Next Article in Journal
Macromolecule–Nanoparticle-Based Hybrid Materials for Biosensor Applications
Previous Article in Journal
A Novel Fluorescence-Based Microplate Assay for High-Throughput Screening of hSULT1As Inhibitors
Previous Article in Special Issue
Morphology Control of Zr-Based Luminescent Metal-Organic Frameworks for Aflatoxin B1 Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Metal–Organic Framework as a Novel Platform for Ratiometric Determination of Cyanide

School of Food and Biological Engineering, Jiangsu University, Zhenjiang 212013, China
*
Author to whom correspondence should be addressed.
Biosensors 2024, 14(6), 276; https://doi.org/10.3390/bios14060276
Submission received: 16 April 2024 / Revised: 13 May 2024 / Accepted: 22 May 2024 / Published: 27 May 2024
(This article belongs to the Special Issue Advanced Nanozyme for Biosensors)

Abstract

:
Metal–organic frameworks (MOFs) are frequently utilized as sensing materials. Unfortunately, the low conductivity of MOFs hinder their further application in electrochemical determination. To overcome this limitation, a novel modification strategy for MOFs was proposed, establishing an electrochemical determination method for cyanides in Baijiu. Co and Ni were synergistically used as the metal active centers, with meso−Tetra(4−carboxyphenyl)porphine (TCPP) and Ferrocenecarboxylic acid (Fc−COOH) serving as the main ligands, synthesizing Ni/Co−MOF−TCPP−Fc through a hydrothermal method. The prepared MOF exhibited improved conductivity and stable ratio signals, enabling rapid and sensitive determination of cyanides. The screen−printed carbon electrodes (SPCE) were suitable for in situ and real−time determination of cyanide by electrochemical sensors due to their portability, low cost, and ease of mass production. A logarithmic linear response in the range of 0.196~44 ng/mL was demonstrated by this method, and the limit of detection (LOD) was 0.052 ng/mL. Compared with other methods, the sensor was constructed by a one−step synthesis method, which greatly simplifies the analysis process, and the determination time required was only 4 min. During natural cyanide determinations, recommended readouts match well with GC−MS with less than 5.9% relative error. Moreover, this electrochemical sensor presented a promising method for assessing the safety of cyanides in Baijiu.

1. Introduction

Cyanide is a toxic compound containing a cyanide group (−CN) or cyanide ion (CN), such as KCN, NaCN, and HCN [1]. Cyanide is highly toxic, causing cell suffocation and death with blood flow [2]. Specific conditions, such as the presence of cyanogenic glycosides [3,4] in raw materials or the utilization of water sources and auxiliary materials contaminated with cyanide compounds, can lead to an increase in the cyanide content in alcoholic beverages. Baijiu is a type of alcoholic beverage that is very popular in China. During the high−temperature fermentation of Baijiu, cyanides are hydrolyzed and then react with ethanol, resulting in the formation of another carcinogen, ethyl carbamate [3,4]. According to national food safety standards, the cyanide content is limited to a maximum of 8 mg/L (calculated as HCN). In the updated “Detailed Rules for the Examination of Baijiu Production Licenses”, it is emphasized that the capability to conduct self−inspections of key food safety indicators, including cyanides and ethyl carbamate, should be possessed by enterprises. The monitoring of cyanide concentrations is instrumental in the assessment and control of potential health risks in alcoholic products.
In laboratories, the determination of cyanides in food is primarily conducted using ion chromatography [5], gas chromatography−mass spectrometry [6], and liquid chromatography [7]. Electrochemical sensors are widely used due to their advantages such as fast response, simple operation, high sensitivity, and low cost [8]. These sensors exhibit adequate precision and accuracy, and their instruments are relatively portable. Owing to their cost−effectiveness and potential for miniaturization [9], electrochemical sensors are considered a promising point−of−care (POC) platform. The stability and repeatability of the sensors can be greatly ensured by the proportional strategy [10]. Ratiometric electrochemical sensors quantify the analyte with a record ratiometric of two signals (one is from the analyte and the other is from the inner reference). A peak intensity ratio (Ianalyte/Iinner reference) is used as the measurement criteria for analytes [11].
Metal–Organic Frameworks (MOFs) are recognized as periodic network−structured organic–inorganic hybrid materials, assembled through coordination bonds between organic ligands and metal ions/clusters [12]. MOFs show high specific surface areas, rich porosity, excellent redox activity, and customizable structure and functionality [13,14,15,16,17]. The problem of poor electrical conductivity can be improved by metal doping combined with ligand regulation, which helps to adjust the composition elements, morphology, specific surface area, and pore size of MOFs [18], as well as improve their catalytic and electrochemical properties [19]. For example, ultrathin Co/Ni− MOF nanosheets were used as a platform for luminol−functionalized AgNPs to construct an immunosensor for tumor marker alpha−fetoprotein [20]. Bimetallic MOFs with superior electrochemical performance grown on electrospun nanofibers (PPNF@M−Ni MOF, M = Co, Zn, Cu, Fe) were prepared by controlling the incorporation of various types of metal ions [21]. ZnZr bimetallic MOFs were prepared as an aptasensor platform for detecting the cancer marker protein tyrosine kinase−7 (PTK7) [22]. The appropriate selection of metal ions and ligands is of fundamental importance to improve the electrochemical performance of MOFs. Ferrocene (Fc) and its derivatives are known for their unique redox and electrocatalytic properties [23,24,25], while porphyrins are known for their large π−conjugated structures and excellent photoelectric properties [26,27,28,29]. Ferrocene (Fc) and its derivatives and porphyrins all show excellent ligands for MOF.
In the present work, ferrocene carboxylic acid (Fc−COOH) and porphyrin (using TCPP as the doping ligand) are introduced into Ni/Co−MOF to enhance the electrochemical performance of Ni/Co−MOF (Figure 1A). Fc−COOH provides an additional metal catalytic center (Fe) that promotes the production of reactive oxygen species. Together with porphyrin as ligands, a greater number of unsaturated sites are formed, enhancing the electro−conductivity of Ni/Co−MOF−TCPP−Fc. After synthesis of Ag on the surface, the metals in Ag−Ni/Co−MOF−TCPP−Fc react with cyanide to form a metal complex, leading to inhibition of the electron transfer on the electrode surface. As the concentration of cyanides increases, the complexation reaction is intensified, leading to a decrease in the electrochemical signal, while the reference signal remains stable (Figure 1B). The ratio value of ICN−/I serves as an excellent indicator for cyanide evaluation. This sensor greatly simplifies the analysis process, improves the determination efficiency, and exhibits the potential to be used for cyanide monitoring in Baijiu.

2. Materials and Methods

2.1. Main Reagents and Materials

Potassium chloride (KCl), anhydrous ethanol, potassium ferricyanide (K3[Fe(CN)6]), potassium ferrocyanide (K4[Fe(CN)6]), and other reagents were supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Polyvinylpyrrolidone, terephthalic acid, ferrocene carboxylic acid (Fc−COOH), N,N−Dimethylformamide (DMF), 2−methylimidazole (2−Mim), Cobaltous nitrate hexahydrate (Co(NO3)2·6H2O), and nickel nitrate hexahydrate (Ni(NO3)2·6H2O) were provided by Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). Meso−tetra(4−carboxyphenyl)porphine (97%) was supplied by Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China). Phosphate buffered saline (PBS, 0.1 M, pH 7.4) and cyanide standards in water were provided by Shanghai Anpel Laboratory Technologies Co., Ltd. (Shanghai, China). The water used was ultrapure water. Screen−Printed Carbon (SPCE) Electrode was purchased from Poten (Beijing, China).

2.2. Main Instruments and Equipment

Composite microstructures were explored with scanning electron microscopy (SEM, JSM−7800M, JEOL, Tokyo, Japan). Bonding characteristics were examined with an X−ray photoelectron spectrometer (XPS, K−Alpha, Thermo, Waltham, MA, USA), using Al−Kα radiation with 1486.6 eV energy. An X−ray diffractometer (XRD, D8 ADVANCE, Bruker, Bremen, Germany) was employed with crystal phases of TCPP, Fc−COOH, and Ni/Co−MOF−TCPP−Fc at a 2θ angle from 5° to 80° with a Cu−Kα radiation source (λ = 1.5405 Å). Fourier transform infrared spectra (FT−IR) were studied with an FT−IR analyzer (Nicolet is50, Nicolet, Wausau, WI, USA). N2 physisorption isotherms were measured on a surface area analyzer (NOVA3000e, Quantachrome, Boynton Beach, FL, USA), in which the Brunauer–Emmett–Teller (BET) model was applied to estimate specific surface area and the Barrett–Joyner–Halenda (BJH) model was employed to analyze pore size. All electrochemical measurements including differential pulse voltammetry (DPV) and cyclic voltammetry (CV) were evaluated with a CHI potentiostat (660E, Shanghai Chen Hua Instrument Co., Ltd., Shanghai, China).

2.3. Preparation of Electrochemical Sensor

Ni/Co−MOF−TCPP−Fc was synthesized via a hydrothermal method following the previously reported procedures with slight modifications [30,31]. A total of 0.1866 g Co(NO3)2·6H2O and 0.2958 g Ni(NO3)2·6H2O, along with 0.5 g polyvinylpyrrolidone, were dissolved in 60 mL solution including DMF and ethanol mixture (3:1, v/v) to prepare solution A. Subsequently, 200 mg TCPP was dissolved in 20 mL solution including DMF and ethanol mixture (3:1, v/v) with thorough stirring until fully dissolved, resulting in solution B. Then, 0.6646 g terephthalic acid and 0.05 g Fc−COOH were dissolved in 20 mL DMF, and the mixture was stirred for 25 min until fully dissolved to obtain solution C.
Then, under vigorous stirring, solutions B and C were added dropwise to solution A. After ultrasonication of the mixture for 25 min, it was transferred to a 100 mL polytetrafluoroethylene−lined autoclave and reacted at 105 °C for 10 h. Upon naturally cooling to room temperature, the solid was collected by centrifugation (4000 rpm, 10 min), washed several times alternately with DMF and anhydrous ethanol, and then dried at 60 °C for 12 h to obtain the Ni/Co−MOF−TCPP−Fc. Then, 100 mg Ni/Co−MOF−TCPP−Fc was dispersed in 50 mL of distilled water and mixed ultrasonically with AgNO3 (10 mL 2 mM) solution, and then NaBH4 solution (5 mL 10 mM) was added drop by drop during agitation. Ag−Ni/Co−MOF−TCPP−Fc was obtained after being magnetically stirred at room temperature for 2 h, washed with ethanol three times, and dried at 60 °C. A total of 10 µL of the prepared Ag−Ni/Co−MOF−TCPP−Fc (1 mg/mL) was drop−coated onto the surface of a screen−printed carbon electrode (SPCE) and dried at room temperature to obtain the Ag−Ni/Co−MOF−TCPP−Fc/SPCE.

2.4. Electrochemical Experiment

A three−electrode system was used: a working electrode (material: carbon, electrode diameter: 4 mm), a counter electrode (material: carbon), and a reference electrode (material: Ag/AgCl). Gradual modification of the electrode was verified by CV and DPV. The CV scan rate was set to 100 mV/s, with scanning windows of −0.6~1 V. In DPV measurements, the potential window, pulse amplitude, pulse width, and pulse period were set to −0.8~0.6 V, 50 mV, 16.7 ms, and 100 ms, respectively.

2.5. Determination of Real Samples

Natural samples including Baijiu, fermented grains, river water, and mineral water were involved to examine the proposed method. Three kinds of Baijiu (sauce−flavor, strong−flavor, and light−flavor) and mineral water were purchased at Jimailong supermarket (Zhenjiang, China). Baijiu samples were diluted to less than 20% alcohol to be measured. Mineral water samples could be directly used for detection. River samples were collected from the campus. They were centrifuged (10,000 rpm, 10 min) and filtered with 0.22 µm microporous membrane to remove undissolved substances. Fermented grains were provided by a winery in Jiangsu Province. In a centrifuge tube, 1 g fermented grains and 50 mL pure water were sonicated for 20 min and then centrifuged for 5 min (4000 rpm). The supernatant was taken for testing. All samples were stored at 4 °C before further analysis. These samples were tested by dropping them onto the Ag−Ni/Co−MOF−TCPP−Fc/SPE sensor.
These samples were also examined with Gas Chromatography−Mass Spectrometry (GC−MS, TQ8040, SHIMADZU, Kyoto, Japan) and recovery assays concurrently. Prior to GC−MS analyses, pretreated samples were filtered with a 0.22 µm microporous membrane. The analytical column was a capillary column (30 m × 0.25 mm id × 0.25 μm film thickness). The flow rate of helium as carrier gas was 1.0 mL/min. The oven temperature program began at 40 °C (held for 2 min), was raised to 60 °C (held for 2 min) at 20 °C/min, and then increased to the final temperature of 200 °C at 50 °C/min (held for 2 min). The inlet temperature was 200 °C, the shunt ratio was 10:1, and the column flow rate was 1.5 mL/min. The ion source temperature of the mass spectrometry was 230 °C, the interface temperature was 220 °C, and the solvent delay time was 1 min. The monitoring mode was ion scanning, the scanning ion m/z was set to 61, 63, and 35, and the quantitative ion m/z was set to 61.
During the recovery test, a pretreated natural sample was analyzed with the proposed sensor to obtain the original CN amount (the 1st analysis). Then, it was spiked with 0.1 µg/mL CN for the 2nd analysis.

3. Results and Discussion

3.1. Morphological Study and Chemical Composition of Ni/Co−MOF−TCPP−Fc

The Ni/Co−MOF−TCPP−Fc, which was synthesized via the direct hydrothermal method, exhibited an irregular block−like structure with small particles on its surface, leading to a relatively rough texture. The size of Ni/Co−MOF−TCPP−Fc was in the range of 14~25 µm. This texture is beneficial for increasing catalytic sites. As evidenced by the SEM image, small particles were prominently present on the surface of the material. These small particles were revealed to be compounds of Co and Ni through surface characterization using XPS and mapping. A uniform distribution of Fe, Ni, and Co atoms within the Ni/Co−MOF−TCPP−Fc was demonstrated by elemental mapping (Figure 2A).
The elemental composition of Ni/Co−MOF−TCPP−Fc was further analyzed with an X−ray photoelectron spectrometer. The presence of Fe, Co, Ni, O, H, and C elements in the full XPS spectrum was consistent with the results of the mapping analysis (Figure 2B). The O 1s spectrum of the high−resolution XPS, shown in Figure 2C, corresponded to C=O and C−O for binding energies at 531.49 eV and 532.99 eV, respectively, which were mainly derived from the carboxyl group of the terphenyl acid ligand [32]. The presence of Ni2+ was confirmed by the Ni 2p3/2 (855.79 eV) and Ni 2p1/2 (873.09 eV) peaks in the Ni 2p high−resolution XPS map (Figure 2D), and the peaks with binding energies of 879.09 eV and 861.29 eV were satellite peaks of Ni [33]. The binding energies at 711.69 eV and 725.09 eV correspond to Fe 2p3/2 and Fe 2p1/2, respectively, both from Fe2+, Fe3+ characteristic peaks in the ferrocene structure, and the introduction of Fc−COOH was confirmed (Figure 2E). The Co 2p spectrum of high−resolution XPS is shown in Figure 2F. The Co 2p emission spectrum had two spin−orbit duplex states, with Co2+ ions with binding energies of 781.19 eV and 796.59 eV and Co3+ ions with binding energies of 780.19 eV and 795.59 eV. There were two oscillatory satellite peaks of 803.79 eV and 785.79 eV, respectively [34].

3.2. Synthetic Properties of Ni/Co−MOF−TCPP−Fc

The crystal structure of Ni/Co−MOF−TCPP−Fc was characterized using XRD, and as shown in Figure 3A, the results were similar to those previously reported for Ni/Co−MOF [34]. The arrangement of the metals Co and Ni as framework nodes within the crystal structure is indicated by the absence of distinct metal peaks in Ni/Co−MOF−TCPP−Fc. Additionally, the crystallinity of Ni/Co−MOF may be affected by the incorporation of the ligands Fc−COOH and TCPP, leading to a deterioration in the crystal form and thus exposing more active sites.
The coordination between the ligands and the MOF material was analyzed using FT−IR (Figure 3B). The attribution of the absorption peak at 1006 cm−1 in Ni/Co−MOF−TCPP−Fc to the stretching vibration of the Me−N bond was indicative of the coordination between metal atoms and N atoms, suggesting that the formation of Co−N [35] and Ni−N [36] bonds was likely to enhance the enzyme−like catalytic activity of the material. The formation of metal–carboxylate bonds was proven by attributing the absorption bands between 1630~1480 cm−1 to the asymmetric and symmetric stretching vibrations of the carboxylate group (−COO−) that was bound to metals (Co or Ni) after coordination. For Ni/Co−MOF−TCPP−Fc, the absorption peak of TCPP at 1270 cm−1 was weakened, indicating that the hydrogen on −OH was replaced by metal ions and formed an Me−O bond. The reduction in peak intensity at 1600 and 1690 cm−1 was possibly attributed to alterations in the electronic structure that occurred when porphyrin was incorporated as a ligand within the MOF [37,38]. A broad absorption band was observed for the pure substance of Fc−COOH between 3310~2500 cm−1 due to the O−H stretching vibration of the carboxyl group in Fc−COOH, and a stretching vibration peak of the C=C in cyclopentadiene was observed at 1475 cm−1 [39,40]. However, in Ni/Co−MOF−TCPP−Fc, a significant weakening of the absorption band and the disappearance of characteristic peaks were observed, indicating that coordination had been established between Fc−COOH and the metal ions (Co and Ni), instead of Fc−COOH entering the pores of MOFs in its free state.
Subsequently, the pore size and specific surface area of Ni/Co−MOF doped with Fc−COOH and TCPP were studied combined with nitrogen adsorption–desorption isotherms. The results are shown in Figure 3C. According to international Union of Pure and Applied Chemistry (IUPAC) classification, it belonged to type I adsorption isotherm, and there were many microporous structures in the material. The BET surface area of Ni/Co−MOF−TCPP−Fc was measured at 1111.35 m2/g, which endowed the material with enhanced catalytic performance due to its large specific surface area. The pore size distribution is shown in Figure 3C. The pore size distribution was not uniform, with a large number of 2.4 nm pores. The micropores of the material were analyzed, and a large number of 0.96 nm pores were included. The addition of Fc−COOH and TCPP to Ni/Co−MOF−TCPP−Fc resulted in a high surface area and abundant microporosity [41], enhancing the mass transfer efficiency in the catalytic process.

3.3. Electrochemical Characterization of Gradual Modification

The electrochemical performance of three types of MOFs (Co−MOF, Ni−MOF, and Ni/Co−MOF) was tested using a standard three−electrode system at a scan rate of 100 mV/s in a solution of 5 mM K3[Fe(CN)6] and 0.1 M KCl. From Figure 4A, within the −0.6~1 V range, the oxidation−reduction potentials of Co−MOF and Ni−MOF were relatively close. Due to the poor conductivity of MOF materials, the peak currents were weaker compared to the bare electrode. Under the cooperative effect of Co and Ni, in the presence of the ligands TCPP and Fc, the best CV response was exhibited by Ni/Co−MOF, with two reduction peaks appearing at −0.2 V and 0.46 V, indicating an enhancement in electrochemical activity. A more conductive electronic structure was formed as Fc−COOH acted as a defect−linker that can coordinate with metal ions through its carboxyl groups. Furthermore, the electron transport capability of MOFs was also improved by the introduction of the π−conjugated system by the TCPP ligand, enhancing the speed and current response of electrochemical reactions. Under scan rates ranging from 10 to 100 mV/s, it was shown that the oxidation−reduction signals of Ag−Ni/Co−MOF−TCPP−Fc significantly increased (Figure 4B). The peak currents of oxidation and reduction showed a good linear relationship with the scan rate, indicating a diffusion−controlled process on the electrode surface. Moreover, the peak separations begin to increase with increasing scan rate. This is consistent with the limitations which arise from charge transfer kinetics. This further confirmed that ion diffusion controls the electrochemical kinetics in the redox reaction (Figure 4C).
The DPV responses of electrodes modified with different materials were studied in 0.1 M PBS buffer. In Figure 4D, a distinct reduction peak at −0.5 V was observed for Ni/Co−MOF−TCPP, attributed to its polycyclic structure and multiple oxidizable redox centers, resulting in a significant current response. Two distinct reduction peaks at 0.26 V and −0.53 V (red curve) were produced by Ag−Ni/Co−MOF−TCPP−Fc, with the peak current reaching a maximum. This enhancement in the electrochemical response was due to the provision of abundant active sites by macrocyclic ligands such as TCPP and Fc to Ag−Ni/Co−MOF. Besides, for the blank solution, two significant peaks were displayed by the sensor. The bare electrode had no obvious electrical response to cyanide (Figure 4E). With the increase in cyanide concentration, the MOF peak was significantly suppressed, while the TCPP peak remained stable (Figure 4F). This phenomenon occurred because strong complexes with metal ions were formed by cyanide, which inhibited electron transfer on the electrode surface, leading to a reduction in the electrical signal.
Ag was the main chemical reaction element, and the metals Co and Ni were the secondary reaction ligands. The silver element may dissolve in the form of AgO and AgO+ in the anode scan. At the end of each CV cycle, it was reduced at the cathode in other forms of silver oxide and redeposited on the surface as Ag metal. The metal center of the MOF also participated in the reaction.
A g O + 2 C N + H 2 O + e A g C N 2 + 2 O H
2 A g + 4 C N + 1 2 O 2 + H 2 O 2 A g C N 2 + 2 O H
C o 2 + + 6 C N C o ( C N ) 6 4
N i 2 + + 4 C N N i ( C N ) 4 2
Hence, a proportional relationship was established between ICN− and I.

3.4. Electrochemical Determination of CN

The relationship between the current ratio and pH was showed in Figure 4. The current ratio gradually increased in the range of 2.8–7.4. This was because HCN was the main form in acidic environment, and acidic conditions inhibited the conversion of HCN to CN. As pH increased, more HCN dissociated to form CN. CN showed strong coordination ability and could form stable cyanogen complexes with various metal ions. Thus, in aiming to obtain better sensitivity for the assays of CN, pH 7.4 was selected for the subsequent analysis.
Under optimal conditions, the DPV curves of the sensor were measured (Figure 5B). With the increase in cyanide concentration, a gradual decrease in the designed sensor ICN− was observed, whereas I was found to remain almost unchanged. This phenomenon was attributed to the formation of stable metal complexes by metal ions with cyanide, which lack conductivity and hinder electron transfer on the electrode surface, further leading to a decrease in signal. In Figure 5C, a good linear relationship between ICN/I and cyanide concentration in the range of 0.196~44 ng/mL was demonstrated by the standard curve, with a correlation coefficient of 0.998. The linear regression equation established was y = 63.10612x + 0.43239 (x is cyanide concentration), and the limit of detection (LOD) was 0.052 ng/mL (S/N = 3). Compared to previously reported sensors (Table 1), a better linear range and determination limit for cyanide were exhibited by the developed sensor.

3.5. Specificity and Stability of the Sensor

Various ions existed in the Baijiu matrix. The selectivity of the sensor towards cyanide was investigated by verifying the presence of potential interfering ions in Baijiu, including Br, CO32−, F, SO42−, Cl, and H2PO4. As shown in Figure 6A, compared to cyanide, the current response of other ions was less than 5.8%. The electrochemical determination of cyanide was not significantly impacted by the coexistence of interfering ions, indicating that the sensor possesses good anti−interference capability for the determination of cyanide. The repeatability of the sensor was evaluated through the determination of 2 ng/mL CN using the same Ag−Ni/Co−MOF−TCPP−Fc sensor. Good repeatability was shown by the current response from six repeated measurements, with only a 4.5% loss (Figure 6B).

3.6. Evaluation of CN in Fermented Grains and Baijiu

To verify the applicability of the Ag−Ni/Co−MOF−TCPP−Fc sensor, the designed sensor was applied to the determination and analysis of cyanide in fermented grains, Baijiu, river water, and mineral water. Baijiu samples were diluted to an alcohol content of below 20%, followed by the addition of different concentrations of cyanide standards (µg/mL). As can be seen from Table 2 and Table 3, the detected values were close to the added standard values, with recovery rates ranging between 98.5% and 103.3%. Besides, the relative standard deviations were low, ranging from 2.5% to 5.9%. The t and F statistical tests showed that there was no significant difference between the prepared electrochemical ratio sensor and GC−MS for the same sample. The electrochemical sensor is appropriate for cyanide assay in natural samples.

4. Conclusions

In summary, a simple and highly sensitive electrochemical sensor was developed based on the synthesis of a novel Ag−Ni/Co−MOF−TCPP−Fc for the determination of cyanides in Baijiu. Moreover, cyanide could be detected by the designed sensor within 4 min. The prepared Ag−Ni/Co−MOF−TCPP−Fc sensor exhibited an integrated reference signal, effectively establishing a ratio strategy that enhanced the stability of the sensor. The electron transport efficiency of Ag−Ni/Co−MOF was significantly improved by the ligands Fc−COOH and TCPP. It was indicated by experimental results that LOD was sufficient for monitoring cyanide levels in Baijiu, providing a new opportunity for the determination of cyanides in Baijiu.

Author Contributions

Conceptualization, Z.S. and X.Z.; methodology, C.L.; validation, Y.G. and Y.Z.; formal analysis, W.G.; resources, Z.S. and X.Z.; data curation, Z.W. and C.L.; writing—original draft preparation, Z.W.; funding acquisition, Z.S. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Research and Development Program of Jiangsu Province, BE2020312.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors gratefully acknowledge the financial support provided by the Key Research and Development Program of Jiangsu Province, BE2020312. The authors wish to thank the timely help given by Qunyong Tang and Jianfeng Wu in the investigation.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bai, C.B.; Zhang, J.; Qin, Y.X.; Huang, H.N.; Xia, Z.N.; Zheng, Q.J.; Dai, H.L.; Lu, P.K.; Miao, H.; Qu, C.Q.; et al. Precise probe design based ESIPT coupled AIE mechanism toward endogenous cyanide in food detection and bioimaging. Chem. Eng. J. 2022, 443, 136445. [Google Scholar] [CrossRef]
  2. Pacher, P. Cyanide emerges as an endogenous mammalian gasotransmitter. Proc. Natl. Acad. Sci. USA 2021, 118, e2108040118. [Google Scholar] [CrossRef] [PubMed]
  3. Shen, T.; Wu, Q.; Xu, Y. Biodegradation of cyanide with Saccharomyces cerevisiae in Baijiu fermentation. Food Control 2021, 127, 108107. [Google Scholar] [CrossRef]
  4. Qin, Y.; Duan, B.; Shin, J.A.; So, H.J.; Hong, E.S.; Jeong, H.G.; Lee, J.H.; Lee, K.T. Effect of fermentation on cyanide and ethyl carbamate contents in cassava flour and evaluation of their mass balance during lab−scale continuous distillation. Foods 2021, 14, 1089. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, D.Y.; Peng, Y.K.; Yan, J. Detection of cyanide in pollution−free livestock product breeding water by ion chromatography. Asian Agric. Res. 2018, 10, 34–36. [Google Scholar]
  6. Xu, J.; Tong, H.W.; Yan, X.; Du, S.; Yao, Z.; Liu, S. Sensitive determination of cyanide in cigarette smoke by capillary GC with a microECD. Chromatographia 2006, 64, 609–612. [Google Scholar] [CrossRef]
  7. Kang, H.I.; Shin, H.S. Derivatization method of free cyanide including cyanogen chloride for the sensitive analysis of cyanide in chlorinated drinking water by liquid chromatography−tandem mass spectrometry. Anal. Chem. 2015, 87, 975–981. [Google Scholar] [CrossRef] [PubMed]
  8. Li, C.; Zhang, X.A.; Tang, Q.Y.; Guo, Y.Q.; Zhang, Z.; Zhang, W.; Zou, X.B.; Sun, Z.B. Molecularly imprinted electrochemical sensor for ethyl carbamate detection in Baijiu based on “on−off” nanozyme−catalyzing process. Food Chem. 2024, 453, 139626. [Google Scholar] [CrossRef] [PubMed]
  9. Dowlatshahi, S.; Abdekhodaie, M.J. Electrochemical prostate−specific antigen biosensors based on electroconductive nanomaterials and polymers. Clin. Chim. Acta 2021, 516, 111–135. [Google Scholar] [CrossRef]
  10. Su, X.Y.; Wang, H.; Wang, C.Q.; Zhou, X.; Zou, X.B.; Zhang, W. Programmable dual−electric−field immunosensor using MXene−Au−based competitive signal probe for natural parathion−methyl detection. Biosens. Bioelectron. 2022, 214, 114546. [Google Scholar] [CrossRef]
  11. Monteiro, M.C.; Winiarski, J.P.; Santana, E.R.; Szpoganicz, B.; Vieira, I.C. Ratiometric electrochemical sensor for butralin determination using a quinazoline−engineered prussian blue analogue. Materials 2023, 16, 1024. [Google Scholar] [CrossRef] [PubMed]
  12. Lal, S.; Singh, P.; Singhal, A.; Kumar, S.; Gahlot, A.P.; Gandhi, N.; Kumari, P. Advances in metal–organic frameworks for water remediation applications. RSC Adv. 2024, 14, 3413–3446. [Google Scholar] [CrossRef] [PubMed]
  13. Shu, H.; Lai, T.R.; Yang, Z.C.; Xiao, X.C.; Chen, X.M.; Wang, Y.D. High sensitivity electrochemical detection of ultra−trace imidacloprid in fruits and vegetables using a Fe−rich FeCoNi−MOF. Food Chem. 2023, 408, 135221. [Google Scholar] [CrossRef] [PubMed]
  14. Li, Y.Z.; Fu, Z.H.; Xu, G. Metal−organic framework nanosheets: Preparation and applications. Coord. Chem. Rev. 2019, 388, 79–106. [Google Scholar] [CrossRef]
  15. Ahmadi, M.; Ebrahimnia, M.; Shahbazi, M.A.; Keçili, R.; Ghorbani−Bidkorbeh, F. Microporous metal–organic frameworks: Synthesis and applications. J. Ind. Eng. Chem. 2022, 115, 1–11. [Google Scholar] [CrossRef]
  16. Yu, K.; Li, M.J.; Chai, H.N.; Liu, Q.; Hai, X.; Tian, M.W.; Qu, L.J.; Xu, T.L.; Zhang, G.Y.; Zhang, X.J. MOF−818 nanozyme−based colorimetric and electrochemical dual−mode smartphone sensing platform for in situ detection of H2O2 and H2S released from living cells. Chem. Eng. J. 2023, 451, 138321. [Google Scholar] [CrossRef]
  17. Wang, Y.; Yan, J.H.; Wen, N.; Xiong, H.J.; Cai, S.D.; He, Q.Y.; Hu, Y.Q.; Peng, D.M.; Liu, Z.B.; Liu, Y.F. Metal−organic frameworks for stimuli−responsive drug delivery. Biomaterials 2020, 230, 119619. [Google Scholar] [CrossRef]
  18. Lu, Y.; Liu, C.Z.; Mei, C.T.; Sun, J.S.; Lee, J.; Wu, Q.L.; Hubbe, M.A.; Li, M.C. Recent advances in metal organic framework and cellulose nanomaterial composites. Coord. Chem. Rev. 2022, 461, 214496. [Google Scholar] [CrossRef]
  19. Shi, X.M.; Wang, Z.; Chen, M.H.; Wu, Q.Q.; Chen, F.Z.; Fan, G.C.; Zhao, W.W. Highly light−harvesting MOF−on−MOF heterostructure: Cascading functionality to flexible photogating of organic photoelectrochemical transistor and bienzyme cascade detection. Anal. Chem. 2024, 96, 3679–3685. [Google Scholar] [CrossRef]
  20. Chen, D.D.; Sun, Q.H.; Han, C.; Guo, Y.Y.; Huang, Q.; Goddard, W.A.; Qian, J.J. Enhanced oxygen evolution catalyzed by in situ formed Fe−doped Ni oxyhydroxides in carbon nanotubes. J. Mater. Chem. A 2022, 10, 16007–16015. [Google Scholar] [CrossRef]
  21. Guo, Y.Y.; Huang, Q.; Ding, J.Y.; Zhong, L.; Li, T.T.; Pan, J.Q.; Hu, Y.; Qian, J.J.; Huang, S.M. CoMo carbide/nitride from bimetallic MOF precursors for enhanced OER performance. Int. J. Hydrogen Energy 2021, 46, 22268–22276. [Google Scholar] [CrossRef]
  22. Zhou, N.; Su, F.F.; Guo, C.P.; He, L.H.; Jia, Z.K.; Wang, M.H.; Jia, Q.J.; Zhang, Z.H.; Lu, S.Y. Two−dimensional oriented growth of Zn−MOF−on−Zr−MOF architecture: A highly sensitive and selective platform for detecting cancer markers. Biosens. Bioelectron. 2019, 123, 51–58. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, D.X.; Liu, J.H.; Liu, H.B.; Yi, J.L.; Xia, F.Q.; Tian, D.; Zhou, C.L. Multiplexed electrochemical aptasensor based on mixed valence Ce (III, IV)−MOF for simultaneous determination of malathion and chlorpyrifos. Anal. Chim. Acta 2022, 1230, 340364. [Google Scholar] [CrossRef] [PubMed]
  24. Hu, M.L.; Abbasi-Azad, M.; Habibi, B.; Rouhani, F.; Moghanni-Bavil-Olyaei, H.; Liu, K.G.; Morsali, A. Electrochemical applications of ferrocene-based coordination polymers. ChemPlusChem 2020, 85, 2397–2418. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, Z.K.; Yu, H.J.; Wang, L.; Liu, X.W.; Lin, T.F.; Haq, F.; Vatsadze, S.Z.; Lemenovskiy, D.A. Ferrocene−contained metal organic frameworks: From synthesis to applications. Coord. Chem. Rev. 2021, 430, 213737. [Google Scholar] [CrossRef]
  26. Zeng, Q.H.; Chen, P.P.; Li, Z.F.; Wen, X.; Wen, W.; Liu, Y.; Zhao, H.L.; Zhang, S.P.; Zhou, H.H.; Zhang, L.Y. Application of a modified porphyrin in a polymer electrolyte with superior properties for all−solid−state lithium batteries. ACS Appl. Mater. Interfaces 2021, 13, 48569–48581. [Google Scholar] [CrossRef] [PubMed]
  27. Cui, J.; Wu, B.W.; Li, Z.Z.; Bai, Y.H.; Kan, L.; Wang, M.H.; He, L.H.; Du, M. Hierarchical CoCoPBA@ PCN−221 nanostructure for the highly sensitive detection of deoxynivalenol in foodstuffs. Food Chem. 2023, 403, 134370. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, Y.; Yuan, J.X.; Huang, S.H.; Chen, Z.Y.; Lu, C.B.; Yang, C.Q.; Zhai, G.Q.; Zhu, J.H.; Zhuang, X.D. Porphyrinic conjugated microporous polymer anode for Li−ion batteries. J. Power Sources 2022, 531, 231340. [Google Scholar] [CrossRef]
  29. Yu, K.; Won, D.I.; Lee, W.I.; Ahn, W.S. Porphyrinic zirconium metal−organic frameworks: Synthesis and applications for adsorption/catalysis. Korean J. Chem. Eng. 2021, 38, 653–673. [Google Scholar] [CrossRef]
  30. Jiang, R.; Pang, Y.H.; Yang, Q.Y.; Wan, C.Q.; Shen, X.F. Copper porphyrin metal−organic framework modified carbon paper for electrochemical sensing of glyphosate. Sens. Actuators B 2022, 358, 131492. [Google Scholar] [CrossRef]
  31. Dong, J.B.; Zheng, J.L.; Hou, J.Z.; Zhao, P.; Liang, Y.; Lei, J.C.; Luo, X.G.; Hou, C.J.; Huo, D.Q. Au nanoparticle/CoFc/metal−organic framework as enzyme−free dual−signal ratiometric electrochemical sensor for in−situ determination of cell−secreted H2O2. ACS Appl. Nano Mater. 2023, 6, 11630–11639. [Google Scholar] [CrossRef]
  32. Ye, G.J.; Luo, P.; Zhao, Y.S.; Qiu, G.L.; Hu, Y.; Preis, S.; Wei, C.H. Three−dimensional Co/Ni bimetallic organic frameworks for high−efficient catalytic ozonation of atrazine: Mechanism, effect parameters, and degradation pathways analysis. Chemosphere 2020, 253, 126767. [Google Scholar] [CrossRef] [PubMed]
  33. Hang, X.X.; Zhao, J.W.; Xue, Y.D.; Yang, R.; Pang, H. Synergistic effect of Co/Ni bimetallic metal–organic nanostructures for enhanced electrochemical energy storage. J. Colloid Interface Sci. 2022, 628, 389–396. [Google Scholar] [CrossRef] [PubMed]
  34. Tian, D.; Song, N.; Zhong, M.X.; Lu, X.F.; Wang, C. Bimetallic MOF nanosheets decorated on electrospun nanofibers for high−performance asymmetric supercapacitors. ACS Appl. Mater. Interfaces 2019, 12, 1280–1291. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, W.Q.; Wei, S.L.; Bai, P.Y.; Yang, C.C.; Xu, L. Robust coal matrix intensifies electron/substrate interaction of nickel−nitrogen (Ni−N) active sites for efficient CO2 electroreduction at industrial current density. Appl. Catal. B 2021, 299, 120661. [Google Scholar] [CrossRef]
  36. Liang, B.L.; Zhang, X.L.; Zhong, M.; Lv, C.C.; Li, K.X. Transition metal (Fe, Co, Ni) and sulfur codoped nitrogen−enriched hydrothermal carbon as high−performance cathode catalyst for microbial fuel cell. J. Power Sources 2021, 506, 230178. [Google Scholar] [CrossRef]
  37. Pang, L.Y.; Jia, X.; Wang, P.; Wang, Y.L.; Yang, Y.H.; Liu, H. Bimetallic synergy boost TCPP(Ni)−Co MOF as the high−performance electrochemical sensor for enhanced detection of trace theophylline. Microchem. J. 2022, 183, 107981. [Google Scholar] [CrossRef]
  38. Chen, G.X.; Dai, W.J.; Hu, C.Y.; Zang, H.; Sun, S.Y.; Zhen, S.J.; Zhan, L.; Huang, C.Z.; Li, Y.F. Ratiometric electrochemiluminescence of zirconium metal–organic framework as a single luminophore for sensitive detection of HPV−16 DNA. Anal. Chem. 2024, 96, 538–546. [Google Scholar] [CrossRef]
  39. Miao, J.C.; Du, K.; Li, X.; Xu, X.T.; Dong, X.; Fang, J.L.; Cao, W.; Wei, Q. Ratiometric electrochemical immunosensor for the detection of procalcitonin based on the ratios of SiO2−Fc–COOH–Au and UiO−66−TB complexes. Biosens. Bioelectron. 2021, 171, 112713. [Google Scholar] [CrossRef]
  40. Liu, D.Y.; Dai, Y.M.; Zou, J.Q.; Wang, S.Y.; Qi, L.; Ling, C. Sandwich−type ferrocene−functionalized magnetic nanoparticles: Synthesis, characterization, and the adsorption of Cr(VI). J. Mater. Sci. Mater. Electron. 2019, 30, 13924–13932. [Google Scholar]
  41. Gan, X.F.; Han, D.B.; Wang, J.M.; Liu, P.; Li, X.R.; Zheng, Q.Y.; Yan, Y.R. A highly sensitive electrochemiluminescence immunosensor for h−FABP determination based on self−enhanced luminophore coupled with ultrathin 2D nickel metal−organic framework nanosheets. Biosens. Bioelectron. 2021, 171, 112735. [Google Scholar] [CrossRef] [PubMed]
  42. Hallaj, R.; Haghighi, N. Photoelectrochemical amperometric sensing of cyanide using a glassy carbon electrode modified with graphene oxide and titanium dioxide nanoparticles. Microchim. Acta 2017, 184, 3581–3590. [Google Scholar] [CrossRef]
  43. Kang, J.; Huo, F.J.; Zhang, Y.B.; Chao, J.B.; Glass, T.E.; Yin, C.X. A novel near−infrared ratiometric fluorescent probe for cyanide and its bioimaging applications. Spectrochim. Acta Part A 2019, 209, 95–99. [Google Scholar] [CrossRef] [PubMed]
  44. Achadu, O.J.; Nyokong, T. Fluorescence “turn−ON” nanosensor for cyanide ion using supramolecular hybrid of graphene quantum dots and cobalt pyrene−derivatized phthalocyanine. Dyes Pigm. 2019, 160, 328–335. [Google Scholar] [CrossRef]
  45. Feng, Y.; Deng, D.Y.; Zhang, L.C.; Liu, R.; Lv, Y. LRET−based functional persistent luminescence nanoprobe for imaging and detection of cyanide ion. Sens. Actuators B 2019, 279, 189–196. [Google Scholar] [CrossRef]
Figure 1. (A) Preparation procedure of Ni/Co−MOF−TCPP−Fc. (B) Schematic representation of electrochemical sensing of CN by using Ni/Co−MOF−TCPP−Fc/SPCE.
Figure 1. (A) Preparation procedure of Ni/Co−MOF−TCPP−Fc. (B) Schematic representation of electrochemical sensing of CN by using Ni/Co−MOF−TCPP−Fc/SPCE.
Biosensors 14 00276 g001
Figure 2. (A) SEM of Ni/Co−MOF−TCPP−Fc and elemental mappings of Co, Fe, and Ni. (B) XPS spectrum of Ni/Co−MOF−TCPP−Fc. The high−resolution XPS spectra of (C) O 1s, (D) Ni 2p, (E) Co 2p, and (F) Fe 2p.
Figure 2. (A) SEM of Ni/Co−MOF−TCPP−Fc and elemental mappings of Co, Fe, and Ni. (B) XPS spectrum of Ni/Co−MOF−TCPP−Fc. The high−resolution XPS spectra of (C) O 1s, (D) Ni 2p, (E) Co 2p, and (F) Fe 2p.
Biosensors 14 00276 g002
Figure 3. (A) XRD patterns and (B) FT−IR spectra of Ni/Co−MOF−TCPP−Fc, TCPP, and Fc−COOH. (C) Nitrogen adsorption–desorption isotherms and (D) the corresponding pore size distribution of Ni/Co−MOF−TCPP−Fc.
Figure 3. (A) XRD patterns and (B) FT−IR spectra of Ni/Co−MOF−TCPP−Fc, TCPP, and Fc−COOH. (C) Nitrogen adsorption–desorption isotherms and (D) the corresponding pore size distribution of Ni/Co−MOF−TCPP−Fc.
Biosensors 14 00276 g003
Figure 4. (A) CV curves (at 100 mV/s) of Co−MOF/SPCE, Ni−MOF/SPCE, and Ni/Co−MOF/SPCE in 5 mM K3[Fe(CN)6] and 0.1 M KCl. (B) Linear relationship between CV peaks and square roots of scan rates (10~100 mV/s, step: 10 mV/s) for Ag−Ni/Co−MOF−TCPP−Fc/SPCE in 5 mM K3[Fe(CN)6] and 0.1 M KCl. (C) Logarithmic fitting relationship curve for peak potential and scan rate. (D) DPV signals of Ni/Co−MOF−Fc/SPCE, Ni/Co−MOF−TCPP/SPCE, and Ag−Ni/Co−MOF−TCPP−Fc/SPCE in PBS buffer. (E) Signal response of bare electrode and modified electrode in PBS buffer (containing 1.67 ng/mL CN). (F) Semiquantitative analyses in PBS buffer (containing 0, 1.67, and 16.7 ng/mL CN).
Figure 4. (A) CV curves (at 100 mV/s) of Co−MOF/SPCE, Ni−MOF/SPCE, and Ni/Co−MOF/SPCE in 5 mM K3[Fe(CN)6] and 0.1 M KCl. (B) Linear relationship between CV peaks and square roots of scan rates (10~100 mV/s, step: 10 mV/s) for Ag−Ni/Co−MOF−TCPP−Fc/SPCE in 5 mM K3[Fe(CN)6] and 0.1 M KCl. (C) Logarithmic fitting relationship curve for peak potential and scan rate. (D) DPV signals of Ni/Co−MOF−Fc/SPCE, Ni/Co−MOF−TCPP/SPCE, and Ag−Ni/Co−MOF−TCPP−Fc/SPCE in PBS buffer. (E) Signal response of bare electrode and modified electrode in PBS buffer (containing 1.67 ng/mL CN). (F) Semiquantitative analyses in PBS buffer (containing 0, 1.67, and 16.7 ng/mL CN).
Biosensors 14 00276 g004
Figure 5. (A) pH optimization of PBS buffers (containing 16.7 ng/mL CN). (B) DPV curves and (C) Linear calibration plots for laboratorial samples containing gradual−level of CN in 0.1 M PBS. Enlarged images in the 0~0.01 µg/mL is presented in the inset of (C).
Figure 5. (A) pH optimization of PBS buffers (containing 16.7 ng/mL CN). (B) DPV curves and (C) Linear calibration plots for laboratorial samples containing gradual−level of CN in 0.1 M PBS. Enlarged images in the 0~0.01 µg/mL is presented in the inset of (C).
Biosensors 14 00276 g005
Figure 6. (A) DPV response at the sensor toward 0.1 ng/mL CN with the presence of different interference at 5 ng/mL. (B) Repeatability for the sensor. Recommended ratiometric readouts show better consistency than raw signals of ICN in experiments of (B) stability. All DPV readouts are conducted with 3 replicates.
Figure 6. (A) DPV response at the sensor toward 0.1 ng/mL CN with the presence of different interference at 5 ng/mL. (B) Repeatability for the sensor. Recommended ratiometric readouts show better consistency than raw signals of ICN in experiments of (B) stability. All DPV readouts are conducted with 3 replicates.
Biosensors 14 00276 g006
Table 1. Summary of this method and previous related studies.
Table 1. Summary of this method and previous related studies.
MethodResponse TimeSensitivityDetermination LimitLinear RangeReference
Ion chromatography0.09071 μg/L10~160 μg/L[5]
HPLC−MS0.00530.07 μg/L0.2~8.42 μg/L[7]
Photoelectrochemical amperometry0.05180.1 μM0.1~60 μM[42]
Fluorescence4.8 min2.452.41 μM0~30 μM[43]
Fluorescence0.5 nM1~50 nM[44]
Fluorescence0.11 μM0.3~40 μM[45]
Electrochemical 4 min63.106120.052 ng/mL0.196~44 ng/mLThis method
Table 2. Analyses of CN−contaminated samples (mean ± SD).
Table 2. Analyses of CN−contaminated samples (mean ± SD).
SpeciesMethod Comparison (µg/mL)
This MethodGC−MSRSD
Fermented grainsSample 10.021 ± 0.005 a0.020 ± 0.007 a5.6%
Sample 20.020 ± 0.003 a0.021 ± 0.005 a5.9%
Sample 30.018 ± 0.005 a0.017 ± 0.006 a4.7%
BaijiuSample 10.024 ± 0.007 a0.025 ± 0.006 a3.4%
Sample 20.029 ± 0.003 a0.030 ± 0.004 a3.8%
Sample 30.021 ± 0.004 a0.020 ± 0.002 a4.2%
River waterSample 10.019 ± 0.003 a0.020 ± 0.002 a5.3%
Sample 20.020 ± 0.002 a0.021 ± 0.002 a3.3%
Sample 30.021 ± 0.003 a0.020 ± 0.004 a2.5%
Mineral waterSample 1Not detectedNot detected
Sample 2Not detectedNot detected
Sample 3Not detectedNot detected
Data are expressed as means ± SD of duplicate assays, Values followed different superscripts in the same column present significantly different (p < 0.05).
Table 3. Determination of CN using the proposed sensor.
Table 3. Determination of CN using the proposed sensor.
SamplesDetected (µg/mL)Added (µg/mL)FoundRecovery
Fermented grains0.021 ± 0.0050.10.119 ± 0.00598.9%
Baijiu0.024 ± 0.0070.128 ± 0.009103.3%
River water0.019 ± 0.0060.121 ± 0.004101.1%
Mineral waterNot detected0.098 ± 0.00398.5%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, Z.; Wu, Z.; Zong, Y.; Li, C.; Guo, W.; Guo, Y.; Zou, X. Construction of Metal–Organic Framework as a Novel Platform for Ratiometric Determination of Cyanide. Biosensors 2024, 14, 276. https://doi.org/10.3390/bios14060276

AMA Style

Sun Z, Wu Z, Zong Y, Li C, Guo W, Guo Y, Zou X. Construction of Metal–Organic Framework as a Novel Platform for Ratiometric Determination of Cyanide. Biosensors. 2024; 14(6):276. https://doi.org/10.3390/bios14060276

Chicago/Turabian Style

Sun, Zongbao, Zhiwei Wu, Yiran Zong, Chen Li, Wang Guo, Yiqing Guo, and Xiaobo Zou. 2024. "Construction of Metal–Organic Framework as a Novel Platform for Ratiometric Determination of Cyanide" Biosensors 14, no. 6: 276. https://doi.org/10.3390/bios14060276

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop