresponding to the (111), (200), and (220) facets of face-centered cubic (fcc) Co (PDF no. 01-1255). Interestingly, the diffraction peak (∼26°) of graphitized carbon is not found in the XRD pattern, but the characteristic spectral lines of two sequence regions of carbon materials appear in the Raman spectrum (Fig. S3†). This phenomenon suggests the poor crystalline nature of graphitized carbon in the Co@N-CNTs material. In Raman spectrum, two characteristic spectral lines representing different chemical structures appear in the first frequency of 1100–1800 cm−1, and the 2D band with a frequency dependent on laser excitation energy appears in the second frequency of 2400–3000 cm−1. There exists a D band for the A1g vibration mode located at 1360 cm−1, showing the presence of disordered carbon in the Co@N-CNTs structure. In addition, the characteristic G band at ca. 1583 cm−1 reflects the in-plane vibration of sp2 carbon with a high degree of graphitization. As it is well known, the different half-peak width and relative intensity of G and 2D peaks can reflect the interaction between multiple graphite layers.27 The relative strength of the 2D band at 2700 cm−1 is weaker than that of the G band, and the peak of the 2D band is obviously wider. This result indicates that multi-layer graphitized carbon was obtained by calcination of carbonaceous organic ligands at high temperatures. Besides, the ratio of ID/IG was determined to be 1.14, indicating a lot of structural defects in the carbon materials, which could provide plentiful active sites for the catalytic reaction.
The energy-dispersive X-ray spectroscopy (EDX) coupling with high-angel annular dark-field scanning transmission electron microscopy (HAADF-STEM) were also employed to characterize the microstructure and elemental distribution in the Co@N-CNTs dodecahedron. The result shown in Fig. 2h indicates that the C, N, O and Co elements exist in the Co@N-CNTs material, wherein the O element may come from the oxidation of surface. The elemental mapping images in Fig. 2i–l show that C and N elements evenly distribute throughout the region of dodecahedron, while the Co element only disperses on the region of nanoparticles, further testifying the novel structure that Co nanoparticles were wrapped in N-doped carbon nanotubes assembled in a hollow dodecahedron framework.
In order to further investigate the surface chemical components and chemical states of the Co@N-CNTs, the X-ray photoelectron spectroscopy (XPS)28,29 measurement was carried out. The survey-scan XPS spectrum of Co@N-CNTs (Fig. S4a†) exhibits the presence of C, N, Co and O elements. The asymmetrical C 1s spectrum in Fig. 3a is made up of three types of bonds, which are assigned to C–O at 288.6 eV, C–N at 285.9 eV, C–C or CC at 284.8 eV.16 In addition, three N types could be distinguished from the N 1s spectrum (Fig. 3b), including graphite N at 401.8 eV, pyrrolic N at 400.9 eV and pyridinic N at 398.8 eV.30 Pyridinic N and pyrrolic N usually act as metal-coordination sites due to the lone-pair electrons.31 Thus, they may serve as catalytic active sites to produce ROS, enhancing enzyme-like activities. This result confirms that N atoms are successfully doped into carbon materials after the high-temperature pyrolysis process. More importantly, owing to higher electronegativity, doping N atoms into a carbon skeleton could induce positive charge sites of adjacent carbon atoms. To confirm the positive charge sites of carbon atoms nearby N atoms in Co@N-CNTs, we prepared the Co@C composite and compared the C 1s XPS spectrum. As displayed in Fig. S4b,† it is not difficult to find that the C 1s XPS peak in the Co@N-CNTs is positively shifted (0.2 eV) compared to that in Co@C, indicating that the small electron density transfers from C atoms to N atoms. This observation proves that N-doping brings about the electron delocalization in the doped region. Therefore, the C and N atoms in the Co/N-CNTs nanozyme have partial positive charge (Cδ+) and negative charge (Nδ−). The positive charge sites on nearby carbon atoms (Cδ+) are beneficial to adsorb dissolved oxygen onto the surface of the Co@N-CNTs catalyst, thus promoting the oxidase-like activity. The Co 2p spectrum is illustrated in Fig. 3c, which is deconvoluted into four components. The fitting peaks at 795.4 and 778.6 eV are assigned to the binding energy of Co in a zero valence state. In addition, the XPS peaks at 796.7 and 780.4 eV are consistent with the Co(II) valence state, corresponding to the binding energy of CoO.32,33 Notably, unlike the XRD data, in which only metallic Co was identified, the presence of the Co–O peak in the XPS spectrum signifies that metal Co nanoparticles are susceptible to inevitable surface oxidation to form amorphous cobalt oxides.34 Besides, the satellite is located at 803.4 eV and 786.3 eV, and the Co–Nx peak is situated at 782.5 eV.35 The O 1s spectrum (Fig. 2d) exhibits three characteristic peaks corresponding to physisorbed-/chemisorbed water on the surface of the Co@N-CNTs material (O1 at 533.4 eV), the dissociative O2 adsorption on coordinatively unsaturated metal atoms (O2 at 532.0 eV), and the typical M–O bond derived from the inevitable surface oxidation (O3 at 531.0 eV). This result further verifies that the material may suffer from slight surface oxidation, in accordance with the observations in the Co 2p spectrum and EDX data. In combination with the TEM, XRD, EDX and XPS analysis, it is not difficult to conclude the successful preparation of the Co@N-CNTs dodecahedron.
![]() | ||
Fig. 3 High-resolution XPS spectrum of (a) C 1s, (b) N 1s, (c) Co 2p and (d) O 1s in the Co@N-CNTs hollow hybrid dodecahedron framework, respectively. |
Similar to natural enzymes, the catalytic performance of the mimic enzyme also follows a factor-dependent manner, which is closely related to the temperature, concentration, pH value and reaction time. Accordingly, we explored the influence of the four factors on the oxidase-like activity of the Co@N-CNTs catalyst. It is not hard to find that the absorbance at 652 nm fluctuates with the change in pH (2.4–6.0) and reaches the optimal value at pH = 4.0 (Fig. S8a and b†). The catalytic activity decreases sharply when the pH exceeds 4.8, owing to the poor solubility of the diamine structure of TMB under high pH conditions. Besides, in Fig. S8c and d,† the Co@N-CNTs catalyst exhibits an outstanding oxidase-like activity at a wide range of temperatures (25–60 °C). As we all know, natural enzymes can only perform the catalytic function within a limited temperature range in consideration of molecular stability; however, the Co@N-CNTs catalyst can operate under extreme conditions of temperature. For an easy operation, room temperature (∼25 °C) was selected as the test temperature. According to Fig. S8e and f,† the catalytic activity increases significantly with the increase in TMB concentration, when the TMB concentration is lower than 80 μg mL−1. However, the catalytic activity increases slowly when the TMB concentration is above 80 μg mL−1. Therefore, 80 μg mL−1 is deemed as the optimum TMB concentration. Similarly, the absorption values at 652 nm are positively correlated with the concentration of the catalyst (Fig. S8g and h†). In order to guarantee the accuracy of measurements, the absorption value should be kept at 1.2, thereby the optimized concentration of the Co@N-CNTs catalyst was determined to be 10 μg mL−1. In addition, as demonstrated in Fig. S8i and j,† the catalytic reaction was completed within 10 min (Fig. S8i and j†). Consequently, optimal conditions (pH = 4.0, T = 25 °C, CTMB = 80 μg mL−1, Ccatalyst = 10 μg mL−1 and Ttime = 10 min) were selected for the subsequent analysis.
Considering the practical utilization, stability and reusability are the two important metrics to judge the performance of nanozymes. As revealed in Fig. S9a,† the oxidase-like activity of the Co@N-CNTs catalyst remains basically steady within 25 days of storage period, displaying the excellent long-term stability. Furthermore, taking the weight loss in each cycle into consideration, the absorbance declines slightly after five reuse cycles (Fig. S9b†). However, the oxidase-like activity still maintains over 70.6% of the initial activity. Additionally, the SEM image (Fig. S10†) of the Co@N-CNTs catalyst after the stability test displays no obvious change, essentially explaining the structural stability. The satisfactory stability and reusability of the Co@N-CNTs catalyst make it possible for long-term applications in the field of catalysis.
Following that, the oxidase-like behavior of the Co@N-CNTs catalyst is profoundly investigated through steady-state kinetic analysis. The typical Michaelis–Menten curve is plotted by monitoring the TMB concentration in a reasonable range (1–120 μg mL−1) while maintaining the concentration of the catalyst at 10 μg mL−1 (Fig. 5a). Obviously, the reaction velocity is high at lower TMB concentrations while the reaction velocity does not proportionally increase with the TMB concentration at higher substrate concentrations demonstrating a saturation effect. The effect might be attributed to the occupancy of catalytically active sites in the Co@N-CNTs catalyst by the substrate at higher TMB concentrations. The corresponding double-reciprocal Lineweaver–Burk plot was obtained, as shown in Fig. 5b and fitted to Michaelis–Menten equation:
V0 = Vm[S]/(Km + [S]) | (1) |
To further figure out the possible reactive oxygen species (ROS) in the Co@N-CNTs catalytic system, several scavengers such as p-benzoquinone (PBQ), isopropanol (IPA) and tryptophan (TRP) were chosen as the free radical quenchers for probing superoxide anions (O2˙−), hydroxyl radicals (˙OH) and singlet oxygen (1O2), respectively.39 As revealed in Fig. 6a, the absorbance of the system with TRP is nearly the same with the blank control group. However, the oxidase-like activity of the Co@N-CNTs catalyst was dramatically decreased when the PBQ exists in the system. In addition, the catalytic activity was also slightly confined by introducing the IPA scavenger. Those results confirm the generation of O2˙− accompanied by small doses of ˙OH, in the oxide-mimicking catalytic process of Co@N-CNTs. In parallel, the electron paramagnetic resonance (EPR) spectrum in Fig. 6b provides with a similar conclusion. The EPR technique would probe the generation of ROS using 5,5-dimethyl-1-pyridine N-oxide (DMPO) as a spin trapping reagent. Six characteristic peaks for O2˙− adducts40 and four characteristic quarter peaks for ˙OH adducts41 were distinctly observed, further disclosing the production of O2˙− and ˙OH in the catalytic reaction solution.
To elucidate the role of O2 in the catalytic reaction, the oxidase-like activity of the Co@N-CNTs catalyst was further explored in O2 and N2-saturated solutions in Fig. 6c. A distinct difference value of characteristic absorbance at 652 nm was observed. Compared with the solution under air conditions, the absorption peak intensity of oxTMB increases under the O2-saturated condition, while the catalytic activity was significantly hindered after saturating with N2, indicating the vital participation of dissolved O2 in the oxidation process, as an electron acceptor for TMB oxidation. Consequently, we speculate the mechanism of the catalytic process: at first, the partial positive charge Cδ+ sites induced by the doping of electronegative N atoms adsorb massive dissolved oxygen onto the surface of the Co@N-CNTs catalyst; then the physically/chemically adsorbed O2 is transferred to the catalytic active sites (Pyridinic N–Co and pyrrolic N–Co sites) to disintegrate into ROS species (O2˙− and ˙OH); finally, the ROS species activate the oxidation reaction of substrates to emerge the oxidase-like activity. In general, the reduction product of the ROS species is probably H2O2. However, the production of H2O2 might trigger the peroxidase-mimicking activity of catalysts to accelerate the oxidation of TMB. Therefore, in order to distinguish the effect of peroxidase-like catalytic activity, further control experiments were carried out. As exhibited in Fig. 6d, the characteristic absorption peak of oxTMB slightly increases, when horse radish peroxidase (HRP) is added into the reaction solution with the Co@N-CNTs catalyst and TMB, suggesting the production of H2O2 after the catalytic reaction. Interestingly, when a dose of H2O2 was added into the reaction solution, the intensity of the characteristic absorption peak at 652 nm rises only slightly. On the contrary, when HPR and H2O2 are simultaneously filled into the reaction solution, an obvious increase in the characteristic absorption peak was detected. Those results testify that the Co@N-CNTs catalyst just owns weak peroxidase activity; hence, the oxidation of TMB mainly depends on its oxidase-like catalytic activity.
System | Signal output | Time of analysis | Linear range (μM) | LOD (μM) | References |
---|---|---|---|---|---|
Cu–Ag/rGO/TMB/H2O2 | Colorimetry | 50 min | 5–30 | 3.6 | 46 |
AgNPs/rGOb/GCE | Amperometry | No report | 10–800 | 9.6 | 47 |
GSH-AuNCs/H2O2/Fe2+ | Fluorescence | 5 min | 5–100 | 5 | 48 |
ClO−/TMB | Colorimetry | 30 s | 1–70 | 0.58 | 49 |
CuO nanoneedles/SPEs | Electrochemical | No report | 100–8000 | 88 | 50 |
GQDs/H2O2/HRP/catechol | Fluorescence | 40 min | 1.11–300 | 0.32 | 51 |
CoOOH/TMB | Colorimetry | 10 min | 0.25–60 | 0.057 | 52 |
MnO2/TMB | Colorimetry | 30 min | 0.1–20 | 0.098 | 53 |
Ag+/OPD | Fluorescence | 10 min | 0.05–1.0, | 0.01 | 54 |
1.0–40 | |||||
CdTe QDs | Fluorescence | No report | 0–800 | 3 | 55 |
PB NCs/ALP | Colorimetry | 15 min | 0.4–4.5 | 0.035 | 56 |
Au NCs | Fluorescence | 24 h | 1.5–10 | 0.2 | 57 |
PNCQDs/Cr(VI) | Fluorescence | 72 h | 5–200 | 1.35 | 58 |
Co@N-CNTs/TMB | Colorimetry | 10 min | 0.1–10, | 0.076 | This work |
10–160 |
In order to evaluate the anti-interference performance of the colorimetric method, the effect of the possible interfering ions including Na+, Mg2+, Al3+, Pb2+, Ca2+, Ni2+, Cr3+, Co2+, Zn2+, Ba2+, and K+ (in 100-fold excess of AA concentration) on AA detection was measured. As depicted in Fig. 8a, through comparing with the absorbance difference (ΔA) of the detection systems containing AA and other interfering substances, it can be found that the presence of interference has no obvious influence on the detection of AA, indicating that the proposed method has good anti-interference ability. In addition, considering the practical application of human serum testing, it is necessary to assess the selectivity of the AA detection platform. Therefore, the selectivity of the proposed colorimetric assay towards AA detection in the presence of multitudinous interfering species such as tryptophan, glycine, histidine, lysine, aspartic acid, serine, threonine, DL-methionine, arginine, valine, and cystine was further investigated by comparing the response of absorbance at 652 nm. Fig. 8b indicates that the absorbance decreased sharply with the blue color disappearance only after addition of AA, while other substances cannot affect the oxidase-like activity of the Co@N-CNTs catalyst. The high sensitivity, anti-interference performance and quick response towards AA endow this sensing platform with excellent specificity to directly detect AA in real samples. Accordingly, we conducted this colorimetric detection approach to measure AA in the human serum (see Fig. S13†). In addition, recovery experiments were executed using the standard addition method, where different concentrations of AA were spiked in human serum samples for analysis. In order to reduce the stochastic error, an average value was obtained by measuring three parallel trials. As exhibited in Table 2, satisfactory recoveries of 96.8–109.1% with a relative standard deviation (RSD) of three replicate detections for each sample lower than 3% were obtained. The above-mentioned results confirm the potential feasibility and wide application prospect of the as-proposed colorimetric method for AA detection in real samples.
Sample | Added | Found (μM) | Recovery (%) | RSD (%, n = 3) |
---|---|---|---|---|
Human serum | 0 | 0.049 | — | 0.1414 |
1 | 5 | 5.45 | 109.10 | 1.6971 |
2 | 10 | 9.68 | 96.80 | 0.6364 |
3 | 15 | 14.74 | 98.27 | 1.5556 |
4 | 20 | 19.89 | 99.45 | 2.1213 |
5 | 25 | 24.99 | 99.96 | 0.0707 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1ra07849f |
This journal is © The Royal Society of Chemistry 2021 |