You are on page 1of 8

Biosensors and Bioelectronics 237 (2023) 115530

Contents lists available at ScienceDirect

Biosensors and Bioelectronics


journal homepage: www.elsevier.com/locate/bios

Regulating highly photoelectrochemical activity of Zr-based mixed-linker


metal-organic frameworks toward sensitive electrogenerated
chemiluminescence sensing of α-glucosidase
Mingzheng Shao a, Yuzhu Sun a, Yuyan Li a, Zhihan Wu a, Xiyan Li b, **, Ruizhong Zhang a, ***,
Libing Zhang a, *
a
Tianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University, Tianjin 300072, China
b
Institute of Photoelectronic Thin Film Devices and Technology, Solar Energy Conversion Center, Key Laboratory of Photoelectronic Thin Film Devices and Technology of
Tianjin, Engineering Research Center of Thin Film Photoelectronic Technology of Ministry of Education, Nankai University, Tianjin, 300350, China

A R T I C L E I N F O A B S T R A C T

Keywords: The conductivity and emission efficiency of metal-organic frameworks (MOFs) remain challenging factors that
Electrogenerated chemiluminescence limit their electrogenerated chemiluminescence (ECL) sensing applications. Herein, we report a facile approach
Metal-organic frameworks to address these challenges by integrating an electroactive linker (H2-TCPP) with an ECL active electrogenerated
Mixed linker
chemiluminescence linker (H4-TBAPy) to construct a highly photoelectrochemical active mixed-linker MOFs
Bandgap regulation
(ML-MOFs). ECL results revealed a remarkable 15.4-fold enhancement for the top-performing ML-MOFs (M6-
α-glucosidase
MOFs), surpassing the single linker MOFs. In addition, M6-MOFs also exhibit a remarkable 73-fold enhancement
in ECL efficiency compared to commercial Ru (bpy)2+ 3 . This improvement should be attributed to the synergistic
effects resulting from the combination of two linkers. Furthermore, M6-MOFs are found to be served as a model
ECLphore for sensitive and selective detection of α-glucosidase for the first time with good potential practica­
bility in human serum samples. This work represents a promising direction to guide for designing good con­
ductivity and high ECL efficiency MOFs in terms of linker functionalization and thus bandgap modulation for
advancing their ECL sensing applications.

1. Introduction Wang et al., 2021b; Zhu et al., 2021). Nevertheless, the efficiency of
electron transfer between the coreactants and ECLphores is constrained
Electrogenerated chemiluminescence (ECL) is a luminescent tech­ by molecular diffusion in the solution, which hampers the amplification
nique that is activated by electrochemical reactions, elicited through of the ECL intensity. Consequently, the integration of high-efficiency
electron transfer between electrogenerated active intermediates, lead­ ECLphores and the nanostructures possessed open channels endowing
ing to the formation of an excited state (Cao et al. 2022, 2023; Cho et al., with efficient mass/electron diffusion capacity could serve as an
2022; Cong et al., 2022; Ding et al., 2002; Gao et al., 2022; Lu et al., essential ECL amplification modality, which could dramatically enhance
2023; Miao, 2008; Wu et al., 2023). By virtue of its facile nature, high the generation and collision efficiency of intermediates produced by the
sensitivity, and quick detection capabilities, ECL has been recognized as coreactant and the ECLphores, thereby substantially elevating the
an indispensable method in the realms of clinical diagnosis, food safety, overall ECL efficiency (Xu et al., 2022a).
bioimaging, and so on (Chen et al., 2022; Feng et al., 2022; Hou et al., Metal-organic frameworks (MOFs) have garnered substantial atten­
2023; Li et al., 2022b; Lv et al., 2022; Ouyang et al., 2023; Wang et al., tion, given their unique properties such as ultrahigh porosity and surface
2023; Wei et al., 2022; Xu et al., 2023; Yu et al., 2022; Zheng et al., area combined with unprecedented tunability, which could facilitate the
2022). To significantly augment the efficiency of ECL, coreactants are mass/electron transfer during the electrochemical process to improve
frequently utilized in tandem with the ECLphores (Tan et al., 2017; the ECL performance (Li et al., 2022a; Zhou et al., 2020). However, the

* Corresponding author.
** Corresponding author.
*** Corresponding author.
E-mail addresses: xiyan.li@nankai.edu.cn (X. Li), zhangrz2019@tju.edu.cn (R. Zhang), libing.zhang@tju.edu.cn (L. Zhang).

https://doi.org/10.1016/j.bios.2023.115530
Received 14 May 2023; Received in revised form 5 July 2023; Accepted 11 July 2023
Available online 16 July 2023
0956-5663/© 2023 Elsevier B.V. All rights reserved.
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

poor conductivity of MOFs has hampered their performance in electro­ the vial and ultra-sounded for another 30 min. The vial was heated to
chemical applications (Jin et al., 2020). To address this limitation, re­ 120 ◦ C for 18 h and then cooled down to room temperature, the product
searchers have introduced an organic linker functionalization strategy was washed three times with DMF and two times with methanol. Finally,
(i.e., mixed-linker), whereby an electroactive organic linker with a the prepared white powder was dried in the vacuum drying oven at
narrow bandgap and strong catalytic activity is involved in MOFs (Jin 120 ◦ C for 24 h for further application. Other details are shown in
et al., 2020). Particularly, the homogeneous incorporation of the Table S1 in Supporting Information.
low-bandgap linkers, such as porphyrin, into MOFs enables the adjust­ Materials and other sections are presented in Supporting
ment of the final bandgap of mixed-linker MOFs (ML-MOFs), thereby Information.
enhancing its ECL performance (Li et al., 2021). Recently, Lei et al. (Jin
et al., 2020) developed an electroactive MOFs by integrating two 3. Results and discussion
organic linkers of hydroquinone-O, O′-diacetic acid, and 1,10-phenan­
throline into the framework, achieving 17-fold enhancement after 3.1. Structure design
pre-reduction. Furthermore, by leveraging the mixed-linker strategy, Ju
et al. (Zhu et al., 2021) found a dual MOFs exhibiting self-enhanced ECL While co-assembling mixed-linker MOFs with well-ordered archi­
without the assistance of the exogenous coreactants, in which 9,10-di tectures can be realized by integrating a variety of topologically distinct
(p-carboxyphenyl) anthracene could act as the robust ECLphore while linkers, there is a greater demand for highly-ordered MOFs through a
1,4-diazabicyclo [2.2.2] octane as the coreactant. Building on these more rational design approach utilizing Zr-based MOFs or other high-
successes, Shan et al. constructed a 2D MOFs by utilizing both zinc (II) valent metal MOFs (Fiankor et al., 2021; Liang et al., 2017). Our
tetrakis (4-carboxyphenyl) porphine (Zn-TCPP) and 2-methylimidazole design strategy is based on two key considerations. Firstly, porphyrin
(MeIm) as the linkers (Li et al., 2022c), in which Zn-TCPP serving as and its derivatives are the electroactive linkers with narrow bandgaps,
the ECLphore and MeIm facilitating electron transfer between Co ions which usually showing outstanding ECL performance (Jin et al., 2020; Li
and oxygen, ultimately yielding exceptional ECL performance. Overall, et al., 2021). Secondly, our previous work has demonstrated the high
the mixed-linker strategy represents a compelling pathway to overcome ECL activity of NU-1000, a (4,8) network composed of a tetrahedral
the conductivity limitation of MOFs in ECL sensing applications. linker (H4-TBAPy) with a csq topology (Shao et al., 2022). Furthermore,
While synthesizing MOFs with a sole linker may be an uncomplicated H4-TBAPy owns a D2h symmetry which is similar with tetratopic
process, the construction of mixed-linker MOFs is indispensable for square-planar ligand of H2-TCPP whose space group is D4h (Fiankor
fabricating multifunctional MOFs, overcoming intrinsic defects, and et al., 2021). Besides, H4-TBAPy shares a similar size (18.208 Å) with
ultimately leading to optimized efficiency. Building upon our prior H2-TCPP (19.911 Å) (Fig. S1), all of which allows for the immobilization
research endeavors (Shao et al., 2022), here two organic linkers of 1,3,6, of two different linkers to construct highly ordered mixed-linker MOFs.
8-Tetra (4′-carboxyphenyl) pyrene (H4-TABPy) and meso-tetra (4-car­
boxyphenyl) porphine (H2-TCPP) served as the principal ECLphore and 3.2. Synthesis and characterization of M6-MOFs
electroactive component, respectively, were co-assembled into the
framework and thus endowing the obtained ML-MOFs with a lower The synthesis of the Zr-based mixed-linker MOFs were achieved via a
bandgap in terms of single linker MOFs. Through precise control of the solvothermal method, whereby a combination of H4-TBAPy and H2-
ratio of two organic linkers, a series of mixed-linker MOFs have been TCPP in different molar ratios was used. Fig. 1A show the schematic
synthesized, designated as M1-MOFs to M8-MOFs. Notably, the optimal illustration for the preparation of the best M6-MOFs (the molar ratio of
M6-MOFs manifested an ECL efficiency of 15.4- and 216.8-fold greater H4-TBAPy and H2-TCPP was 4:1). The morphological microstructure of
than those of the single linker MOFs (NU-1000 made from H4-TABPy, M6-MOFs was elucidated by means of scanning electron microscopy
PCN-222 from H2-TCPP). This outstanding enhancement in ECL per­ (SEM) and transmission electron microscopy (TEM). SEM images
formance can be attributed to the significantly augmented electroactive (Fig. 1B and C) reveal that the as-synthesized M6-MOFs exhibits a uni­
surface area giving the superior mass/electron transfer, as well as the form fusiform-like shape with an average diameter of 1 μm and length of
synergistic interplay between the two organic linkers from the optimal 5 μm (insets in Fig. 1B). TEM image (Fig. 1D) and the corresponding
M6-MOFs. Furthermore, a comprehensive analysis of the energy levels elemental maps (Fig. 1E–H) of M6-MOFs confirm the homogeneous
of mixed-linker MOFs disclosed the critical role of bandgap modulation distribution of C, N, O and Zr throughout the structure. In conjunction
in achieving maximum ECL performance. with energy dispersive spectrum (EDS) analysis (Fig. S2), these results
Significantly, the M6-MOFs, serving as the optimal ECLphore, have collectively confirm the successful synthesis of M6-MOFs.
been demonstrated its prowess as a cutting-edge tool for the ultrasen­ X-ray photoelectron spectroscopy (XPS) was utilized to elucidate the
sitive detection of α-glucosidase, with an unprecedentedly low limit of element composition and chemical state of M6-MOFs. The XPS survey
detection (LOD) of 0.00088 U mL− 1. α-Glucosidase, a carbohydrate- spectrum depicted in Fig. 2A, exhibiting peaks at 182.6, 284.8, 399.9
hydrolase that exists in intestinal epithelial cells and plays a critical and 531.2 eV, which can be attributed to Zr 3d, C 1s, N 1s and O 1s,
role in regulating blood sugar levels, can cause hyperglycemia, diabetes, respectively, confirming the presence of C, N, O, and Zr. Fig. S3A shows
and other disorders when its levels exceed normal physiological ranges the high-resolution XPS spectrum of Zr 3d spectrum with two peaks at
(Ao et al., 2017; Chiba, 1997; Wang et al., 2021a). It is noteworthy that 182.7 and 185.1 eV which belong to Zr 3d3/2 and Zr 3d5/2 of Zr. The C 1s
this is the first time that the α-glucosidase has been detected by using the high-resolution XPS spectrum of M6-MOFs (Fig. S3B) exhibits four peaks
ECL method. Additionally, the M6-MOFs based α-glucosidase ECL sensor with binding energies at 284.8, 286.6, 288.6, and 291.1 eV, which can
exhibited exceptional selectivity, underscoring its immense potential for be attributed to the sp2 C–C bonds, the carbon from the pyrrole ring, and
a broad range of bioanalytical applications. the carbon from carbonyl groups (Shao et al., 2022). For N 1s (Fig. S3C),
two fitted peaks at 397.7 and 399.9 eV are assigned to the pyrrolic ni­
2. Experimental section trogen (–NH–) and the imine nitrogen (-C– – N-) of pyrrole, respectively
(Da Silva et al., 2018). Its O 1s XPS spectrum (Fig. S3D) shows two peaks
2.1. Synthesis of ML-MOFs at 531.1 and 532.0 eV, which could be ascribed to the (-C-OH) and
(-C–– O), respectively (Da Silva et al., 2018). Collectively, the XPS
The ML-MOFs were prepared by mixing ZrOCl2⋅8H2O (97 mg), 40 μL spectra provide compelling evidence for the formation of M6-MOFs.
trifluoroacetic acid, and 1.6 g benzoic acid into a glass vial with 8 mL M6-MOFs was further subjected to rigorous characterization to
DMF. Then, the mixture was ultra-sounded for 10 min to dissolve all the elucidate its crystal structure, porosity, and thermal stability. The crystal
agents. Then different masses of H4-TBAPy and H2-TCPP were added to structure of M6-MOFs was firstly characterized by powder X-ray

2
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

Fig. 1. (A) Schematic illustration for the preparation of M6-MOFs. (B) SEM and (C) HRSEM images of M6-MOFs, the insets show the length (blue) and diameter (red)
distribution histograms of M6-MOFs. (D) TEM images of M6-MOFs and the corresponding elemental maps of Zr (E), C (F), N (G) and O (H).

diffraction (PXRD). The PXRD pattern of M6-MOFs (Fig. 2B) is consis­ molecules (DMF/water) in the pores. The weight losses at 368 ◦ C and
tent with the simulated single linker MOFs (NU-1000 and PCN-222), 500 ◦ C correspond to the removal of H2-TCPP and H4-TBAPy, respec­
which suggests that M6-MOFs maintains the structure assembled from tively. These results demonstrate the good thermal stability of
H4-TBAPy and H2-TCPP with Zr6O8 cluster (Bai et al., 2016; Li et al., M6-MOFs.
2022c). However, more detailed observation on M6-MOFs peak loca­
tions find a pronounced leftward shift when compared to the simulated 3.3. Optical property of M6-MOFs
XRD pattern for NU-1000 and a rightward shift relative to the simulated
XRD pattern for PCN-222 (inset in Fig. 2B). These shifts are attributed to The formation of M6-MOFs was also confirmed by the UV–vis ab­
a change in the lattice constant of M6-MOFs resulting from size differ­ sorption and photoluminescence (PL) spectroscopies. As depicted in
ences of the two linkers (Fig. S1) (Ren et al., 2022). The porosity Fig. 2E, the optical spectrum of M6-MOFs exhibits an absorption band at
property of M6-MOFs was then characterized by N2 absorption at 77 K. 413 nm which could be ascribed to the H4-TBAPy and characteristic four
As shown in Fig. 2C, M6-MOFs exhibit a type IV isotherm, with a bands of 531, 564, 597 and 648 nm, ascribed to the Q bands of H2-TCPP
Brunauer-Emmett-Teller (BET) surface area of 2336.73 m2 g− 1 and an (Xu et al., 2022b). Upon excitation at 400 nm, the PL spectrum of
average pore size distribution of 2.31 nm by BJH model (inset in M6-MOFs (Fig. 2E) displays three peaks at 470 nm (from H4-TBAPy
Fig. 2C), suggesting the high surface area and mesoporous structure of emission), and 660 and 720 nm (from H2-TCPP emission). The results of
M6-MOFs. These findings suggest that M6-MOFs possess an extensive UV–vis and PL spectra support the presence of two linkers in the
surface area and highly mesoporous structure, thereby facilitating M6-MOFs. Additionally, there is significant spectra overlap between the
electron and reactive intermediate conduction during electrochemical absorption of H2-TCPP and emission of H4-TBAPy, along with proximity
processes. The thermal stability of M6-MOFs was assessed by thermog­ between H4-TBAPy and H2-TCPP in NU-1000 and PCN-222, respectively
ravimetric analysis (TGA) under N2 (Fig. 2D). The initial weight loss (Shao et al., 2022). These results indicate the feasibility of the Förster
before 200 ◦ C could be ascribed to the removal of the guest solvent resonant energy transfer (FRET) happening between the energy donor of

3
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

Fig. 2. (A) XPS survey spectrum of as-prepared M6-MOFs. (B) XRD patterns of NU-1000, M6-MOFs and PCN-222 (inset shows the magnified images of XRD pattern
in low degree range). (C) N2 absorption-desorption isotherm at 77 K and BJH average pore size distribution (the inset) of M6-MOFs. (D) Thermogravimetric analysis
of M6-MOFs. (E) UV–vis spectra and PL emission spectra of M6-MOFs. The inset shows the CIE diagram of M6-MOFs. (F) PL lifetime curve of NU-1000, PCN-222 and
M6-MOFs.

H4-TBAPy and the energy acceptor of H2-TCPP (Fiankor et al., 2021). absorption features from both H4-TBAPy and H2-TCPP (Fig. S15A).
Time-resolved emission experiments employ the time-correlated sin­ Remarkably, the ratio of the emission peaks at 470 nm and 660 nm for
gle-photon counting (TCSPC) technique were employed to confirm the the ML-MOFs exhibits a pronounced increase and reached its maximum,
presence of FRET (Fig. 2F). Compared with the average lifetime of the giving rise to a diverse array of emission colors for the ML-MOFs
NU-1000 (5.30 ns collected at 470 nm, Table S2) and the PCN-222 (2.38 (Figs. S15B–D).
ns collected at 660 nm, Table S2), the emission of the average lifetime of
H4-TBAPy in M6-MOFs is shortened (1.02 ns, Table S2) while H2-TCPP
in M6-MOFs is prolonged (6.06 ns, Table S2), which is consistent with 3.5. ECL enhancement mechanism of ML-MOFs
the FRET process assigned above (Fiankor et al., 2021), and which is also
evidence for constructing M6-MOFs successfully. The rational linker ratio of electroactive H2-TCPP and high-efficiency
ECLphore of H4-TBAPy, as well as the channels in ML-MOFs, endow
them with enormous potential for further enhancing ECL performance.
3.4. Regulation of the linker ratios To this end, the ECL behaviors of ML-MOFs were first investigated, as
indicated in Fig. 3A and S16-17 that the ECL intensity of ML-MOFs rises
In our design, the implementation of the narrow bandgap linker H2- sharply with increasing the proportion of H2-TCPP, and reaches a
TCPP facilitates bandgap regulation, ultimately enhancing the conduc­ highest value of 19127.1 a. u. For M6-MOFs in the presence of 50 mM
tivity of MOFs. Another linker, H4-TBAPy, functionalized as a potent K2S2O8. Followed by a decrease as the H2-TCPP ratio further increasing,
ECLphore. Nevertheless, the ECL performance of PCN-222 made from signified that the H4-TBAPy/H2-TCPP ratio plays a critical role in the
H2-TCPP is much weaker than the NU-1000 whose linker is H4-TBAPy ECL performance of ML-MOFs. To demonstrate the superior ECL per­
(Shao et al., 2022). Consequently, the linker ratio can be fine-tuned to formance from M6-MOFs, series control experiments conducted from
greatly augment the ECL performance of ML-MOFs. Various ML-MOFs the single-linker MOFs (NU-1000 and PCN-222), the pure ECL active
with varying linker ratios were synthesized and assigned designations linker (H4-TBAPy), as well as the physical mixture of NU-1000 and PCN-
ranging from M1-MOFs to M8-MOFs (Table S1). Except for M6-MOFs 222 (the molar ratio between NU-1000 and PCN-222 is 4:1, similar with
described above, the morphological microstructures of NU-1000, M6-MOFs confirmed by the 1HNMR, Fig. S18). As shown in Fig. 3B, the
PCN-222, and other ML-MOFs were also revealed by the SEM M6-MOFs exhibit the strongest ECL intensity among others. And the ECL
(Figs. S4–S12). NU-1000 and PCN-222 exhibit short-rod morphology efficiency of them is determined and summarized in Table S3. Impres­
(Figs. S4–S5). For the ML-MOFs, as illustrated in Figs. S6–S12, ML-MOFs sively, the ECL efficiency of M6-MOFs is enhanced by 15.4-, 216.8-, 3.9-,
underwent transformations from dumbbell-like to fusiform-like, and and 8.2-fold compared to NU-1000, PCN-222, the physical mixture of
ultimately to rod-like shapes with the increase in H4-TBAPy ratio, likely NU-1000 and PCN-222, and H4-TBAPy, respectively. Compared to
due to disparities in size between H2-TCPP and H4-TBAPy. Notably, the commercial Ru (bpy)2+ 3 standard, M6-MOFs also exhibit 73-fold ECL
lengths of all prepared ML-MOFs are maintained at the micron level efficiency, showing its superior ECL performance. Such enhancement
(Fig. S13). The phase purity of synthesized ML-MOFs was assessed by can be attributed to the porous frame structure of M6-MOFs and rational
PXRD. As shown in Fig. S14, the phase purity of ML-MOFs has a strong ration of the two distinctive linkers (Zhu et al., 2021). Furthermore, the
tolerance on the linker ratios and the main peaks of these ML-MOFs also ECL intensity of M6-MOFs under consecutive potential scans remained
have slight shifts compared with both PCN-222 and NU-1000. very stable, with a relative standard deviation (RSD) of 0.69%, as
The optical properties of as-prepared ML-MOFs were all evaluated by demonstrated in Fig. 3C. To gain further insights into the ECL
means of the UV–vis absorption and PL spectroscopies (Figs. S15A–B). enhancement from M6-MOFs, here the electronic energy level structures
The UV–vis absorption spectra of all ML-MOFs show the presence of the of M6-MOFs, NU-1000 and PCN-222 were explored by valence band XPS

4
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

Fig. 3. (A) ECL intensity of as-prepared various ML-MOFs. (B) ECL-voltage curves for M6-MOFs, NU-1000, PCN-222, TBAPy and physically mixed linkers. (C)
Continuous ECL signals of 20 μg M6-MOFs modified GCE. All the measurements are operated with a potential scan between 0 and − 1.9 V in 0.1 M PB (pH = 7.4) with
50 mM K2S2O8 as the coreactant and 0.1 M KCl as the supporting electrolyte. The scan rate was 100 mV s− 1. (D) XPS VB spectrum of M6-MOFs. (E) Solid UV–Vis
diffuse reflection absorption spectrum and corresponding Tauc plots of M6-MOFs. (F) Energy-level plots of M6-MOFs, NU-1000 and PCN-222.

(VB-XPS) spectra and UV–vis diffuse reflectance spectrum (UV-DRS). spectrum of NU-1000 and PCN-222 are given in Figs. S19–20. Their
Fig. 3D shows the XPS VB spectrum of M6-MOFs and discloses the VB corresponding energy levels of M6-MOFs, NU-1000 and PCN-222 are
edge at 1.58 eV. The optical band gap evaluated by the UV-DRS spectra shown in Fig. 3F. It is evident that M6-MOFs possess the narrowest
is 1.81 eV, which is given in Fig. 3E. And the conduction band minimum bandgap, making it highly efficient for electron transfer and light
(CBM) is estimated to be − 0.23 eV. The VB spectrum and UV-DRS emission. Moreover, M6-MOFs possess a lower CBM value (− 0.23 eV),

Fig. 4. (A) Square wave voltammogram of 20 μg M6-MOFs modified Pt electrode in 0.1 M DCM solution containing 0.1 M TBAP as the supporting electrolyte. (B) The
spooling ECL spectra of 20 μg M6-MOFs modified on GCE in the presence of 50 mM K2S2O8 in 0.1 M PB (pH = 7.4) containing 0.1 M KCl. The insets show the spooled
ECL spectra: (red) evolution and (blue) devolution. (C) Accumulated ECL spectrum of the same M6-MOFs/K2S2O8 system.

5
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

indicating its stronger ability to accept electrons from the electrode and to generate M6-MOFs (TBAPy)•+ (Equation (6)) and act as an efficient
thus easier reduction during the ECL process (Gao et al., 2021). This electron acceptor, receiving an electron from the electrode to generate
indicates the success of our strategy to regulate the band gap to improve M6-MOFs (TBAPy)•- (Equation (7)) and facilitate the emission of light
the ECL performance. (512 nm, Equations (8) and (9)). In light of these findings, the ECL
mechanisms of M6-MOFs can be articulated by the ensuing equations:
3.6. ECL mechanism of M6-MOFs
S2 O2−8 + e− → SO2−4 + SO•−4 (1)
To further verify the electrochemical property of M6-MOFs, square
M6 − MOFs(TCPP) + SO•−4 →M6 − MOFs (TCPP)•+ + SO2−4 (2)
wave voltammetry (SWV) and cyclic voltammetry (CV) were employed
to explore the redox behavior of M6-MOFs. As shown in Fig. 4A, M6-
M6 − MOFs(TCPP) + e− → M6 − MOFs (TCPP)•− (3)
MOFs undergoes two main oxidation process with peak potentials at
1.07 and 1.45 V vs. Ag/AgCl, which could be ascribed to the generation
M6 − MOFs (TCPP)•+ + M6 − MOFs (TCPP)•− →M6 − MOFs(TCPP) + hν1
of H2-TCPP and H4-TBAPy cation radicals. By scanning the potential in a
(4)
reverse direction, one main reduction peaks could be observed at − 1.12
V vs. Ag/AgCl which could be ascribed to the reduction of H2-TCPP to
M6 − MOFs (TCPP)•− + SO•−4 →M6 − MOFs(TCPP) + SO2−4 + hν1 (5)
form TCPP anion radical (Goswami et al., 2021; Li et al., 2022c; Yu et al.,
2018). For a real-time monitoring evolution and devolution of ECL
M6 − MOFs(TBAPy) + SO•−4 → M6 − MOFs (TABPy)•+ + SO2−4 (6)
during this potentiodynamic process, spooling ECL spectroscopy was
performed. As shown in Fig. 4B, the spooling ECL spectra of M6-MOFs M6 − MOFs(TBAPy) + e− →M6 − MOFs (TBAPy)•− (7)
show a peak of 690 nm at − 0.8 V vs. Ag/AgCl, which is corresponding
to the ECL emission from H2-TCPP in M6-MOFs. And a second peak at M6− MOFs(TABPy)•+ +M6− MOFs(TABPy)•− →M6− MOFs(TBAPy)+hν2
512 nm could be ascribed to the ECL emission from H4-TBAPy when
(8)
scanning at or more negatively than − 1.75 V vs. Ag/AgCl, which cor­
responds to the formation of excited state of H4-TBAPy* (Goswami et al.,
M6 − MOFs (TBAPy)•− + SO•−4 →M6 − MOFs(TBAPy) + SO2−4 + hν2
2021; Yu et al., 2018). Besides, compared with the spooling ECL spectra
(9)
of single-linker PCN-222 and NU-1000 (Figs. S21–22), an apparently
ECL energy transfer for ML-MOFs would be occurred at − 1.85 V or more
negative potentials (Figs. S23–29). Furthermore, the ECL accumulated 3.7. ECL detection α-glucosidase based on M6-MOFs
spectrum (Fig. 4C) acquired by executing two potential sweeping scans,
manifesting two discernible peaks at 512 and 690 nm, which is in good Given the exceptional ECL performance, the M6-MOFs were utilized
coincide with its spooling ECL spectra. It is noteworthy that only one to design a sensor platform for detecting the presence of α-glucosidase.
ECL emission peak (690 nm) from H2-TCPP in M6-MOFs can be Fig. 5A depicts the schematic representation of the detection strategy
observable, which is distinct with its PL (two emission peaks at 660 and based on the ECL quenching effect of hydroquinone, in which hydro­
720 nm, Fig. 2E). This is probably due to their different concentrations, quinone could be oxidized into benzoquinone by the SO•- 4 (Gao et al.,
as PL obtained in micron molar while ECL in millimolar. Similar 2022) and result in a ECL intensity quenching of M6-MOFs. Under the
red-shifted wavelength of ECL from H4-TBAPy in M6-MOFs (512 nm) catalytic effect of α-glucosidase, the hydrolysis of α-arbutin generates
can be seen when compared to its PL (470 nm, Fig. 2E), all these changes hydroquinone, which further leads to a reduction in the ECL intensity of
are likely to a result of the inner-filter effect brought about by M6-MOFs. The level of ECL quenching is directly proportional to the
self-absorption (Maar et al., 2019). concentration of hydroquinone produced as a result of α-glucosidase-­
To gain a comprehensive understanding of the ECL mechanism of the mediated hydrolysis of α-arbutin.
M6-MOFs, we proceeded to estimate the enthalpy of the radical ion To further validate the effectiveness of our hypothesis, the impact of
reaction through the employment of the equation: ΔHK2S2O8 = E (SO•- 4/ hydroquinone on the ECL intensity was investigated. As shown in Fig. 5B
SO2− 2−
4 ) - ERed1 - 0.16 (eV), where E (SO4 /SO4 ) exhibited a value of 3.15
•-
and S30A that ECL intensity gradually decreased as the concentration of
V (Miao, 2008; Zhang et al., 2021). The consequential enthalpy values hydroquinone increased, indicating the quenching effect of hydroqui­
for H2-TCPP and H4-TBAPy in M6-MOFs were determined to be 4.05 eV none. Additionally, the ECL intensity of M6-MOFs was found to be lin­
and 4.74 eV, respectively. Moreover, an estimate was derived for the early related to the concentration of hydroquinone, as plotted in
requisite energy to produce the lowest excited singlet state (Es) of Fig. S30B, within the range of 0.01–1 mmol L− 1. Moreover, the effects of
M6-MOFs as 1.81 eV (as illustrated in Fig. 3F), a reflection of the α-arbutin, α-glucosidase and the combination of α-arbutin and α-gluco­
energy-sufficient ECL system of M6-MOFs/K2S2O8 and following by sidase on the ECL of M6-MOFs were also evaluated. As illustrated in
S-route, leading to a high-efficiency ECL system. Fig. 5B, there are negligible effects on ECL performance of M6-MOFs
Based on the preceding discourse, the possible ECL mechanisms of when α-arbutin or α-glucosidase existed independently. By contrast,
M6-MOFs were proposed. Firstly, K2S2O8 would be reduced to generate the ECL intensity dramatically decreases in the presence of both
a highly potent oxidizing intermediate (SO•- 4 , Equation (1)) as its peak α-arbutin and α-glucosidase, demonstrating the feasibility of our strat­
reduction potential (~− 0.9 V) is much positive than M6-MOFs (− 1.12 egy. The hydrolysis reaction time of α-glucosidase was initially deter­
V, Fig. 4A). This intermediate then can oxidize H2-TCPP in M6-MOFs to mined (as exhibited in Fig. S31), the quenching degree of the sensing
result in the formation of M6-MOFs (TCPP)•+ (Equation (2)). Subse­ platform significantly increases with an increase of the hydrolysis re­
quently, the H2-TCPP within M6-MOFs undergoes electrochemical action time from 0 to 60 min. Under the appropriate incubation time of
reduction to form the M6-MOFs (TCPP)•- (Equation (3)) by scanning 60 min, the ECL intensity of the sensing platform containing α-arbutin
potential in further negative values, and the M6-MOFs (TCPP)•- could (10 μmol L− 1) was measured after adding varying levels of α-glucosi­
annihilate with former M6-MOFs (TCPP)•+ to generate its excited singlet dase. Fig. 5C depicts the declining trend in the ECL response of the
state, which relaxed to the ground state and emitted at 690 nm (Equa­ sensing platform as the concentration of α-glucosidase increases. As
tion (4)). In addition, SO•-4 could remove an electron from the highest plotted in Fig. 5D and the inset, the logarithm of the concentration of
occupied molecular orbital (HOMO) of the electrochemical formed M6- α-glucosidase from 0.001 to 0.1 U mL− 1 exhibits an excellent linear
MOFs (TCPP)•- to produce the excited singlet state and thereby show correlation. The regression equation of the calibration curve is IECL =
light emission as well (690 nm, Equation (5)). Furthermore, the H4- − 7243.33 lg [Cα-glucosidase] - 2532.98 with a coefficient of 0.997. The
TBAPy in M6-MOFs could remove an electron with the presence of SO•- 4 limit of detection (LOD) was thus estimated to be as low as 0.00088 U

6
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

Fig. 5. (A) Diagrammatic illustration of α-glucosi­


dase detection based on the quenching effect of the
hydroquinone. (B) ECL-voltage curves for 20 μg M6-
MOFs modified GCE with none (red), α-arbutin
(green), α-glucosidase (yellow), the combination of
α-arbutin and α-glucosidase (purple) and hydroqui­
none (blue). (C, D) The correlation between ECL in­
tensity and the concentration of α-glucosidase. The
inset of (D) shows the linearity of the ECL intensity on
the logarithm of α-glucosidase concentration. (E)
Selectivity and anti-interference of the ECL sensor for
interferences, and the mixture of α-glucosidase with
various interferences. The concentration of α-gluco­
sidase was 0.1 U mL− 1, other bioenymes were 1.0 U
mL− 1 and biomolecules were 1.0 mM.

mL− 1 (3σ/k), suggesting that the proposed sensing platform is highly


Table 1
sensitive and possess great potential in detecting α-glucosidase. Notably,
The recovery results of α-glucosidase in human serum samples.
this is the first time that α-glucosidase has been determined using the
ECL method. As summarized in Table S4, the detection performance of Sample Added (U mL− 1) Found (U mL− 1) Recovery (%) RSD (%, n = 3)
current work is comparable or even better than those of reported. 1 0 0.782 – 3.304
Especially, the LOD is better than most of reported methods, demon­ 2 0.5 1.269 97.4 1.722
3 1 1.795 101.3 1.529
strating the outstanding sensitivity of current sensing platform resulted
4 5 5.907 102.5 1.302
from both the high ECL efficiency of M6-MOFs and low background of
ECL technique (Miao, 2008).
To evaluate the selectivity of the proposed ECL sensing platform, of α-Glu (0.5, 1, and 5 U mL− 1) were added, with a corresponding
various potential interferences, such as glucose oxidase (GOx), acetyl­ relative standard deviation (RSD) of less than 1.7% (n = 3). These results
cholinesterase (Ace), glucose (Glu), bovine serum albumin (BSA), and further confirm the reliability and accuracy of the sensing platform.
glutathione (GSH), were investigated in the presence and absence of
α-glucosidase. By mixing 0.05 U mL− 1 α-glucosidase with 10- fold or 4. Conclusion
more interferences, the ECL intensity of the sensor remained almost
unchanged (Fig. 5E). The results demonstrate that the interfering sub­ In conclusion, this study presents a novel approach to construct a
stances had minimal impact on the ECL performance of the M6-MOFs, series of mixed-linker MOFs using ECL active H4-TBAPy in combination
indicating the high specificity of the sensing platform. Furthermore, with electroactive and narrow-bandgap H2-TCPP as organic linkers.
the stability and reproducibility were investigated to evaluate the reli­ Among them, M6-MOFs exhibited the highest ECL efficiency, surpassing
ability of the ECL sensor. The ECL intensity was measured using four corresponding single-linker MOFs (NU-1000, PCN-222), as well as
distinct electrodes that were incubated with 0.005 U mL− 1 α-Glucosi­ physical mixture of NU-1000 and PCN-222, and H4-TBAPy alone. This
dase. As shown in Fig. S32, the ECL intensity of four individual sensors highlights the effectiveness of bandgap regulation through mixed-linker
show quite similar values with a relative standard deviation (RSD) of strategy. The ECL enhancement mechanism of M6-MOFs was attribut­
2.31%, demonstrating an excellent stability and reproducibility of the able to its smallest energy bandgap. Furthermore, M6-MOFs was utilized
ECL sensing platform. to construct a ECL sensing platform for detecting α-glucosidase,
Given that excellent performance, M6-MOFs was further used to demonstrating high sensitivity, good selectivity and practicability. By
determine the α-glucosidase in human serum samples. The serum sam­ focusing on linker functionalization and bandgap modulation, this work
ples were diluted 100-fold before analysis. As illustrated in Table 1, the provides valuable insights to guide the design and utilization of MOFs
recovery of α-Glu ranged from 97.3% to 102.5% when different amounts with improved properties for ECL sensing applications.

7
M. Shao et al. Biosensors and Bioelectronics 237 (2023) 115530

CRediT authorship contribution statement Feng, Y., Wang, N., Ju, H., 2022. Sci. China Chem. 65, 2417–2436.
Fiankor, C., Nyakuchena, J., Khoo, R.S.H., Zhang, X., Hu, Y., Yang, S., Huang, J.,
Zhang, J., 2021. J. Am. Chem. Soc. 143 (48), 20411–20418.
Mingzheng Shao: Validation, Writing – original draft. Yuzhu Sun: Gao, H., Wei, X., Li, M., Wang, L., Wei, T., Dai, Z., 2021. Small 17, 2103424.
Validation, Investigation. Yuyan Li: Formal analysis, Investigation. Gao, N., Zeng, H., Wang, X., Zhang, Y., Zhang, S., Cui, R., Zhang, M., Mao, L., 2022.
Zhihan Wu: Investigation. Xiyan Li: Conceptualization. Ruizhong Angew. Chem. Int. Ed. 61, e202204485.
Goswami, S., Yu, J., Patwardhan, S., Deria, P., Hupp, J.T., 2021. ACS Energy Lett. 6 (3),
Zhang: Conceptualization, Writing – review & editing, Resources, 848–853.
Funding acquisition, Supervision. Libing Zhang: Conceptualization, Hou, Y., Fang, Y., Zhou, Z., Hong, Q., Li, W., Yang, H., Wu, K., Xu, Y., Cao, X., Han, D.,
Writing – review & editing, Resources, Funding acquisition, Supervision. Liu, S., Shen, Y., Zhang, Y., 2023. Adv. Opt. Mater. 11, 2202737.
Jin, Z., Zhu, X., Wang, N., Li, Y., Ju, H., Lei, J., 2020. Angew. Chem. Int. Ed. 59, 10446.
Li, B., Huang, X., Lu, Y., Fan, Z., Li, B., Jiang, D., Sojic, N., Liu, B., 2022a. Adv. Sci. 9 (35),
Declaration of competing interest 2204715.
Li, J., Yang, H., Cai, R., Tan, W., 2022b. ACS Appl. Mater. Interfaces 14 (39),
44222–44227.
The authors declare that they have no known competing financial Li, Y.-J., Cui, W.-R., Jiang, Q.-Q., Wu, Q., Liang, R.-P., Luo, Q.-X., Qiu, J.-D., 2021. Nat.
interests or personal relationships that could have appeared to influence Commun. 12 (1), 4735.
the work reported in this paper. Li, Y.-X., Li, J., Zhu, D., Wang, J.-Z., Shu, G.-F., Li, J., Zhang, S.-L., Zhang, X.-J.,
Cosnier, S., Zeng, H.-B., Shan, D., 2022c. Adv. Funct. Mater. 32, 2209743.
Liang, C.-C., Shi, Z.-L., He, C.-T., Tan, J., Zhou, H.-D., Zhou, H.-L., Lee, Y., Zhang, Y.-B.,
Data availability 2017. J. Am. Chem. Soc. 139 (38), 13300–13303.
Lu, Y., Huang, X., Wang, S., Li, B., Liu, B., 2023. ACS Nano 17 (4), 3809–3817.
Lv, H., Zhang, R., Cong, S., Guo, J., Shao, M., Liu, W., Zhang, L., Lu, X., 2022. Anal.
Data will be made available on request.
Chem. 94 (10), 4538–4546.
Maar, R.R., Zhang, R., Stephens, D.G., Ding, Z., Gilroy, J.B., 2019. Angew. Chem. Int. Ed.
Acknowledgements 58 (4), 1052–1056.
Miao, W., 2008. Chem. Rev. 108 (7), 2506–2553.
Ouyang, X., Wu, Y., Guo, L., Li, L., Zhou, M., Li, X., Liu, T., Ding, Y., Bu, H., Xie, G.,
This work was supported by the National Natural Science Foundation Shen, J., Fan, C., Wang, L., 2023. Angew. Chem. Int. Ed. 62, e202300893.
of China (No. 21904095 and 22004089), Peiyang Talents Project of Ren, H., Zheng, L., Li, Y., Ni, Q., Qian, J., Li, Y., Li, Q., Liu, M., Bai, Y., Weng, S.,
Tianjin University, Young Thousand Talented Program, Program of Wang, X., Wu, F., Wu, C., 2022. Nano Energy 103, 107765.
Shao, M., Li, Y., Chen, M., Liu, W., Sun, Y., Chu, Y., Sun, Y., Li, X., Zhang, R., Zhang, L.,
Tianjin Science and Technology Major Project and Engineering 2022. Chemelectrochem 9, e202200866.
(19ZXYXSY00090) and Open Funds of the State Key Laboratory of Tan, X., Zhang, B., Zou, G., 2017. J. Am. Chem. Soc. 139 (25), 8772–8776.
Electroanalytical Chemistry (SKLEAC202304). Wang, M., Zhou, X., Wang, X., Wang, M., Su, X., 2021a. Sensor. Actuat. B-Chem. 345,
130407.
Wang, Y., Zhao, G., Chi, H., Yang, S., Niu, Q., Wu, D., Cao, W., Li, T., Ma, H., Wei, Q.,
Appendix A. Supplementary data 2021b. J. Am. Chem. Soc. 143 (1), 504–512.
Wang, Y., Ding, J., Zhou, P., Liu, J., Qiao, Z., Yu, K., Jiang, J., Su, B., 2023. Angew.
Chem. Int. Ed. 62, e202216525.
Supplementary data to this article can be found online at https://doi. Wei, X., Chu, K., Adsetts, J.R., Li, H., Kang, X., Ding, Z., Zhu, M., 2022. J. Am. Chem. Soc.
org/10.1016/j.bios.2023.115530. 144 (44), 20421–20433.
Wu, K., Chen, R., Zhou, Z., Chen, X., Lv, Y., Ma, J., Shen, Y., Liu, S., Zhang, Y., 2023.
Angew. Chem. Int. Ed. 62, e202217078.
References
Xu, W., Zhao, X., Zhan, F., He, Q., Wang, H., Chen, J., Wang, H., Ren, X., Chen, L., 2022a.
Energy Storage Mater. 53, 79–135.
Ao, H., Feng, H., Huang, X., Zhao, M., Qian, Z., 2017. J. Mater. Chem. C 5 (11), Xu, W., Wu, Y., Wang, X., Qin, Y., Wang, H., Luo, Z., Wen, J., Hu, L., Gu, W., Zhu, C.,
2826–2832. 2023. Angew. Chem. Int. Ed. 62, e202304625.
Bai, Y., Dou, Y., Xie, L.-H., Rutledge, W., Li, J.-R., Zhou, H.-C., 2016. Chem. Soc. Rev. 45 Xu, Y., Gong, H., Ren, H., Fan, X., Li, P., Zhang, T., Chang, K., Wang, T., He, J., 2022b.
(8), 2327–2367. Small 18, 2203917.
Cao, Y., Wu, R., Zhou, Y., Jiang, D., Zhu, W., 2022. Adv. Funct. Mater. 32, 2203005. Yu, J., Park, J., Van Wyk, A., Rumbles, G., Deria, P., 2018. J. Am. Chem. Soc. 140 (33),
Cao, Y., Wang, J.-X., Lin, C., Geng, Y.-Q., Ma, C., Zhu, J.-J., Wang, L., Zhu, W., 2023. 10488–10496.
Adv. Funct. Mater. 33, 2214294. Yu, S., Du, Y., Niu, X., Li, G., Zhu, D., Yu, Q., Zou, G., Ju, H., 2022. Nat. Commun. 13,
Chen, M.-M., Xu, C.-H., Zhao, W., Chen, H.-Y., Xu, J.-J., 2022. Angew. Chem. Int. Ed. 61, 7302.
e202117401. Zhang, R., Cheng, J., Yang, L., Wong, J.M., Ralph Adsetts, J., Wang, R., Liu, J., Ding, Z.,
Chiba, S., 1997. Biosc. Biotech. Biochem. 61 (8), 1233–1239. Wang, H.-B., 2021. Chemelectrochem 8, 547–557.
Cho, Y.K., Kim, H., Bénard, A., Woo, H.-K., Czubayko, F., David, P., Hansen, F.J., Lee, J. Zheng, Y., Yang, H., Zhao, L., Bai, Y., Chen, X., Wu, K., Liu, S., Shen, Y., Zhang, Y., 2022.
I., Park, J.H., Schneck, E., Weber, G.F., Shin, I.-S., Lee, H., 2022. Sci. Adv. 8 (38), Anal. Chem. 94 (7), 3296–3302.
eabq4022. Zhou, J., Li, Y., Wang, W., Tan, X., Lu, Z., Han, H., 2020. Biosens. Bioelectron. 164,
Cong, S., Jiang, Z., Zhang, R., Lv, H., Guo, J., Zhang, L., Lu, X., 2022. Anal. Chem. 94 112332.
(18), 6695–6702. Zhu, D., Zhang, Y., Bao, S., Wang, N., Yu, S., Luo, R., Ma, J., Ju, H., Lei, J., 2021. J. Am.
Da Silva, E.S., Moura, N.M.M., Neves, M.G.P.M.S., Coutinho, A., Prieto, M., Silva, C.G., Chem. Soc. 143 (8), 3049–3053.
Faria, J.L., 2018. Appl. Catal. B Environ. 221, 56–69.
Ding, Z., Quinn, B.M., Haram, S.K., Pell, L.E., Korgel, B.A., Bard, A.J., 2002. Science 296
(5571), 1293–1297.

You might also like