Next Article in Journal
Reversible Colorimetric and Fluorescence Solid Sensors Based on Aryl Hydrazone Derivatives of 1,8-Naphthalimides for Caustic Media and Biogenic Amine Vapors
Next Article in Special Issue
Automated Multistep Lateral Flow Immunoassay Using a Smartphone for the Quantification of Foodborne Bacteria from Fresh Lettuce
Previous Article in Journal
The Advances of Hydrosol–Gel Transition-Based Sensors
Previous Article in Special Issue
A High-Precision Three-Dimensional Probe Array Temperature Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Non–Enzymatic Electrochemical Sensor of Cu@Co–MOF Composite for Glucose Detection with High Sensitivity and Selectivity

1
Key Laboratory of Macromolecular Science of Shaanxi Province, Shaanxi Key Laboratory for Advanced Energy Devices, Shaanxi Engineering Laboratory for Advanced Energy Technology, School of Chemistry & Chemical Engineering, Shaanxi Normal University, Xi’an 710062, China
2
Shaanxi Key Laboratory of Chemical Additives for Industry, College of Chemistry and Chemical Engineering, Shaanxi University of Science and Technology, Xi’an 710021, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Chemosensors 2022, 10(10), 416; https://doi.org/10.3390/chemosensors10100416
Submission received: 10 August 2022 / Revised: 7 October 2022 / Accepted: 9 October 2022 / Published: 12 October 2022

Abstract

:
A 3D cobalt metal–organic framework (Co–MOF), [Co3(BDC)3(DMU)2], was utilized to prepare Cu@Co–MOF composite in a deposition–reduction process. Cu@Co–MOF/GCE (GCE = glassy carbon electrode) electrode was prepared by “drop–coating” method. Cu@Co–MOF/GCE shows excellent electrocatalytic activity for Glu detection. The chronoamperometric response of Cu@Co–MOF/GCE to Glu concentration (CGlu) displays linear relationships in two CGlu sections with calculated sensitivities of 282.89 μA mM−1 cm−2 within 0.005–0.4 mM Glu and 113.15 μA mM−1 cm−2 within 0.4–1.8 mM Glu. The detection limit is calculated as 1.6 μM at S/N = 3. Cu@Co–MOF/GCE also exhibits a rapid current response, high anti–interference, stability, and repeatability to Glu detection. Cu@Co–MOF/GCE was applied to detect Glu in human serum and orange juice. All found CGlu are very close to those added CGlu with low RSDs and high recoveries. Cu@Co–MOF/GCE as a non–enzymatic electrochemical sensor of Glu has high sensitivity, selectivity, accuracy, and reliability.

1. Introduction

As the most important monosaccharide, glucose (Glu) plays a crucial role in bodily functions. Glu provides the main energy source of the body, which hydrolyzes in the body and is stored as glycogen. Glu can protect and detoxify the liver, and also promotes the metabolism of poisons. Glu is the most important and basic substance in medicine. It can be adsorbed fast by human tissues [1]. However, the fluctuation of blood Glu hazards the health. For example, hypoglycemia leads to brain dysfunction, induces cardiovascular diseases of arrhythmia, myocardial infarction and stroke, and causes nervous diseases; hyperglycemia will result in dehydration, metabolic dysfunction, water electrolyte disorder, fatigue, decreased resistance, and especially, diabetes. Diabetes is a major predisposition to other serious illnesses of kidney failure, stroke, and blindness, causing a huge financial burden [2]. Thus, an accurate, low–cost, and easy measurement method for sensing Glu in the human body is greatly needed [3,4]. The measurement methods of Glu mainly include chromatography, spectroscopy, and electrochemistry [5,6]. The electrochemical Glu sensors have the advantages of high reliability, low–cost, and outstanding operability, which attract a lot of interest [7,8]. Electrochemical Glu sensors are divided into enzymatic and non–enzymatic sensors. The enzymatic Glu sensors occupy the main part in the market of Glu sensor industry, but the enzymatic Glu sensors show enzyme activity–dependent sensitivities [9]. Their detection performances are seriously influenced by some environmental factors such as room temperature, humidity, pH value, etc. [10]. Some limitations are related with enzymes, such as enzyme payload, enzyme activity, poor reproducibility, and poor stability during process [11]. Overall, enzymes have become a difficulty to some degree in the development and application of enzymatic Glu sensors.
Compared with enzymatic Glu sensors, non–enzymatic sensors are easy to realize, long–term, and stable Glu detection, whose detection performances are not influenced by environmental factors, providing a support for high–performance continuous monitoring of Glu [12]. Non–enzymatic Glu sensors have much more choices in material platforms to achieve larger surface area rather than the stacking of enzymes. For example, metal–organic frameworks (MOFs) are a kind of porous materials constructed by the self–assembly of metal ions with organic ligands into highly ordered structures [13,14]. Their poor conductivity blocks raw MOFs as electrocatalysts for sensing Glu. Metal nanoparticles (NPs), such as nano gold and silver, have high specific surface areas, catalytic activity, and electrical conductivity [15]. Doping metal NPs into the electrochemical Glu sensors can greatly improve the conductivities and the detection performances of the MOF–based electrochemical Glu sensors [16,17]. Furthermore, MOFs and metal NPs integrate into modified electrodes as Glu sensors with the effective reaction areas and the interface electron migration rates. It shows the advantages of high catalysis and high sensitivity for the oxidation of Glu. Without limitations in biological enzymes, the working electrodes of NPs@MOFs composites are of great significance as non–enzymatic Glu sensors with high sensitivities and anti–interference [18,19,20]. Due to the high cost of nano novel metal–based electrochemical sensors, the application of nano transition metals in electrochemical sensors is expected to reduce the cost but not sacrifice the advantages of doping nano novel metal. The preparation and large–scale application of low–cost and high–performance transition metal–based electrochemical devices are a support for Glu detection [21,22,23]. There are a lot of explorations on transition metal NPs@MOF as electrochemical Glu sensors. For example, as a nonenzymatic electrochemical analytical chip, a cobalt MOF/carbon cloth/paper button–sensor quantitatively determines Glu. The sensor was applied in complex biosystems of human serum, urine, and saliva, showing high selectivity and durability [24].
Based on the aforementioned discussions, we deposit Cu NPs in a cobalt MOF (Co–MOF), aiming at the integrated advantages: porous structure in Co–MOF with active sites can host Cu NPs in a restricted area, and thus control Cu NPs growing in a small size and prevent Cu NPs from migration and agglomeration in preparation and electrochemical measurements. Co–MOF integrating with Cu NPs can effectively improve the electrocatalytic activity. Herein, our work presents a Co–MOF, [Co3(BDC)3(DMU)2], as a non–enzymatic Glu sensor, which was prepared under ionothermal conditions. The Co–MOF was characterized in detail by single crystal X–ray diffraction (SCXRD), powder X–ray diffraction (XRD), Fourier transform infrared spectroscopy (FT–IR), thermogravimetric analysis (TGA), energy dispersive X–ray spectrometry (EDS), and X–ray photoelectron spectroscopy (XPS). Cu@Co–MOF and its composite electrode Cu@Co–MOF/GCE (GCE = glassy carbon electrode) were prepared through a sequential deposition–reduction process. Cu@Co–MOF/GCE electrode exhibits a high electrocatalytic performance to Glu detection, which returns accurate Glu concentrations in human serum and orange juice. Cu@Co–MOF/GCE can be regarded as a non–enzymatic electrochemical Glu sensor with high sensitivity, anti–interference, reproducibility, stability, and rapid current response.

2. Materials and Methods

2.1. Materials and Instruments

All chemicals are in analytical grade and obtained from commercial sources without purification. Co(NO3)2·6H2O, CuSO4, terephthalic acid (H2BDC), ethanol, methanol, N,N–dimethylformamide (DMF), NaBH4, Al2O3 (0.5 μm and 50 nm) powder, HNO3, H2SO4, KCl, K3Fe(CN)6 NaOH, Glu were bought from Sinopharm Chemical (Shanghai, China). Interferents of D–Mannose (D–Man), D–fructose (D–Fru), aspartic acid (AA), dopamine (DA) urea, and uric acid (UA) were obtained from Aladdin (Shanghai, China). Choline chloride (ChCl) and 1,3–dimethylurea (DMU) were purchased from Tokyo Chemical (Tokyo, Japan). Nafion solution of 5 wt% was bought from Sigma–Aldrich (Shanghai, China). KBr pellets of FT–IR grade was bought from Aldrich. Minutemaid orange juice was bought from a local supermarket.
XRD were measured on a Rigaku MiniFlex 600 (Japan) diffractometer under 40 kV and 15 mA at room temperature. FT–IR spectra were collected on a Bruker Tensor 27 FT–IR spectrometer in the wave number range of 4000–400 cm−1. TGA was performed on a SDT Q600 V8.3 Build 101 instrument with a heating rate of 10 °C·min−1 and in a N2 atmosphere with a flow rate of 20 cm3·min−1. The surface elemental contents were determined by EDS on a Philips–FEI Quanta 200 scanning electron microscope. XPS was used to investigate the chemical states of the surface elements in Co–MOF on an Axis Ultra (Kratos Analytical Ltd., Manchester, UK). A CHI 660E (CH Instruments, Inc. Shanghai, China) electrochemical workstation from CH Instruments, Inc. (Shanghai, China) inspected all electrochemical measurements at room temperature.

2.2. Preparation

Co–MOF. Co(NO3)2·6H2O (0.40 mmol, 0.1164 g) and H2BDC (0.20 mmol, 0.0332 g) were mixed with a deep eutectic solvent (DES)–type ionic liquid of ChCl (1 mmol, 0.1396 g) and DMU (2 mmol, 0.1762 g) as in a crystallization vial. The mixture was heated at 100 °C for 3 days, and naturally cooled to room temperature. The as–synthesized purple crystals of Co–MOF were washed with DMF and ethanol and collected for the further characterizations. FT–IR data (in KBr, cm−1): 3475 (w), 3411 (m), 3352 (m), 2927 (w), 2370 (w), 2073 (w), 1595 (s), 1388 (s), 1142 (w), 1013 (m), 886 (m), 822 (s), 748 (s), 543 (s).
Cu2+/Co–MOF. Three groups of ca. 0.1238 g Co–MOF were added into 4 mL CuSO4 methanol solution with concentrations of 5, 10, and 15 mmol∙L−1 (mM), respectively. After 6 h of continuous stirring, the suspensions were centrifuged and dried in a vacuum oven at 60 °C for 12 h to obtain Cu2+@Co–MOF.
Cu@Co–MOF. Three groups of 10, 20, and 30 mg NaBH4 were dissolved in 4 mL methanol, and then mixed with the above Cu2+/Co–MOF with a 30 min strong agitation. The mixture turned to black immediately, the suspensions were filtered, washed with methanol and dried at 60 °C to obtain Cu@Co–MOF.
Cu@Co–MOF/GCE electrode. Amounts of 0.5 μm and 50 nm alumina slurry on a polishing cloth were continuously used to polish a GCE with a 5 mm diameter to a mirror surface, then washed with HNO3 solution of Vwater:VHNO3 = 1:1, anhydrous ethanol and ultrapure water by sonication. GCE was further activated by 0.5 M H2SO4 and 0.1 M KCl/5 mM K3Fe(CN)6, respectively. A total of 5 mg ground Cu@Co–MOF powder was dispersed in 1 mL anhydrous ethanol with 20 min sonication to form a uniform Cu@Co–MOF suspension of 5 mg∙mL−1. Cu@Co–MOF/GCE electrode was obtained by a 10 μL Cu@Co–MOF suspension and a 5 μL nafion solution of 5 wt% was dropped on the GCE surface, then dried at room temperature.

2.3. Electrochemical Measurements

Three–electrode system was used in the electrochemical measurements: The as–prepared Cu@Co–MOF/GCE is as the working electrode, a platinum wire as the counter electrode and KCl saturated Ag/AgCl as the reference electrode. The electrochemical performance of Cu@Co–MOF/GCE for sensing Glu was estimated by cyclic voltammetry (CV) and amperometric It curve methods.

2.4. Crystal Structure Determination

A suitable single crystal was mounted and collected the crystal data on a Bruker D8 Quest CCD diffractometer with a Cu–Kα radiation with λ = 1.54184 Å at 120 K. The crystal data was reduced via the ω scan technique by Bruker XSCANS program, and the direct method was used to solve the crystal structure of Co–MOF with SHELXL–2014/7 crystallographic software package. All non–hydrogen atoms were refined anisotropically. The positions of all hydrogen atoms were calculated according to their carriers and then were refined isotropically in the final refinement stage.
Crystal data (excluding structure factors) of Co–MOF has been deposited in the Cambridge Crystallographic Data Centre (CCDC) with No. of 2195853. The data can be obtained for free from CCDC via www.ccdc.cam.ac.uk/data_request/cif (accessed on 8 October 2022).

3. Results

3.1. Structure Description

[Co3(BDC)3(DMU)2] (Co–MOF) crystalizes in the triclinic Pī space group, whose asymmetric unit contains one and a half Co(II) centers, three separate BDC2− ligands of 0.5 occupancy and a DMU molecule. BDC2− ligand adopts two coordination modes of μ4−bis–bidentate bridging and μ4–bis–monodentate bridging/chelating (Supplementary Information (SI), Figure S1). Six–coordinated Co1 locates in an octahedron constructed by six carboxylate O atoms from six separate BDC2− ligands. Five–coordinated Co2 is in a square pyramidal geometry, in which four carboxylate O atoms from two BDC2− ligands shape the basal plane, and an O atom from DMU is in the apex. Central Co2 connects two symmetric terminal Co1 by two pairs of μ2–bidentate COO groups and one pair of μ2–monodentate bridging/chelating COO groups, respectively, and one DMU ligand binds to each terminal Co1 into a linear trinuclear [Co3(COO)6(DMU)2] secondary building unit (SBU) (SI, Figure S2). Neighboring [Co3(COO)6(DMU)2] SBUs are connected by the three separate BDC2− ligands with O11/O12 and O31/O32 (μ4–bis–bidentate bridging), or O21/O22 (μ4–bis–monodentate bridging/chelating) into three 1D chains along the [110], [011] and b–directions, respectively (SI, Figure S3a–c). The chains along the [110] and b–directions weave a 2D layer, which topologizes a [4,4] network. The 2D layers further stack into a 3D framework with the fabrication of the chains along the [011] direction (Figure 1). With the rest phenyl rings of the three separate BDC2− ligands dummied as the 2–connected sticks and the [Co3(COO)6(DMU)2] SBUs as the six–connected nodes, the point symbols of the rest phenyl rings are {8}, and the one of [Co3(COO)6(DMU)2] is {812.123}. Therefore, the 3D framework is topologized as a 2,6–connected {812.123}{8}3 network.

3.2. Characterizations

XRD. The experimental and simulated XRD patterns of Co–MOF, and those of Co–MOF and Cu@Co–MOF were compared (Figure 2). The experimental XRD pattern of Co–MOF is in good agreement with the one simulated from the single crystal data, indicating that the bulk sample of Co–MOF is in high crystallinity and purity for the following characterizations. In the synthetical process of Cu@Co–MOF, metal precursors contained in organic solvents were used, a part of which will deposit on the surface of Co–MOF after drying, and then tend to aggregate into nanosheets, while high concentration NaBH4 solution reduces Cu2+, thus avoiding Cu NPs unevenly aggregating on the surface of Co–MOF [25]. The XRD diffraction peaks of Co–MOF can be easily identified from those of Cu@Co–MOFs, indicating that Co–MOF stays the structural stability after Cu NPs are loaded. Comparing all Cu@Co–MOFs with Cu2+ ranging 5, 10, and 15 mM, the same XRD patterns demonstrate the isomorphism and the independence of Co–MOF on Cu2+ concentrations.
FT–IR. The characteristic absorption peaks of Co–MOF, Cu@Co–MOF and free H2BDC ligand are analyzed by FT–IR (SI, Figure S4). The peaks within 3410–3310 cm−1 and 1650–1550 cm−1 are related with the symmetric and asymmetric stretching vibration peaks of –NH from the secondary amine DMU in Co–MOF and Cu@Co–MOF, which do not appear in H2BDC. Those in 3100–3000 cm−1 are attributed to very weak vC–H from phenyl rings in Co–MOF, Cu@Co–MOF and H2BDC. Compared to 1682 cm−1 of carbonyl in free H2BDC, the peaks, respectively move to 1720–1550 cm−1 and 1450–1250 cm−1, attributed to the asymmetric and symmetric stretching vibrations (vas(–COO) and vs(–COO)) of the carboxyl groups in Co–MOF and Cu@Co–MOF. The wavelength differences are larger than 200 cm−1, supporting the carboxyl groups are in coordination [26,27], agreeing with the structural analysis. Comparing Cu@Co–MOF with Co–MOF, the peaks at 3348 and 3388 cm−1 become weak, concerning with the interactions between Cu NPs and –NH groups after Cu2+ deposing–reducing on Cu–MOF; similarly, the peaks at 1552 and 1626 cm−1 are weakened, but the one at 1594 cm−1 concerning with the in–plane bending vibration of –NH group is strengthened. All changes support that Cu2+ loads on Co–MOF through the interaction between Cu2+ and –NH groups during the composite process [28,29].
TGA. The TGA curve shows there are three steps in the thermal decomposition of Co–MOF (Figure S5). There is a weight loss of only ca. 1.85% before 340 °C, coming from the loss of absorbed water or organic solvents. The second weigh loss of 20.68% is attributed to the removal of DMU as the temperature rises to 400 °C, very close to the calculated value of 20.85%. Followed with the collapse of the MOF skeleton, the third decomposition ends at 570 °C with 44.75% weight loss, relating with the decomposition of BDC2− ligands.
EDS. Surface elemental contents of Co–MOF and Cu@Co–MOF were analyzed by EDS. C, N, O and Co elements can be observed on the surface of Co–MOF with the contents of 65.10, 7.47, 21.57 and 5.85 at%, respectively (Figure S6a), basically in accord with the formula C30H28Co3N4O14. Besides C, N, O, and Co elements with closed contents (64.10, 6.13, 25.23, and 4.40 at% for C, N, O, and Co, respectively) in Cu@Co–MOF, there is Cu element of 0.22 at%, demonstrating the load of Cu NPs on the surface of Co–MOF (Figure S6b).
XPS. Surface electronic states and the compositions of Co–MOF and Cu@Co–MOF were inspected by XPS. XPS of Co 2p in Co–MOF, Cu2+/Co–MOF and Cu@Co–MOF are listed in Figure 3, which mainly includes Co 2p3/2 and Co 2p1/2 characteristic regions. The Co 2p3/2 region in Co–MOF contains the main peak at 781.29 eV and the accompanying one at 786.22 eV (Figure 3a). Similarly, the main peak at 797.15 eV and its accompanying peak at 802.70 eV can be found in the Co 2p1/2 region in Co–MOF; in Cu2+/Co–MOF, the main and accompanying peaks in Co 2p3/2 region are at 781.48 and 786.13 eV, respectively, those in Co 2p1/2 region are at 797.32 and 802.60 eV (Figure 3b); the corresponding main and accompanying peaks of Cu@Co–MOF are at 781.59 and 786.20 eV in Co 2p3/2 region, and at 797.39 and 802.69 eV in Co 2p1/2 region, respectively (Figure 3c). By comparing the above peaks, Co–MOF, Cu2+/Co–MOF and Cu@Co–MOF show the same main and the accompanying peaks both in Co 2p3/2 and Co 2p1/2 characteristic regions, indicating the incorporation with Cu2+ or Cu NPs does not change the valence state of Co(II) cation.
XPS of Cu 2p3/2 and Cu 2p1/2 characteristic regions in Cu2+/Co–MOF and Cu@Co–MOF are also compared (Figure 4). In Cu2+/Co–MOF, the peaks at 933.26 eV in Cu 2p3/2 and at 953.78 eV in Cu 2p1/2 correspond to Cu(II) (Figure 4a). Differently, there exists four peaks in Cu 2p3/2 and Cu 2p1/2 regions of Cu@Co–MOF: those at 932.80 eV in Cu 2p3/2 and 952.55 eV in Cu 2p1/2 are attributed to Cu(0); those are at 935.27 eV in Cu 2p3/2 and 954.82 eV in Cu 2p1/2 related with Cu(II) (Figure 4b) [30,31,32]. It indicates that Cu2+ deposits in the Co–MOF to form Cu2+/Co–MOF, but only a part of deposited Cu2+ were reduced to Cu(0). Cu2+ and Cu(0) co–exist in Cu@Co–MOF.

3.3. Electrochemically Sensing Glu by Non–Enzymatic Cu@Co–MOF

CV test. Cu@Co–MOF was modified on the surface of GCE to prepare Cu@Co–MOF/GCE as working electrode. CV tests were carried out in 0.01 M NaOH at working potential of +0.6 V (Figure 5). The redox peaks of CoII/CoIII can be obviously observed in all the CV curves of Cu@Co–MOF with feeding Cu2+ concentrations (CCu2+) within 5–15 mM. With CCu2+ increasing, the redox peaks weaken, indicating CCu2+ of 5 mM shows the most sensitive current response (Figure 5a); 5 mM Cu2+ undertakes the synergistic role with Co–MOF. Higher Cu2+ concentration will occupy more active sites and prevent the electron transfer. After the reduction of Cu2+, high Cu2+ concentration easily leads to the aggregation of Cu NPs, decreasing electrocatalytic performance. It supports the decreased trend of electrocatalysis with Cu2+ concentration. However, lower Cu2+ concentration has a difficulty in generating appropriate Cu NPs to form sufficient electron transfer path. Therefore, 5 mM is selected as the feeding Cu2+ concentration for the subsequent tests. By comparing Cu@Co–MOF/GCE and Co–MOF/GCE in the potential vs. current response, Cu@Co–MOF/GCE shows a much stronger redox peak, indicating the conductivity of Co–MOF improved by Cu NPs. After 1 mM Glu is added, the redox peak of Co–MOF/GCE slightly weaken, suggesting an insensitivity of Co–MOF/GCE to Glu. Differently, Cu@Co–MOF/GCE causes the anode and cathode currents increased greatly at the initial potential of +0.4 V as 1 mM Glu is added (Figure 5b). It suggests that Cu@Co–MOF/GCE has a better catalytic oxidation effect on Glu with the improved conductivity.
The detection mechanism of Cu@Co–MOF sensing Glu is illustrated in Scheme 1, which concerns a Glu oxidizing process explained by Reactions 1 and 2. Blank Co–MOF/GCE compared with blank Cu@Co–MOF/GCE, an obvious improved electroactivity is found, showing Cu NPs effectively increase conductivity of Co–MOF/GCE. When 1 mM Glu is added into 0.01 mM NaOH, Co–MOF/GCE exhibits a negligible electroactivity. This concerns the poor electroconductivity of Co–MOF, resulting in Co–MOF/GCE’s weak electrocatalysis to Glu, while as Glu contacts with Cu@Co–MOF/GCE, a significant enhancement occurs. The above comparisons demonstrate the improvement of Cu NPs on the electrocatalytical performance of Co–MOF/GCE. The sensing mechanism is related with a synergistic effect between Co–MOF and Cu NPs. Cu is oxidized to CuO and CuOOH in an alkaline medium, and therefore, Cu reaches the high oxidation state of Cu [33]. An intermediate enediol from the deprotonation of Glu in alkaline medium can be oxidized into gluconolactone when contacting with CuOOH [25,34]. The improved electrocatalysis of Cu@Co–MOF/GCE mainly benefits from the porous structure of Co–MOF. Porous structure in Co–MOF can host Cu NPs in a restricted area, and thus control Cu NPs growing in a small size and prevent Cu NPs from agglomeration, migration and electrochemical dissolution in preparation and electrocatalytic processes. The full use of pores in Co–MOF provides large active surface area and active sites to encapsulate Cu NPs on the –NH groups. Co–MOF with encapsulated Cu NPs can effectively improve the electrocatalytic activity. Furthermore, the comparison of Cu@Co–MOF before and after the electrochemical detection indicates Cu@Co–MOF can keep enough stability to preserve the crystallinity from the suffering of NaOH and current (Figure S7).
CoII–MOF + OH − e → CoIII–MOF
CoIII–MOF + glucose → CoII–MOF + gluconolactone
Amperometric It curve test. The chronoamperometric response of Cu@Co–MOF/GCE to Glu concentrations (CGlu) ranging from 0.005 to 1.8 mM was recorded in 0.01 M NaOH at working potential of +0.6 V [35]. With Glu added, the current response gradually increases with increased CGlu (Figure 6a). By depicting the plot of detected current vs. CGlu, the linear relationships can be figured out in the CGlu sections of 0.005–0.4 mM and 0.4–1.8 mM (Figure 6b). The linear equations are as follows with high correlation coefficient R2:
CGlu = 0.005–0.4 mM:
j Cu @ Co MOF / GCE   mA   cm 2 = 0.020 C Glu μ M + 8.266   R 2 = 0.9802
CGlu = 0.4–1.8 mM:
j Cu @ Co MOF / GCE   mA   cm 2 = 0.012 C Glu μ M + 6.096   R 2 = 0.9800
EDS analysis shows only 0.22 at% Cu deposits on Co–MOF, meaning ca. one Cu corresponding to 30 CoII. It hints a saturation effect in Cu@Co–MOF/GCE sensing Glu. It comes from the fact that all active sites of Cu NPs loading on Co–MOF are occupied by Glu molecules. The sensitivities of Cu@Co–MOF/GCE are calculated as 282.89 μA mM−1 cm−2 within CGlu = 0.005–0.4 mM, and 113.15 μA mM−1 cm−2 within CGlu = 0.4–1.8 mM. The detection limit was dived to 1.6 μM at a signal–to–noise (S/N) ratio of 3 [36].
Cu@Co–MOF/GCE is compared with other reported non–enzymatic electrochemical Glu sensors in Table 1 [37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58]. Its detection limit of 1.6 μM is lower than a lot of non–enzymatic electrochemical Glu sensors, and comparable to Cu@HHNs (1.97 μM) [39], Au@Ni–BTC (1.5 μM) [46], CPO–27–NiII (1.46 μM) [47], and CoII–MOF/Acb (1.7 μM) [56], but higher than Ni/Co(HHTP)MOF/CC (0.1 μM) [37], NiCo–LDH/CC (0.12 μM) [40], Co–MOF/EG (0.58 μM) [44], Cu(OH)2@CoNi–LDH NT–NSs/GSPE (0.6 μM) [45], SiCNPs–ENFM (0.56 μM) [3], NiCo–MOF (0.29 μM) [49], UiO–67@Ni–MOF (0.98 μM) [50] and Ag@TiO2@ZIF–67 (0.99 μM) [55]. Similarly, the sensitivity of 282.89 μA mM−1 cm−2 within CGlu = 0.005–0.4 mM is more sensitive than most of the reported electrochemical Glu sensors, but lower than some sensors, such as Ni/Co(HHTP)MOF/CC (3250 μA mM−1 cm−2) [37], CoZn–BTC/GCE (1218 μA mM−1 cm−2) [22], Cu@HHNs (1594.2 μA mM−1 cm−2) [39] and Ni0.7Co0.3(OH)2 (1541 μA mM−1 cm−2) [42]. Overall, Cu@Co–MOF/GCE shows a low detection limit and a high sensitivity to Glu electrochemical detection.
Response time and anti–interference. Cu@Co–MOF/GCE exhibits a rapid current response of less than 7 s to CGlu change (Figure S8). The anti–interference performance was evaluated by amperometric method. In the human body, the normal blood Glu concentration locates in 3–8 mM and the one in diabetes higher than 11.1 mM; while compared to the concentrations of UA, AA, and DA are around 0.13–0.46 mM, 0.23 μM, and 0.1 mM, respectively, in physiological conditions. Therefore, we carried out the anti–interference test in the presence of 0.1 mM Glu and 0.01 mM of each interferent. The current responses were measured by continuous dropping method when 2 μM, 0.1 mM Glu was dropped into 0.01 M NaOH comparing with those of 0.01 mM interferents including D–Fru, D–Man, urea, AA, UA, and DA. As 0.1 mM Glu is added every injection, remarkable current signals always immediately appear (Figure 7). Because of the dilution effect, the current responses raised by Glu become weak, while when the interferents are added one by one into the system, the current responses of oxidizing interferents are observed to be very faint. If setting the percentage amperometric response of the first adding of Glu as 100%, those amperometric responses of the interferents are 3.17, 6.35, 1.59, 9.52, 4.76, and 11.11%. Therefore, the high anti–interference of Cu@Co–MOF is expectable in electrochemically sensing Glu.
Stability and repeatability. Stored at 4 °C for 3–15 days, Cu@Co–MOF/GCE as non–enzymatic electrochemical sensor detects 1.0 mM Glu in 0.01 M NaOH every three days by amperometric method, whose current changes were recorded in Figure 8a. The current of the fresh Cu@Co–MOF/GCE is set as 100%. In the first three days, the current remains 96.50%, then 85.78%, 77.06%, 72.39%, and 69.25% for the following three–day sections. The current downward trend becomes slow. Even after the 15 days, Cu@Co–MOF/GCE still remains nearly 70% current response to Glu, indicating Cu@Co–MOF/GCE stays enough sensitivity and stability to Glu within 15 days. Five parallel measurements of sensing Glu were carried out with five groups of fresh Cu@Co–MOF/GCE sensors to evaluate the repeatability of Cu@Co–MOF/GCE (Figure 8b). The relative standard deviation (RSD) of the five parallel measurements is 5.84%, indicating the high repeatability of Cu@Co–MOF/GCE.

3.4. Sensing Glu in Human Serum and Orange Juice

Standard addition method was used to evaluate the application performance of Cu@Co–MOF/GCE sensing Glu in human serum and orange juice. A total of 50 μL human serum was diluted by 50 mL, 0.01 M NaOH, then 1.0 mM Glu followed to prepare 50, 100, and 200 μM Glu standard solutions. The current responses of each CGlu were recorded for five times by CV method (Table 2). Through the linear equation of j Cu @ Co MOF / GCE   mA   cm 2 = 0.020 C Glu   μ M + 8.266 , the found CGlu were calculated as 50.01, 94.93, and 188.97 μM based on the found currents. The recoveries (defined as the found CGlu/added CGlu) were calculated as 100.02%, 94.93%, and 94.49%. The RSDs (n = 5) range from 2.27% to 4.30%, all less than 5%.
The supernatant of minutemaid organic juice was obtained by 8000–rpm centrifugation for 5 min; 50 mL, 0.01 M NaOH diluted 50 μL extracted supernatants. Similar sensing Glu processes were carried out in a high–sugar orange juice. The found CGlu are 49.04, 93.27, and 189.54 μM, whose recoveries correspond to 98.08%, 93.27%, and 94.77% with small RSDs ranging 1.61% to 3.08%. In total, the found CGlu in human serum and orange juice are close to the added CGlu, confirming that non–enzymatic Glu sensor of Cu@Co–MOF/GCE can detect Glu with high accuracy and reliability in the environments of rich endogenous biomolecules and high sugar content.

4. Conclusions

A cobalt metal–organic framework (Co–MOF), [Co3(BDC)3(DMU)2], was prepared via an ionothermal reaction, which features three 1D chains along the [110], [011] and b–directions. The chains along the [110] and b–directions interconnect into a 2D [4,4] topological network, and finally shape a 3D 2,6–connected {812.123}{8}3 network by the connection of the [011] chains with the 2D layers. Co–MOF was further synthesized into a Cu@Co–MOF composite material through a sequential deposition–reduction process. Cu@Co–MOF/GCE composite electrode acts as a non–enzymatic electrochemical sensor to detect Glu in 0.01 M NaOH. Cu@Co–MOF/GCE shows excellent electrocatalytic activity for Glu detection. The chronoamperometric response of Cu@Co–MOF/GCE to CGlu displays two linear relationships of j = 0.020·CGlu + 8.266 (within 0.005–0.4 mM Glu) and j = 0.012·CGlu + 6.096 (within 0.4–1.8 mM Glu) at a +0.6 V working potential. The sensitivities of Cu@Co–MOF/GCE are calculated as 282.89 and 113.15 μA mM−1 cm−2, respectively, for the two CGlu ranges. The detection limit is calculated as 1.6 μM at S/N = 3. The low detection limit and high sensitivities of Cu@Co–MOF/GCE to Glu detection are better than or comparable with the other reported MOF–based non–enzymatic electrochemical Glu sensors. Furthermore, Cu@Co–MOF/GCE exhibits a rapid current response of less than 7 s to CGlu change, and a high anti–interference with 0.01 mM D–Man, D–Fru, AA, DA, urea and UA as the interferents. Within the storage time of 15 d, Cu@Co–MOF/GCE shows a gradually decreased current trend of 96.50%, 85.78%, 77.06%, 72.39% and 69.25% for every three days’ Glu electrochemical detection. The nearly 70% current response to Glu indicates an enough sensitivity and stability of Cu@Co–MOF/GCE within 15 days. Five parallel measurements with RSD of 5.84% also demonstrate the high repeatability of Cu@Co–MOF/GCE. Glu detection were carried out in human serum and orange juice. All found CGlu are very close to those added CGlu with low RSDs and high recoveries. It indicates Cu@Co–MOF/GCE as a non–enzymatic electrochemical sensor has a high accuracy and a reliability for Glu detection in real samples with rich endogenous biomolecules and sugars.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemosensors10100416/s1, Figure S1: The coordination modes of BDC2− ligand in Co–MOF; Figure S2: The structure motif of the linear trinuclear [Co3(COO)6(DMU)2] SBU; Figure S3: [Co3(COO)6(DMU)2] SBUs connect BDC2− ligands with O11/O12 (a), O21/O22 (b), and O31/O32 (c) into three 1D chains along the [110], b– and [011] directions, respectively; Figure S4: FT–IR spectra of Co–MOF, Cu@Co–MOF and free H2BDC ligand; Figure S5: TGA curve of Co–MOF; Figure S6: EDS of Co–MOF (a) and Cu@Co–MOF (b); Figure S7: The comparison of XRD patterns of Co-MOF, Cu@Co-MOF, Cu@Co-MOF/Nifion and Cu@Co-MOF/Nifion after the electrochemical detection; Figure S8: The current response of Cu@Co–MOF/GCE to Glu.

Author Contributions

Conceptualization, L.X. and H.J.; data curation, Z.-Z.M. and Y.-S.W.; formal analysis, Z.-Z.M. and Y.-S.W.; funding acquisition, L.X. and H.J.; investigation, B.L. and Z.-Z.M.; methodology, L.X. and H.J.; project administration, L.X.; software, B.L., Y.-S.W. and Z.-Z.M.; validation, B.L.; writing—original draft, Z.-Z.M.; writing—review and editing, L.X. All authors have read and agreed to the published version of the manuscript.

Funding

The project was sponsored by Fundamental Research Funds for the Central Universities (GK202003032), Science and Technology program of Xi’an (201805027YD5CG11(3)), the Natural Science Foundation of Shaanxi province (2018JM2042).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wei, M.; Qiao, Y.X.; Zhao, H.T.; Liang, J.; Li, T.S.; Luo, Y.L.; Lu, S.Y.; Shi, X.F.; Lu, W.B.; Sun, X.P. Electrochemical non-enzymatic glucose sensors: Recent progress and perspectives. Chem. Commun. 2020, 56, 14553–14569. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, D.; Hu, D.; Chen, H.; Shi, G.; Fetahu, I.S.; Wu, F.; Rabidou, K.; Fang, R.; Tan, L.; Xu, S.; et al. Glucose-regulated phosphorylation of TET2 by AMPK reveals a pathway linking diabetes to cancer. Nature 2018, 559, 637–641. [Google Scholar] [CrossRef] [PubMed]
  3. Puttananjegowda, K.; Takshi, A.; Thomas, S. Silicon carbide nanoparticles electrospun nanofibrous enzymatic glucose sensor. Biosens. Bioelectron. 2021, 186, 113285. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, M.; Zhang, Y.; Liu, Q.; Huang, H.; Wang, X.; Shi, Z.; Li, Y.; Liu, S.; Xue, L.; Lei, Y. A smart hydrogel system for visual detection of glucose. Biosens. Bioelectron. 2019, 142, 111547. [Google Scholar] [CrossRef]
  5. Teymourian, H.; Barfidokht, A.; Wang, J. Electrochemical glucose sensors in diabetes management: An updated review (2010–2020). Chem. Soc. Rev. 2020, 49, 7671–7709. [Google Scholar] [CrossRef]
  6. Heller, A.; Feldman, B. Electrochemical glucose sensors and their applications in diabetes management. Chem. Rev. 2008, 108, 2482–2505. [Google Scholar] [CrossRef] [Green Version]
  7. Du, J.; Li, F.; Sun, L. Metal–organic frameworks and their derivatives as electrocatalysts for the oxygen evolution reaction. Chem. Soc. Rev. 2021, 50, 2663–2695. [Google Scholar] [CrossRef]
  8. Dong, Q.; Ryu, H.; Lei, Y. Metal oxide based non-enzymatic electrochemical sensors for glucose detection. Electrochim. Acta 2021, 370, 137744. [Google Scholar] [CrossRef]
  9. Rahmawati, I.; Einaga, Y.; Ivandini, T.A.; Fiorani, A. Enzymatic biosensors with electrochemiluminescence transduction. ChemElectroChem 2022, 9, e202200175. [Google Scholar] [CrossRef]
  10. Khor, S.M.; Choi, J.; Won, P.; Ko, S.H. Challenges and strategies in developing an enzymatic wearable sweat glucose biosensor as a practical point-of-care monitoring tool for type II Diabetes. Nanomaterials 2022, 12, 221. [Google Scholar] [CrossRef]
  11. Osuna, V.; Vega–Rios, A.; Zaragoza–Contreras, E.A.; Estrada–Moreno, I.A.; Dominguez, R.B. Progress of polyaniline glucose sensors for diabetes mellitus management utilizing enzymatic and non-enzymatic detection. Biosensors 2022, 12, 137. [Google Scholar] [CrossRef]
  12. Wang, Q.Q.; Liu, S.Q.; Tang, Z.Y. Recent progress in the design of analytical methods based on nanozymes. J. Mater. Chem. B 2021, 9, 8174–8184. [Google Scholar] [CrossRef]
  13. Huang, W.; Huang, S.M.; Chen, G.S.; Ouyang, G.F. Biocatalytic metal–organic frameworks: Promising materials for biosensing. ChemBioChem 2022, 23, 202100567. [Google Scholar] [CrossRef]
  14. Chen, J.E.; Yang, Z.J.; Koh, H.U.; Shen, J.; Cai, Y.C.; Yamauchi, Y.; Yeh, L.H.; Tung, V.; Wu, K.C.W. Current progress and scalable approach toward the synthesis of 2D metal–organic frameworks. Adv. Mater. Interfaces 2022, 9, 2102560. [Google Scholar] [CrossRef]
  15. Pandey, S.; Nanda, K.K. Au Nanocomposite based chemiresistive ammonia sensor for health monitoring. ACS Sens. 2016, 1, 55–62. [Google Scholar] [CrossRef]
  16. Li, N.; Xu, J.; Feng, R.; Hu, T.L.; Bu, X.H. Governing metal–organic frameworks towards high stability. Chem. Commun. 2016, 52, 8501–8513. [Google Scholar] [CrossRef]
  17. Xie, L.S.; Skorupskii, G.; Dincă, M. Electrically conductive metal–organic frameworks. Chem. Rev. 2020, 120, 8536–8580. [Google Scholar] [CrossRef] [Green Version]
  18. Kajal, N.; Singh, V.; Gupta, R.; Gautam, S. Metal organic frameworks for electrochemical sensor applications: A review. Environ. Res. 2022, 204, 112320. [Google Scholar] [CrossRef]
  19. Adeel, M.; Asif, K.; Rahman, M.M.; Daniele, S.; Canzonieri, V.; Rizzolio, F. Glucose detection devices and methods based on metal–organic frameworks and related materials. Adv. Funct. Mater. 2021, 31, 2106023. [Google Scholar] [CrossRef]
  20. Shi, L.B.; Zhu, X.; Liu, T.T.; Zhao, H.L.; Lan, M.B. Encapsulating Cu nanoparticles into metal–organic frameworks for nonenzymatic glucose sensing. Sens. Actuators B Chem. 2016, 227, 583–590. [Google Scholar] [CrossRef]
  21. Zheng, L.; Liu, Y.; Zhang, C. A sample-to-answer, wearable cloth-based electrochemical sensor (WCECS) for point-of-care detection of glucose in sweat. Sens. Actuators B Chem. 2021, 343, 130131. [Google Scholar] [CrossRef]
  22. Ataei Kachouei, M.; Shahrokhian, S.; Ezzati, M. Bimetallic CoZn–MOFs easily derived from CoZn–LDHs, as a suitable platform in fabrication of a non-enzymatic electrochemical sensor for detecting glucose in human fluids. Sens. Actuators B Chem. 2021, 344, 130254. [Google Scholar] [CrossRef]
  23. Wu, H.Y.; Zheng, W.; Jiang, Y.; Xu, J.C.; Qiu, F.X. Construction of a selective non-enzymatic electrochemical sensor based on hollow nickel nanospheres/carbon dots–chitosan and molecularly imprinted polymer film for the detection of glucose. New J. Chem. 2021, 45, 21676–21683. [Google Scholar] [CrossRef]
  24. Wei, X.F.; Guo, J.L.; Lian, H.T.; Sun, X.Y.; Liu, B. Cobalt metal–organic framework modified carbon cloth/paper hybrid electrochemical button–sensor for nonenzymatic glucose diagnostics. Sens. Actuators B Chem. 2021, 329, 129205. [Google Scholar] [CrossRef] [PubMed]
  25. Shang, N.Z.; Gao, S.T.; Zhou, X.; Feng, C.; Wang, Z.; Wang, C. Palladium nanoparticles encapsulated inside the pores of a metal–organic framework as a highly active catalyst for carbon–carbon cross–coupling. RSC Adv. 2014, 4, 54487–54493. [Google Scholar] [CrossRef]
  26. Deacon, G.B.; Phillips, R.J. Relationships between the carbon–oxygen stretching frequencies of carboxylato complexes and the type of carboxylate coordination. Coord. Chem. Rev. 1980, 33, 227–250. [Google Scholar] [CrossRef]
  27. Wei, Z.; Zhang, Z.H.; Wang, M.M.; Xu, L.; Liu, B.; Jiao, H. Combination effect of ligands and ionic liquid components on the structure and properties of manganese metal–organic frameworks. CrystEngComm 2017, 19, 5402–5411. [Google Scholar] [CrossRef]
  28. Gross, P.; Höppe, H.A. Unravelling the urea–route to boron nitride: Synthesis and characterization of the crucial reaction intermediate ammonium bis(biureto)borate. Chem. Mater. 2019, 31, 8052–8061. [Google Scholar] [CrossRef]
  29. Xue, R.; Guo, H.; Yue, L.G.; Wang, T.; Wang, M.Y.; Li, Q.; Liu, H.; Yang, W. Preparation and energy storage application of a long–life and high rate performance pseudocapacitive COF material linked with –NH– bonds. New J. Chem. 2018, 42, 13726–13731. [Google Scholar] [CrossRef]
  30. Zhang, J.H.; Zhou, T.; Wen, L.; Zhang, A.M. Fabricating metallic circuit patterns on polymer substrates through laser and selective metallization. ACS Appl. Mater. Interfaces 2016, 8, 33999–34007. [Google Scholar] [CrossRef]
  31. Zhang, J.H.; Zhou, T.; Wen, L. Selective metallization induced by laser activation: Fabricating metallized patterns on polymer via metal oxide composite. ACS Appl. Mater. Interfaces 2017, 9, 8996–9005. [Google Scholar] [CrossRef]
  32. Shao, Y.Y.; Gao, Y.; Yue, Q.Y.; Kong, W.J.; Gao, B.Y.; Wang, W.G.; Jiang, W.Q. Degradation of chlortetracycline with simultaneous removal of copper (II) from aqueous solution using wheat straw-supported nanoscale zero–valent iron. Chem. Eng. J. 2020, 379, 122384. [Google Scholar] [CrossRef]
  33. Meher, S.K.; Rao, G.R. Archetypal sandwich-structured CuO for high performance non-enzymatic sensing of glucose. Nanoscale 2013, 5, 2089–2099. [Google Scholar] [CrossRef]
  34. Niu, X.; Li, Y.; Tang, J.; Hu, Y.; Zhao, H.; Lan, M. Electrochemical sensing interfaces with tunable porosity for nonenzymatic glucose detection: A Cu foam case. Biosens. Bioelectron. 2014, 51, 22–28. [Google Scholar] [CrossRef]
  35. Wang, N.; Liang, S.; Zhang, L.J.; Cao, P.F.; Xu, L.; Lin, M. Ionic liquid supported nickel-based metal–organic framework for electrochemical sensing of hydrogen peroxide and electrocatalytic oxidation of methanol. Colloid Surf. A 2020, 603, 125199. [Google Scholar] [CrossRef]
  36. Wang, N.; Dai, H.X.; Sai, L.T.; Ma, H.Y.; Lin, M. Copper ion–assisted gold nanoparticle aggregates for electrochemical signal amplification of lipopolysaccharide sensing. Biosens. Bioelectron. 2019, 126, 529–534. [Google Scholar] [CrossRef]
  37. Xu, Z.H.; Wang, Q.Z.; Hui, Z.S.; Zhao, S.; Zhao, Y.J.; Wang, L. Carbon cloth-supported nanorod-like conductive Ni/Co bimetal MOF: A stable and high-performance enzyme-free electrochemical sensor for determination of glucose in serum and beverage. Food Chem. 2021, 349, 129202. [Google Scholar] [CrossRef]
  38. Kong, X.Y.; Xia, B.; Xiao, Y.W.; Chen, H.H.; Li, H.F.; Chen, W.Z.; Wu, P.; Shen, Y.; Wu, J.S.; Li, S.; et al. Regulation of cobalt–nickel LDHs’ structure and components for optimizing the performance of an electrochemical sensor. ACS Appl. Nano Mater. 2019, 2, 6387–6396. [Google Scholar] [CrossRef]
  39. Zhu, Q.Z.; Hu, S.Y.; Zhang, L.Q.; Li, Y.; Carraro, C.; Maboudian, R.; Wei, W.; Liu, A.R.; Zhang, Y.J.; Liu, S.Q. Reconstructing hydrophobic ZIF–8 crystal into hydrophilic hierarchically–porous nanoflowers as catalyst carrier for nonenzymatic glucose sensing. Sens. Actuators B Chem. 2020, 313, 128031. [Google Scholar] [CrossRef]
  40. Wang, X.D.; Zheng, Y.Y.; Yuan, J.H.; Shen, J.F.; Hu, J.G.; Wang, A.J.; Wu, L.J.; Niu, L. Three-dimensional NiCo layered double hydroxide nanosheets array on carbon cloth, facile preparation and its application in highly sensitive enzymeless glucose detection. Electrochim. Acta 2017, 224, 628–635. [Google Scholar] [CrossRef]
  41. Wang, F.; Chen, X.Q.; Chen, L.; Yang, J.L.; Wang, Q.X. High-performance non-enzymatic glucose sensor by hierarchical flower-like nickel(II)-based MOF/carbon nanotubes composite. Mater. Sci. Eng. C 2019, 96, 41–50. [Google Scholar] [CrossRef] [PubMed]
  42. Sun, F.C.; Wang, S.T.; Wang, Y.Q.; Zhang, J.T.; Yu, X.P.; Zhou, Y.; Zhang, J. Synthesis of Ni–Co hydroxide nanosheets constructed hollow cubes for electrochemical glucose determination. Sensors 2019, 19, 2938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Martín–Yerga, D.; Carrasco–Rodríguez, J.; Fierro, J.L.G.; Alonso, F.J.G.; Costa–García, A. Copper–modified titanium phosphate nanoparticles as electrocatalyst for glucose detection. Electrochim. Acta 2017, 229, 102–111. [Google Scholar] [CrossRef]
  44. Liu, B.B.; Wang, X.Y.; Liu, H.Q.; Zhai, Y.Y.; Li, L.; Wen, H.R. 2D MOF with electrochemical exfoliated graphene for nonenzymatic glucose sensing: Central metal sites and oxidation potentials. Anal. Chim. Acta 2020, 1122, 9–19. [Google Scholar] [CrossRef]
  45. Shahrokhian, S.; Sanati, E.K.; Hosseini, H. Advanced on-site glucose sensing platform based on a new architecture of free-standing hollow Cu(OH)2 nanotubes decorated with CoNi–LDH nanosheets on graphite screen-printed electrode. Nanoscale 2019, 11, 12655–12671. [Google Scholar] [CrossRef]
  46. Chen, J.L.; Yin, H.Y.; Zhou, J.L.; Wang, L.; Gong, J.Y.; Ji, Z.G.; Nie, Q.L. Efficient nonenzymatic sensors based on Ni–MOF microspheres decorated with Au nanoparticles for glucose detection. J. Electron. Mater. 2020, 49, 4754–4763. [Google Scholar] [CrossRef]
  47. Lopa, N.S.; Rahman, M.M.; Ahmed, F.; Sutradhar, S.C.; Ryu, T.; Kim, W. A Ni-based redox–active metal–organic framework for sensitive and non-enzymatic detection of glucose. J. Electroanal. Chem. 2018, 822, 43–49. [Google Scholar] [CrossRef]
  48. Wei, H.G.; Xue, Q.Z.; Li, A.; Wan, T.; Huang, Y.; Cui, D.P.; Pan, D.; Dong, B.B.; Wei, R.B.; Naik, N.; et al. Dendritic core–shell copper–nickel alloy@metal oxide for efficient non-enzymatic glucose detection. Sens. Actuators B Chem. 2021, 337, 129687. [Google Scholar] [CrossRef]
  49. Li, W.W.; Wang, S.L.Y.; Zhang, L.; Cui, X.Q. Nanoporous gold induced vertically standing 2D NiCo bimetal–organic framework nanosheets for non-enzymatic glucose biosensing. Sens. Actuators B Chem. 2019, 281, 652–658. [Google Scholar] [CrossRef]
  50. Lu, M.X.; Deng, Y.J.; Li, Y.C.; Li, T.B.; Xu, J.; Chen, S.W.; Wang, J.Y. Core–shell MOF@MOF composites for sensitive nonenzymatic glucose sensing in human serum. Anal. Chim. Acta 2020, 1110, 35–43. [Google Scholar] [CrossRef]
  51. Wang, X.; Wang, M.Z.; Feng, S.Y.; He, D.P.; Jiang, P. Controlled synthesis of flower-like cobalt phosphate microsheet arrays supported on Ni foam as a highly efficient 3D integrated anode for non-enzymatic glucose sensing. Inorg. Chem. Front. 2020, 7, 108–116. [Google Scholar] [CrossRef]
  52. Yang, N.; Guo, K.L.; Zhang, Y.W.; Xu, C.L. Engineering the valence state of ZIF–67 by Cu2O for efficient nonenzymatic glucose detection. J. Mater. Chem. B 2020, 8, 2856–2861. [Google Scholar] [CrossRef]
  53. Peng, X.R.; Wan, Y.; Wang, Y.Y.; Liu, T.; Zou, P.; Wang, X.X.; Zhao, Q.B.; Ding, F.; Rao, H.B. Flower-like Ni(II)–based metal–organic framework–decorated Ag nanoparticles: Fabrication, characterization and electrochemical detection of glucose. Electroanalysis 2019, 31, 2179–2186. [Google Scholar] [CrossRef]
  54. Arif, D.; Hussain, Z.; Sohail, M.; Liaqat, M.A.; Khan, M.A.; Noor, T. A non-enzymatic electrochemical sensor for glucose detection based on Ag@TiO2@metal–organic framework (ZIF–67) nanocomposite. Front. Chem. 2020, 8, 573510. [Google Scholar] [CrossRef]
  55. Wen, Y.Y.; Meng, W.; Li, C.; Dai, L.; He, Z.X.; Wang, L.; Li, M.; Zhu, J. Enhanced glucose sensing based on a novel composite CoII–MOF/Acb modified electrode. Dalton Trans. 2018, 47, 3872–3879. [Google Scholar] [CrossRef]
  56. He, J.; Yang, H.; Zhang, Y.Y.; Yu, J.; Miao, L.F.; Song, L.F.Y.H.; Wang, L. Smart nanocomposites of Cu–hemin metal–organic frameworks for electrochemical glucose biosensing. Sci. Rep. 2016, 6, 36637. [Google Scholar] [CrossRef] [Green Version]
  57. Song, Y.H.; Shen, Y.; Gong, C.C.; Chen, J.Y.; Xu, M.L.; Wang, L.Y.; Wang, L. A novel glucose biosensor based on Tb@mesoporous metal–organic frameworks/carbon nanotube nanocomposites. ChemElectroChem 2017, 4, 1457–1462. [Google Scholar] [CrossRef]
  58. Shi, L.B.; Li, Y.F.; Cai, X.; Zhao, H.L.; Lan, M.B. ZIF–67 derived cobalt–based nanomaterials for electrocatalysis and nonenzymatic detection of glucose: Difference between the calcination atmosphere of nitrogen and air. J. Electroanal. Chem. 2017, 799, 512–518. [Google Scholar] [CrossRef]
Figure 1. The structure construction of Co–MOF.
Figure 1. The structure construction of Co–MOF.
Chemosensors 10 00416 g001
Figure 2. The experimental XRD patterns of Co–MOF and Cu@Co–MOF.
Figure 2. The experimental XRD patterns of Co–MOF and Cu@Co–MOF.
Chemosensors 10 00416 g002
Figure 3. XPS of Co 2p in (a) Co–MOF, (b) Cu2+/Co–MOF, and (c) Cu@Co–MOF.
Figure 3. XPS of Co 2p in (a) Co–MOF, (b) Cu2+/Co–MOF, and (c) Cu@Co–MOF.
Chemosensors 10 00416 g003
Figure 4. XPS of Cu 2p in (a) Cu2+/Co–MOF and (b) Cu@Co–MOF.
Figure 4. XPS of Cu 2p in (a) Cu2+/Co–MOF and (b) Cu@Co–MOF.
Chemosensors 10 00416 g004
Figure 5. (a) CV curves of Cu@Co–MOF/GCE with Cu2+ concentration within 5–15 mM; (b) CV curves of Cu@Co–MOF/GCE and Co–MOF/GCE in 0.01 M NaOH with (solid line) or without (dash line) 1 mM Glu added.
Figure 5. (a) CV curves of Cu@Co–MOF/GCE with Cu2+ concentration within 5–15 mM; (b) CV curves of Cu@Co–MOF/GCE and Co–MOF/GCE in 0.01 M NaOH with (solid line) or without (dash line) 1 mM Glu added.
Chemosensors 10 00416 g005
Scheme 1. The mechanism of Cu@Co–MOF/GCE sensing Glu.
Scheme 1. The mechanism of Cu@Co–MOF/GCE sensing Glu.
Chemosensors 10 00416 sch001
Figure 6. (a) The chronoamperometric curve of Cu@Co–MOF/GCE to CGlu ranging from 0.005 to 1.8 mM in 0.01 M NaOH at working potential of +0.6 V; (b) the calibration curves of current vs. CGlu.
Figure 6. (a) The chronoamperometric curve of Cu@Co–MOF/GCE to CGlu ranging from 0.005 to 1.8 mM in 0.01 M NaOH at working potential of +0.6 V; (b) the calibration curves of current vs. CGlu.
Chemosensors 10 00416 g006
Figure 7. The anti–interference of Cu@Co–MOF/GCE sensing 0.1 mM Glu with 0.01 mM D–Man, D–Fru, AA, DA, urea and UA as the interferents in 0.01 M NaOH at a working potential of +0.6 V.
Figure 7. The anti–interference of Cu@Co–MOF/GCE sensing 0.1 mM Glu with 0.01 mM D–Man, D–Fru, AA, DA, urea and UA as the interferents in 0.01 M NaOH at a working potential of +0.6 V.
Chemosensors 10 00416 g007
Figure 8. (a) The histogram of current vs. storage time of Cu@Co–MOF/GCE sensing 1.0 mM Glu in 0.01 M NaOH with Cu@Co–MOF/GCE stored at 4 °C for 3–15 days; (b) the histogram of five parallel Glu detections by fresh Cu@Co–MOF/GCE sensing Glu.
Figure 8. (a) The histogram of current vs. storage time of Cu@Co–MOF/GCE sensing 1.0 mM Glu in 0.01 M NaOH with Cu@Co–MOF/GCE stored at 4 °C for 3–15 days; (b) the histogram of five parallel Glu detections by fresh Cu@Co–MOF/GCE sensing Glu.
Chemosensors 10 00416 g008
Table 1. Comparison of detection performances of reported non–enzymatic electrochemical Glu sensors with Cu@Co–MOF.
Table 1. Comparison of detection performances of reported non–enzymatic electrochemical Glu sensors with Cu@Co–MOF.
ElectrodesDetection Limit (μM)Detection Range (mM)Sensitivity (μA·mM−1·cm−2)Refs.
Cu@Co–MOF1.60.005–0.4282.89This work
0.4–1.8113.15
Ni/Co(HHTP)MOF/CC a0.10.3–2.3123250[37]
CoZn–BTC/GCE b4.70.001–0.2551218[22]
0.255–2.53510
Co0.33Ni0.67–HLDH c3.10.01–2242.9[38]
Cu@HHNs d1.970.005–31594.2[39]
NiCo–LDH/CC e0.120.001–1.55.12[40]
Ni(TPA)–SWCNTf4.60.02–4.4[41]
Ni0.7Co0.3(OH)23.420.5–2.51541[42]
CuTiPNPs g70.25–27.81[43]
Co–MOF/EG h0.580.001–3.3330[44]
Cu(OH)2@CoNi–LDH NT–NSs/GSPE i0.60.002–3.21895[45]
3.2–7.71322
Au@Ni–BTC1.50.005–7.41447.1[46]
CPO–27–NiII j1.460.04–640.95[47]
SiCNPs–ENFM k0.560.5–2030.75[3]
Cu–in–ZIF–8/SPCE l2.760–0.7412[20]
Cu–Ni/NF m20.001–0.611340[48]
MWCNTs–PB n4.950.01–1105.93[21]
NiCo–MOF0.290.001–3.8684.4[49]
UiO–67@Ni–MOF0.980.005–3.9203.4[50]
CoPO MA/NF o10.001–1.163.55[51]
Cu2O@ZIF–676.50.01–10307.02[52]
10–16.3181.34
AgNPs/MOF–74 p4.70.01–41.29[53]
Ag@TiO2@ZIF–670.990.048–178.8[54]
CoII–MOF/Acb q1.70.005–1255[55]
Cu–hemin MOFs r2.730.009–3622.77[56]
Tb@mesoMOFs–CNT s80.025–17[57]
Co NP/Porous C5.690.1–1.1227[58]
Notes: a HHTP: triphenylene–2,3,6,7,10,11–hexaol; CC: carbon cloth; b H3BTC: 1,3,5–benzene tricarboxylic acid; GCE: glassy carbon electrode; c HLDH: hollow LDHs; d HHNs = hydrophilic hierarchically porous nanoflowers; e LDHs = layered double hydroxides; f TPA = terephthalic acid; SWCNT = single–walled carbon nanotubes; g CuTiPNPs = copper–modified titanium phosphate nanoparticles; h EG = exfoliated graphene; i NT: nanotube; NSs = Nanosheets; GSPE = graphite screen–printed electrode; j CPO–27–NiII: Ni2(dihydroxyterephthalic acid); k SiCNPs = silicon carbide nanoparticles; ENFM = electrospun–nanofibrous–membrane; l ZIF–8 = zeolitic imidazolate framework 8; SPCE = screen–printed carbon electrodes; m NF = nickel foam; n MWCNTs–PB: multi–walled carbon nanotubes and Prussian blue; o CoPO MA/NF = Cobalt phosphate microsheet arrays supported on Ni foam; p NP = nano particles; q Acb = acetylene black; r mesoMOFs = mesoporous metal–organic frameworks; s CNTs = carbon nanotubes.
Table 2. The Glu sensing in human serum and orange juice by Cu@Co–MOF/GCE.
Table 2. The Glu sensing in human serum and orange juice by Cu@Co–MOF/GCE.
SamplesAdded (μM)Found ± SD (μM)Recovery (%)RSD
Serum5050.01 ± 2.15100.024.30%
10094.93 ± 2.6194.932.75%
200188.97 ± 4.2394.492.24%
Orange juice5049.04 ± 1.6098.081.61%
10093.27 ± 2.8793.273.08%
200189.54 ± 4.4694.772.35%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, Z.-Z.; Wang, Y.-S.; Liu, B.; Jiao, H.; Xu, L. A Non–Enzymatic Electrochemical Sensor of Cu@Co–MOF Composite for Glucose Detection with High Sensitivity and Selectivity. Chemosensors 2022, 10, 416. https://doi.org/10.3390/chemosensors10100416

AMA Style

Ma Z-Z, Wang Y-S, Liu B, Jiao H, Xu L. A Non–Enzymatic Electrochemical Sensor of Cu@Co–MOF Composite for Glucose Detection with High Sensitivity and Selectivity. Chemosensors. 2022; 10(10):416. https://doi.org/10.3390/chemosensors10100416

Chicago/Turabian Style

Ma, Zhen-Zhen, Yue-Shu Wang, Bing Liu, Huan Jiao, and Ling Xu. 2022. "A Non–Enzymatic Electrochemical Sensor of Cu@Co–MOF Composite for Glucose Detection with High Sensitivity and Selectivity" Chemosensors 10, no. 10: 416. https://doi.org/10.3390/chemosensors10100416

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