Next Article in Journal
Sensitivity Enhancement of Hybrid Two-Dimensional Nanomaterials-Based Surface Plasmon Resonance Biosensor
Next Article in Special Issue
Sulfonated Starch-Graft-Polyaniline@Graphene Electrically Conductive Nanocomposite: Application for Tyrosinase Immobilization
Previous Article in Journal
G-Optrode Bio-Interfaces for Non-Invasive Optical Cell Stimulation: Design and Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in the Development of Biosensors Based on Peptide–Copper Coordination Interaction

1
College of Chemistry and Chemical Engineering, Anyang Normal University, Anyang 455000, China
2
College of Chemistry and Chemical Engineering, Henan University of Technology, Zhengzhou 450052, China
*
Authors to whom correspondence should be addressed.
Biosensors 2022, 12(10), 809; https://doi.org/10.3390/bios12100809
Submission received: 27 August 2022 / Revised: 21 September 2022 / Accepted: 27 September 2022 / Published: 30 September 2022

Abstract

:
Copper ions, as the active centers of natural enzymes, play an important role in many physiological processes. Copper ion-based catalysts which mimic the activity of enzymes have been widely used in the field of industrial catalysis and sensing devices. As an important class of small biological molecules, peptides have the advantages of easy synthesis, excellent biocompatibility, low toxicity, and good water solubility. The peptide–copper complexes exhibit the characteristics of low molecular weight, high tenability, and unique catalytic and photophysical properties. Biosensors with peptide–copper complexes as the signal probes have promising application prospects in environmental monitoring and biomedical analysis and diagnosis. In this review, we discussed the design and application of fluorescent, colorimetric and electrochemical biosensors based on the peptide–copper coordination interaction.

1. Introduction

Copper is an essential trace metal element for redox chemistry and the growth and development of organic organisms. In the human body, copper is the third largest trace element besides iron and zinc, which plays an important role in human metabolism. Copper ions are involved in the synthesis of many enzymes, such as cytochrome oxidase, peroxide dismutase, tyrosinase, ascorbic acid oxidase, lysyl oxidase, and ceruloplasmin [1]. As the active centers of enzymes, copper ions participate in many electron transfer and redox reactions involving oxygen in various biological systems and life processes. However, they exhibit great toxicity when their content exceeds the inherent steady-state regulation range of organisms [2]. In organisms, copper ions mainly exist in the form of complexes with proteins or peptides, although small molecule-based metal chelators can be used to maintain the concentration of free copper ions below picomolar levels. In peptide–copper complexes, the terminal amino group, deprotonated amide nitrogen and various side chain donors are the binding sites of copper ions [3,4,5,6,7,8,9]. Histidine residues in the peptide sequence, especially in the first three positions of the N-terminal (denoted as His1, His2, and His3), are the main sites for the binding of copper ions. For example, the His1-contained peptide can bind Cu2+ by two nitrogen atoms in the free amine of the N-terminal and the imidazole group of histidine residue [3,9], the His2-contained peptide can bind Cu2+ by three nitrogen atoms in the free amine of N-terminal, the second amide as well as the imidazole group of histidine residue [6], and the His3-contained peptide, also denoted as the amino-terminal copper and nickel (ATCUN) binding peptide, can chelate Cu2+ and Ni2+ ions with femtomolar affinity. Cu2+ is coordinated by ATCUN peptide through the N-terminal free amine, the second and third amides, and the imidazole group of histidine residue [6,7,8,9]. In general, the location of histidine in the peptide sequence has a decisive impact on the characteristics of peptide–copper complexes, including the coordination mode, binding affinity, redox potential, and catalytic activity. In addition, divalent copper ions are paramagnetic and easy to combine with electron chelating groups with appropriate cavity volume. This will cause the change of electron or energy transfer mode when the organic molecules are excited, resulting in the reduction or quenching of the fluorescence signal [10,11,12,13,14]. The design of biosensors based on the intrinsic fluorescence quenching ability and the redox activity of copper ions has been widely covered by researchers.
Previously, several review papers have overviewed the design of peptide-derived biosensors by using peptides as the recognition elements or enzyme substrates [15,16,17,18,19]. However, there is little overview of the development of biosensors based on the formation of peptide–copper complexes. In this work, we, for the first time, review the progress in the design of fluorescent, colorimetric, and electrochemical biosensors based on the peptide–copper coordination interaction (Scheme 1). In view of the fluorescence quenching of Cu2+ by electron or energy transfer, dye-labeled peptide monomers or their assemblies have been designed to develop fluorescent biosensors for the sensitive detection of metal ions and proteases. Due to the specific coordination mode, copper ions can regulate the aggregation of plasmonic metal nanoparticles alongside with the change of color solution. Meanwhile, the peptide–copper complexes with catalytic properties can mediate the in situ growth of metal nanoparticles and accelerate the chromogenic reaction, which facilitate the design of colorimetric biosensors. In electrochemical biosensors, the peptides can be modified on the electrode surface to determine copper ions. More interestingly, the peptide–copper complexes with electrocatalytic ability could be used as the signal tags for the design of novel electrochemical biosensors.

2. Fluorescent Biosensors

2.1. Peptide Fluorescent Probes for Copper Ions

Traditional methods for the detection of metal ions mainly include atomic spectrometry, inductively coupled plasma, chromatography, and mass spectrometry. Although these methods can achieve accurate determinations, they require the use of large and expensive instruments with complex operation procedures. As an important class of homogeneous detection methods, fluorescent probes have widely used in the field of bioassays in vitro and in vivo. Peptides have been widely used to develop various biosensors due to their characteristics of easy synthesis, good biocompatibility, and low toxicity [15,17,18]. Moreover, their rich structural composition and variable structural characteristics are favorable to designing sequence-specific peptides for the selective coordination with various analytes, including metal ions. Based on the type of fluorescent groups, peptide-based fluorescent biosensors can be divided into endogenous fluorophores-based and exogenous fluorophores-labeled biosensors. Several amino acids, such as tryptophan (Trp), tyrosine (Tyr), and phenylalanine (Phe), exhibit endogenous fluorescence [20,21,22,23,24]. The peptide of HDSGWEVHH has been used as a label-free fluorescent probe for the detection of Cu2+ and S2− [23]. However, the fluorescence intensity of endogenous fluorophores is low and the sensitivity and selectivity of such methods are poor for the detection of copper ions. Therefore, exogenous fluorescent molecules or metal nanoclusters (NCs) are often used to modify peptides, including dansyl (Dns), carboxyl fluorescein (FAM), fluorescein isothiocyanate (FITC), and pyrene sulfonyl chloride (PySO2) (Table 1). Among them, dansyl is a commonly used fluorescent group to modify peptide for the detection of metal ions [25,26,27,28,29,30,31,32,33,34]. Meanwhile, the sulfonamide group in it can facilitate the formation of peptide–metal complexes [35]. Regarding this, Wang et al. reported a peptide fluorescent probe (dansyl-Ser-Pro-Gly-His-NH2) for the detection of Cu2+ in 100% aqueous solution [29]. Histidine (His) and serine (Ser) separated by the spacer (Pro-Gly) are the binding sites to coordinate with Cu2+. After being complexed by the peptide, Cu2+, with a paramagnetic property, quenched the fluorescence of the dansyl group near the binding center via electron or energy transfer from the excited fluorophore to the coordinated Cu2+. In addition, metal NCs as the fluorescent reporters have attracted extensive attention because of their promising properties (e.g., easy synthesis, good biocompatibility, and high stability) [36,37,38]. In order to evaluate the selectivity and anti-interference of the sensor, other common metal ions (K+, Na+, Mg2+, Fe2+, Cu+, Co2+) and plasma were determined as the control experiments. The results indicated that these species showed negligible interference for Cu2+ detection. Finally, tap water was analyzed as a real sample with excellent results. Qian et al. developed a Cu2+ biosensor by using the peptide of lipoic acid-Arg-Gly-Asp-Asp-Lys-lipoic acid (LA-RGDK-LA) as the templet and ligand to synthesize AuNCs [37]. Zhuang et al. reported the detection of Cu2+ with the peptide (CCYWDAHRDY)-modified AuNCs, in which bovine serum albumin (BSA) was used as the template for the synthesis of fluorescent AuNCs [38]. The binding of Cu2+ ions to the peptide-modified AuNCs caused the fluorescence quenching.
Although many peptide-based sensors for Cu2+ detection have been developed, most of them work in the “signal-off” formation and suffer from the intrinsic fluorescence background of biomolecules in biological media. To solve this problem, Falcone et al. designed two Tb3+-based luminescent ATCUN peptides for reversible and “turn-on” detection of Cu2+ in biological media [39,40]. The ATCUN motif in the probe exhibited high affinity toward Cu2+. The Tb3+-DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid) complex was sensitized to the Trp residue via the “antenna effect”, providing a long-lifetime luminescence. After the formation of ATCUN–Cu2+ complex, the luminescence of both Trp and Tb3+ was quenched. In addition, Jung et al. developed a dansyl-labeled tripeptide fluorescent probe (DNS-LLC) to detect Cu+ for the first time [41]. This probe formed a 1:1 complex with Cu+ in aqueous solution and showed a “signal-on” response to Cu+. The position of cysteine residue in the peptide chain showed a significant effect on the selectivity. The DNS-LLC could penetrate live RKO cells and monitor Cu+ in Golgi matrix with a “signal-on” fluorescent response.
Multiple metal ions can also be determined by one peptide probe by reasonable design of the peptide sequence. For example, Pang et al. reported that dansyl-His-Pro-Gly-Trp-NH2 with a dansyl fluorophore and a Trp residue exhibited a different response toward Cu2+ and Hg2+ [27]. After the coordination of peptide with Hg2+, the distance between the Trp residue (donor) and the dansyl fluorophore (acceptor) was shortened, increasing the efficiency of the fluorescent resonance energy transfer (FRET). As a result, the fluorescence intensity of Trp at 355 nm decreased and that of dansyl at 550 nm increased with a 45 nm blue shift. The fluorescent probe showed good selectivity and sensitivity toward Hg2+ and Cu2+ among 17 metal ions and 6 anions in N-2-hydroxyethylpiperazine N-2-ethane acid (HEPES) buffer solution. Moreover, Wang et al. synthesized a peptide fluorescent probe (dansyl-Gly-Cys-NH2, DGC) for the selective detection of Cu2+ and Cd2+ in aqueous solution and within live LNCaP cells [32]. At the same time, they designed and synthesized two dansyl-labeled fluorescent peptide probes (dansyl-Ser-Pro-Gly-His-NH2 and dansyl-His-Pro-Gly-Glu-NH2) for the “signal-on” detection of Zn2+ and “signal-off” detection of Cu2+ through two different mechanisms [30,33]. The detection limits for Zn2+ and Cu2+ are as low as 4.9 nM and 15 nM, respectively.

2.2. Peptide-Cu2+ Complexes for Fluorescent Detection of Anions

The formed peptide–Cu2+ complexes can be further used to detect other substances that can bind to Cu2+ with high binding constant. Thus, peptides can be employed as the dual-functional probes to develop “on–off–on” biosensors for multiple analytes [28,29,34,42,43,44,45,46,47,48,49]. For example, anions, such as S2− and CN, can sequester Cu2+ from the peptide–Cu2+ complexes, leading to the restoration of the quenched fluorescence. Recently, Hu et al. designed a dansyl-labeled hexapeptide for the detection of Cu2+ and S2− [34]. The peptide formed a 1:1 complex selectively with Cu2+ without interference by other metal ions, resulting in the fluorescence quenching. After the addition of S2−, Cu2+ was released from the complex and the fluorescence was recovered. Finally, the dansyl-labeled peptide was successfully utilized for the fluorescence imaging of Cu2+ and S2− in living HeLa cells and zebrafish. Wang’s group used a series of fluorophore-labeled peptides to detect Cu2+ and S2− or H2S by both colorimetric and fluorescent methods [29,44,45,46]. For instance, FITC-Ahx-Ser-His-NH2 was synthesized to continuously monitor Cu2+ and S2− in buffer solution and within living cells [45]. The probe showed high sensitivity for colorimetric analysis of Cu2+, and the color change from yellow to orange was observed with naked eye, reflecting the fluorescence “signal-off” response to Cu2+. Then, the peptide-Cu2+ complex was used for colorimetric and fluorescent analysis of S2−. The system could continuously and alternately determine Cu2+ and S2− with good reversibility. Moreover, Hao et al. designed and synthesized a fluorescent probe of FITC-Ahx-His-Glu-Phe-Cys-NH2 [47]. Histidine and cysteine residues in the peptide chain acted as the Cu2+-binding groups. Through photon-induced electron transfer (PET), the coordination between Cu2+ and peptide caused the electron transfer, which led to the fluorescence quenching and realized the specific detection of Cu2+. The addition of S2− caused the removal of Cu2+ from the system and led to the fluorescence recovery. By mimic the binding mode of the ceruloplasmin for Cu2+, Wang et al. reported a peptide-based multifunctional probe (FITC-Ahx-Gly-His-Lys-NH2) for the fluorescence detection of Cu2+ and H2S in live colon cancer cells [46]. Other metals did not interfere with the ability of the probe to detect Cu2+ ions. In addition, five common soluble copper salts were tested for anionic interference. The results showed that the anions had no obvious interference with Cu2+ ions. Meanwhile, they synthesized a FITC-labeled peptide ligand (FITC-Ahx-Ser-Pro-Gly-His-NH2), in which His and Ser residues were the coordination sites for Cu2+ [44]. The peptide–Cu2+ complex was further used to detect H2S in living cells and zebrafish larvae. Taking advantage of the capability of histidine residue to coordinate with Zn2+ and Cu2+, Wang et al. synthesized a peptide probe (dansyl-His-Lys-His-dansyl) with lysine and histidine residues as the main chain for the detection of multiple analytes in live cells [28]. The binding of histidine residue and Zn2+ inhibited the photo-induced electron transfer (PET) from the indole moiety to the dansyl fluorophore, generating a “turn-on” fluorescence response. Conversely, the complexation of Cu2+ quenched the fluorescence and resulted in a “turn-off” response. The detection of S2− was further achieved through the formation and separation of CuS. The probe showed good cell permeability, low biotoxicity, and high selectivity as well as sensitivity.
Cyanide can impair the cellular functions of metalloenzymes by binding with the metal cofactors. Thus, cyanide is considered to be one of the most toxic anions, even at a very low concentration. Jung et al. designed a peptide-based system for the detection of cyanide in aqueous solution based on the ATCUN motif, in which the tripeptide Ser-Ser-His (NBD-SSH) was labeled by 7-nitro-2,1,3-benzoxadiazole at the N-terminal (Figure 1) [48]. The formation of the NBD-SSH–Cu2+ complex resulted in the color change of the probe from yellow to orange, which was accompanied by the fluorescence quenching. The addition of CN successfully removed Cu2+ from the complex. As a result, the solution color turned yellow, and the fluorescence signal was enhanced. This method realized the highly sensitive detection of CN through the colorimetric and fluorescent analysis. This method is the first example of using a peptide-integrated system to sensitively detect cyanide in 100% aqueous solution. Other anions, such as F, Cl, Br, I, NO3−, SCN, AcO, HCO3−, SO42−, ClO4−, HPO42−, and HAsO42−, showed no significant effect on the colorimetric detection of cyanide. Recently, Wang et al. reported a bifunctional peptide probe (FLH) based on fluorescence “on-off” response and colorimetric analysis [43]. In this method, FLH could detect Cu2+ quickly and selectively with a detection limit as low as 1.5 nM, and the formed FLH-Cu complex showed a colorimetric change and a “signal-on” response to CN with a detection limit of 12.7 nM. Interference experiments showed that other ions did not induce the change of solution color from light yellow to orange under natural light, and the effect of other ions on the fluorescence intensity was negligible. Finally, the probe was successfully applied to monitor Cu2+ and CN in living cells through fluorescence imaging.

2.3. Other Fluorescent Biosensors Based on Peptide-Cu2+ Complexes

Based on the high binding affinity of ATCUN peptide and Cu2+ and the high inherent quenching ability of Cu2+ toward the fluorophore, Deng et al. proposed a single-labeled peptide probe for the detection of protease (Figure 2) [50]. In this work, two proteases, caspase-3 and β-secretase, were determined in a “signal-on” and “signal-off” format, respectively. For the detection of caspase-3, the Dns group was labeled at the C-terminal of peptide. Cu2+ could be coordinated by the peptide probe SGHDEVDK-Dns through the interaction with the ATCUN motif (SGH) at the N-terminal, thus leading to the fluorescence quenching. Once the peptide was catalytically cleaved by caspase-3, the released fragment of K-Dns exhibited high fluorescence signal even in the presence of Cu2+. For the assay of β-secretase, the fluorophore was labeled at the N-terminal of peptide. The probe EVNLDAHFWADK-Dns had a low binding affinity toward Cu2+ and showed a high fluorescence signal. Once the probe was cleaved by β-secretase, Cu2+ would be coordinated by the ATCUN moiety (DAH) in the released DAHFWADK-Dns segment, thus leading to the fluorescence quenching.
Conventional small-molecule organic dyes may confront the problem of aggregation-caused quenching and photobleaching. To overcoming this disadvantage, organic luminaries with aggregation-induced luminescence (AIE) characteristics have attracted extensive attention in recent years [51,52]. In the dispersed state of AIE molecules, the vibration and rotation inside the molecules lead to the rapid non-radiative attenuation of the excited state. Thus, the fluorescence of AIE molecules in dilute solution is very weak. However, in the aggregated state, the non-radiative attenuation of restriction of intra-molecular rotation (RIR) is suppressed, resulting in the strong fluorescence. Liu et al. reported an AIE-based probe (TPE-Py-EEGTIGYG) for the highly selective detection of Cu2+ in aqueous solution and living cells [53]. The peptide of EEGTIGYG exhibited good water solubility and specific binding ability toward cell membrane. The TPE-Py-EEGTIGYG monomers could self-assemble into aggregates at a high concentration (25 μM) in aqueous solution and thus produce a strong fluorescence signal. Cu2+ could be selectively and sensitively determined by quenching the fluorescence of nanoaggregates. Interestingly, TPE-Py-EEGTIGYG at a low concentration (5 μM) exhibited weak fluorescence intensity in aqueous solution. When it was anchored to the cell membrane, the fluorescence was significantly enhanced. Thus, the probe can be used to determine Cu2+ entering and leaving the cell. Moreover, Gour et al. developed a pyridothiazole-based AIE probe for the in vitro detection of amyloid peptide [54]. The probe of 4-(5-methoxythiazolo [4,5-b] pyridin-2-yl) benzoic acid (PTC1) shows AIE characteristics. However, in the presence of Cu2+, PTC1 molecules were dissociated or destroyed, resulting in the fluorescence quenching. The coordination of Cu2+ by amyloid peptide caused the significant recovery of fluorescence signal. Cai et al. reported the detection of Cu2+ and caspase-3 based on the self-assembly of 4-nitro-2,1,3-benzoxadiazole (NBD)-labeled peptides with the sequence of NBD-FFYEEGGH and NBD-FFFDEVDGGH (Figure 3) [55]. The peptides could self-assemble into nanofibers with enhanced cellular uptake and brighter fluorescence. The coordination of Cu2+ to the nanofibers by the interaction between Cu2+ and GGH segment caused the fluorescence quenching, which allowed for the determination of Cu2+ in cells. Furthermore, the NBD-FFFDEVDGGH–Cu2+ nanofibers were used to determine caspase-3 activity in vitro and within cells. Upon the enzymatic cleavage of the peptide between NBD-FFFDEVD and GGH, the Cu2+–GGH complex was removed, resulting in the fluorescence recovery. Moreover, Kim et al. prepared two types of pyrene-labeled peptide nanofibrils for “light-off” and “light-up” fluorescent detection of Cu2+ and Ag+ in HeLa cells, respectively [56]. The peptides consist of a hydrophobic linear- or branched-alkyl chain and a hydrophilic histidine-rich peptide (HGGGHGHGGGHG). The nanofibrils were further used as the template scaffolds for the preparation of monodispersed and highly stable silver nanoparticles (AgNPs).

3. Colorimetric Biosensors

3.1. Aggregation and Growth of Metal Nanoparticles Mediated by Peptide–Cu2+ Complexes

Colorimetric biosensors with apparent color changes are particularly attractive because of their high sensitivity, minimum instrumental investment and visual signal-readout [57]. Gold nanoparticles (AuNPs) are the commonly used indicators for colorimetric assays due to their good stability and high extinction coefficient. The dispersed AuNPs with small diameters usually exhibit wine red in the solution. However, when AuNPs are subjected to aggregation, the color changes from red to blue due to the transfer of surface plasmon absorption to a longer wavelength. Peptides and proteins can be rapidly adsorbed or assembled on the surface of AuNPs to trigger or prevent the aggregation of AuNPs. Based on the coordination of Cu2+ with peptide, AuNPs-based colorimetric methods have been used for the quantification of peptide and Cu2+ (Table 2) [58,59,60,61]. For example, Wang et al. investigated the interaction between amyloid-β peptides and Cu2+/Zn2+ ions using AuNPs as the indicators [59]. Zhou et al. reported the AuNPs-based colorimetric assay of Aβ(1–40) (one of the biomarkers of Alzheimer’s disease) by the interaction between Aβ(1–40) and Cu2+, in which the coordination of Cu2+ with Aβ(1–40) induced the aggregation and color change of the Aβ(1–40)-AuNPs conjugates (Figure 4) [60]. The detection limit of this method for Aβ(1–40) was found to be 0.6 nM. Moreover, Pelin et al. found that Cu2+ could induce the aggregation of amphiphile peptide-modified AuNPs, thus achieving the colorimetric detection of Cu2+ in aqueous solution with a detection limit of 0.19 μM [61].
In contrast to the AuNPs-based assays, AgNPs-based colorimetric analysis exhibits the advantages of lower cost and higher extinction coefficient. For this view, Ghodake et al. reported the colorimetric detection of Cu2+ by using casein peptide-functionalized AgNPs as the indicators [62]. Hu et al. developed a colorimetric immunosensor for the specific detection of Aβ(1–40/1–42) by using antibody-modified AgNPs (Ab-AgNPs) as the indicators and Cu2+ ions as the linkers [63]. The coordination between Cu2+ and Aβ(1–40/1–42) captured by the Ab-AgNPs induced the aggregation of AgNPs, which was accompanied by the color change from yellow to red. The method allowed for the determination of Aβ(1–40/1–42) with a detection limit down to 86 pM.
Peptides with cysteine and positively charged amino acid residues can induce the aggregation of AuNPs through the Au–S and electrostatic interactions. The cleavage of the peptide by protease may prevent the aggregation of AuNPs, thus facilitating the colorimetric detection of protease [64]. However, in the aggregation-based colorimetric assays, the peptide substrate must be well-designed for interacting with gold surface. Recently, Liu and coworkers reported a colorimetric method for protease detection based on the in-situ growth of AuNPs and the peptide-mediated catalytic activity of Cu2+ (Figure 5) [65]. In this method, AuNPs were generated by the reduction of chloroauric acid with AA as the reducing reagent, but the Cu2+-catalytic oxidation of AA limited the formation of AuNPs. Because the peptide substrate and the proteolytic products exhibited different binding affinity toward Cu2+, the oxidation of AA and the formation of AuNPs were mediated by proteolytic reaction. Thus, the color change was dependent upon the concentration and activity of protease. To exclude the interference from matrix components in biological fluids, the peptide substrates were modified on the surface of magnetic microbeads. The cleavage of the peptides on the surface of magnetic microbeads by protease led to the generation or detachment of Cu2+-binding ATCUN peptide segments. This sensing strategy is fundamentally different from the aggregation-based colorimetric assays, thus providing a new concept for the design of protease biosensors.

3.2. Peptide-Cu2+ Self-Assembles as the Signal Tags

Phenylalanine residues in the peptide chains can promote the self-assembly of peptides into various molecular conformations via hydrogen-bonding and π-stacking interactions [66,67]. The histidine residues play an important role in the coordination of divalent metal ions as the active centers of many natural enzymes [1,68]. Thus, incorporation of metal ions with peptides may grant for assembles with catalytic activity or new functionality. Sun et al. prepared the biotinylated peptide-Cu2+ nanoparticles (Cu-P NPs) by “one-pot” self-assembly with biotin-FFKGH as the precursor (Figure 6) [69]. The Cu–P NPs contained extensive active center of Cu2+ ions. Under acidic conditions, Cu2+ ions with intrinsic peroxidase activity could be released to promote the catalytic oxidization of 3,3′,5,5′-tetramethylbenzidine (TMB). The Cu–PNPs were further used as the labels for the colorimetric immunoassays of prostate specific antigen (PSA). The target PSA attached on the ELISA was recognized by the streptavidin-conjugated secondary antibody (Ab2–SA), allowing for the capture of Cu–PNPs by the streptavidin–biotin interaction. Then, the Cu2+ ions were released to catalyze the oxidation of the colorless TMB into blue oxTMB in the presence of H2O2. The biosensor has a linear range of 0.001–1 ng/mL with a detection limit of 1 pg/mL. In addition, Qi and coworkers synthesized a laccase-like nanozyme by using Cu2+ and dipeptide Cys-His (CH) as the precursors [70]. The CH–Cu nanozyme formed thorough the coordination interaction between Cu2+/Cu+ and CH exhibited high catalytic activity and thermostability even at extreme pH and high temperature. It has been successfully used for the degradation of chlorophenol and bisphenol and the determination of epinephrine by a smart phone. Zhang et al. prepared various self-assembles with histidine-containing peptides as the precursors [71]. The peptide conformation changed from random coil to β-sheet structure and nanotube with hydrolase activity. After integration with Cu2+, the morphology of self-assembles was altered and the resulting hybrid nanomaterials showed excellent peroxidase-like activity. These works suggest that artificial enzymes could be designed with histidine-containing self-assemble peptides by incorporation of copper ions.

4. Electrochemical Biosensors

4.1. Peptide-Modified Electrodes for Copper Detection

A variety of optical, electrochemical, and mass biosensors have been developed for the detection of metal ions with peptides as the probes. Among them, electrochemical biosensors have the characteristics of easy miniaturization, rapid response, and high sensitivity. Therefore, it is of great significance to design electrochemical biosensors for real-time detection of copper ions. Compared with other biological macromolecules, peptides have the characteristics of small molecular weight, are easily modified, and have a simple synthesis procedure. The peptide-modified electrodes have been prepared to detect copper ions, including the simplest ATCUN tripeptide Gly-Gly-His (GGH) [72,73,74,75,76,77]. For example, Yang et al. prepared an electrochemical biosensor for rapid detection of Cu2+ by modification of tripeptide GGH onto 3-mercaptopropionic acid (MPA)-modified gold electrode via carbodiimide-mediated amine coupling reaction [72]. The redox behavior of GGH–Cu2+ complexes on the electrode surface has been investigated by cyclic voltammetry. In order to test the performance of the modified electrode for the analysis of real samples, copper ions in diluted tap water and lake samples were determined. The concentration of Cu2+ in tap water was 0.27 ppm, which was consistent with that (0.33 ppm) determined by inductively coupled plasma mass spectrometry (ICP-MS). Subsequently, they reported the direct synthesis of tripeptide GGH on the gold electrode surface for the binding of Cu2+ [73]. The covalent coupling of Gly, Gly, and His through three steps on the MPA self-assembled monolayers (SAMs)-covered electrode surface was confirmed by characterizing the electrochemically desorbed peptide with mass spectrometry. Compared with previous reports, the direct synthesis of GGH on the electrode surface effectively provided the recognition site of Cu2+ and thus improved the detection sensitivity. However, the SAMs formed through the Au–S covalent bonds on the gold surface may be affected by temperature, thus reducing the stability of biosensors. In order to solve this problem, Gooding’s group used aryldiazonium salt to modify the glassy carbon electrode and micro planar electrode for the immobilization of Cu2+-binding peptide [74,75]. In addition, the strategy of covalent immobilization of peptides through silicon–carbon bonds has attracted much attention. Sam et al. prepared a Cu2+ biosensor by covalent modification of tripeptide GGH on the surface of porous silicon electrode through the silicon-carbon interaction [76]. These methods solved the problem of instability of SAMs on the gold surface, making the commercialization of Cu2+ biosensors possible.
The binding of Cu2+ with peptide on the gold surface may induce the change in the peptide conformation and electron conductance of electrode interface [78]. For example, Chen et al. used surface plasmon resonance (SPR) spectroscopy to study the conformational change of Cu2+-binding peptide of Ac-CGGGSIRKLEYEIEELRLRIG-NH2 [79]. After the injection of Cu2+ onto the peptide-modified gold chip, the SPR angle increased significantly, indicating that Cu2+ was successfully captured by the peptide. Electrochemical impedance spectroscopy (EIS) can measure the electrical resistance of the metal/solution interface and is capable of delivering measurable signal change resulting from the difference in analyte concentration. This technique has been used to achieve the highly sensitive detection of Cu2+ ions with peptide-modified electrodes. For example, Jiang et al. developed an EIS biosensor for Cu2+ detection with GGH-covered AgNPs-modified glassy carbon electrode [80]. In the absence of Cu2+, the charge transfer resistance (Rct) between the peptide-modified electrode and the [Fe(CN)6]3−/4− probe is high. However, the value decreased significantly after Cu2+ ions were captured by the sensing electrode. In addition, Mervinetsky et al. prepared a Cu2+ biosensor based on the impedance change by covalently coupling of lipoic acid-functionalized tripeptide GGH to the gold electrode [81]. With the increase of Cu2+ concentration, the conformation of peptide on the electrode surface changed and the SAMs became more and more dense, thus resulting in the significant change of Rct value.

4.2. Electrochemical Biosensors with Peptide-Cu2+ Complexes as the Signal Reporters

ATCUN peptide is characterized by a free amine group at the N-terminus and a histidine residue in the third position. Cu2+ was coordinated by the free amine, the second and third amides, and the imidazole group of histidine residue [6,7,8,9]. In actual fact, when the ATCUN peptide was anchored by the covalent coupling reaction between the free N-terminus amine group and the activated carboxyl group on the electrode surface, the binding mode and redox activity of the immobilized peptide–Cu2+ complex might be essentially different from the ATCUN–Cu2+ complex [82,83]. Interestingly, Deng et al. investigated the redox activity of various ATCUN–Cu2+ complexes and found that they exhibited an electrocatalytic ability for water oxidation even at neural pH and low oxidation potential (Figure 7) [84]. However, the acetylated ATCUN peptide did not significantly change the redox property of Cu2+. Based on this fact, nanocatalysts were prepared by modifying the ATCUN–Cu2+ complexes on the surface of AuNPs by the Au–S bonds and then used as the signal labels to develop a sandwich-like biosensor for the detection of DNA. The analytical performances were shown in Table 3. This is the first electrochemical biosensor by using H2O molecule as the catalytic substrate. Furthermore, Deng et al. developed a “signal-on” electrochemical biosensor for the detection of caspase-3 activity by the formation of ATCUN peptide on the electrode surface [85]. In this work, the peptide substrates specific to caspase-3 were anchored on the graphene electrode surface through the hydrophobic and π-stacking interactions. The enzymatic hydrolysis of peptide substrates led to the generation of ATCUN motif. The formed ATCUN–Cu2+ complexes on the electrode surface exhibited high electrocatalytic ability for water oxidation. As a result, caspase-3 has been determined at the concentration range of 0.5 pg/mL~2 ng/mL with a detection limit of 0.2 pg/mL. The biosensor was further used to evaluate drug-induced cell apoptosis with satisfactory results. Meanwhile, Liu’s group developed an electrochemical immunosensor by using protease as the signal reporter to catalyze the generation of Cu2+-binding ATCUN peptide [86]. In this work, carbon nanotubes (CNTs) were used as the carriers to load the recognition antibody and protease for signal amplification. Protease on the surface of CNTs promoted the generation of ATCUN peptides by catalyzing the cleavage of the substrates. The resulting ATCUN–Cu2+ complexes showed good electrocatalytic property toward water oxidation, thus leading to the second signal amplification through the redox cycling reaction.
Recently, we studied the redox activity of the Cu2+ complexes formed with the His1- and His2-containing peptides [87]. It was found that the Cu2+ complexes formed with His1-containing peptides exhibited excellent catalytic activities for the electrochemical reduction of oxygen and chemical oxidation of ascorbic acid (AA) by O2. However, the His2-containing peptides were able to coordinate with Cu2+ ions and depress the catalytic activity of Cu2+. Finally, the Cu2+ complexes formed with His1-containing peptides were modified on the surface of AuNPs to produce nanocatalysts for oxygen reduction or AA oxidation. The resulting nanocatalysts were further used as the signal labels for the design of electrochemical and fluorescent biosensors with PSA as the model target (Figure 8). In the electrochemical immunoassays, the nanocatalysts promoted the electrocatalytic reduction of oxygen. In the fluorescent immunoassays, the nanocatalysts catalyzed the oxidation of AA, and the oxidization product of dehydroascorbic acid (DHA) was then reacted with o-phenylenediamine (OPD) to produce fluorescent 3-(dihydroxyethyl)furo [3,4-b]quinoxaline-1-one (DFQ). The electrochemical and fluorescent methods show a detection limit of 0.40 and 1.00 pg/mL, respectively. Meanwhile, it was found that the Cu2+ complexes performed with the His1-containing peptides reveal high activity for the catalytic oxidation of OPD to produce fluorescent 2,3-diaminophenazine (OPDox) and maintained excellent selectivity for PSA detection even in the presence of high concentration of interfering proteins [88]. Based on the difference in the redox potential and current of Cu2+ before and after being coordinated by the His1- or His3-containing peptide, Feng et al. proposed a general, label-free, and homogeneous electrochemical biosensor for the detection of protease activity [89]. The angiotensin-converting enzyme (ACE) and thrombin were tested as the analyte examples. In this work, ACE catalyzed the hydrolysis of peptide substrates to produce His1-containing peptides and thrombin promoted the generation of His3-containing peptides. The resulting peptide–Cu2+ complexes exhibited a couple of redox peaks, which are essentially different from the free Cu2+ ions or the substrate–Cu2+ complexes.

5. Conclusions

In summary, this work provided an overview on the advancement of different biosensors based on the peptide–copper complexes, including fluorescence, colorimetric, and electrochemical methods. Because of their advantages of easy preparation, high binding affinity, and good biocompatibility, the peptide–copper complexes have been broadly employed in diverse fields of environment monitoring, food safety, and disease diagnosis. For example, copper ions can be sensitively detected by fluorescent peptide probes through quenching the endogenous fluorescence of amino acid residues or the exogenous fluorescence of dyes labels due to its paramagnetic capability. The peptide–copper complexes with excellent catalytic and redox properties have been elaborately used as signal labels to construct colorimetric and electrochemical sensing platforms.
Although some important progress has been made, there are still some problems to be further explored. For example, novel copper complexes with more enzyme-mimicking abilities (e.g., laccases and glucose oxidase) are desired and the efficiency and specificity toward substrates should be further improved. The binding and catalytic mechanism of peptide–copper complexes need to be further studied under uniform standards and experimental conditions, because experimental parameters have a significant influence on the redox properties, including pH, the ratio of copper ions and peptide, and scan rate in electrochemical techniques. Moreover, new detection principles and technologies should be developed and the commercialization of sensing devices needs to be promoted. It is believed that more and more peptide-based biosensors will be discovered with the continuous exploration and exploration of metal complexes.

Author Contributions

Conceptualization, N.X.; investigation, L.T.; writing—original draft preparation, G.L. and N.X.; writing—review and editing, Z.S. and L.L.; project administration, N.X.; funding acquisition, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Program for Innovative Research Team of Science and Technology in the University of Henan Province (21IRTSTHN005) and the National Natural Science Foundation of China (21905002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tisato, F.; Marzano, C.; Porchia, M.; Pellei, M.; Santini, C. Copper in diseases and treatments, and copper-based anticancer strategies. Med. Res. Rev. 2010, 30, 708. [Google Scholar] [CrossRef]
  2. Kozlowski, H.; Potocki, S.; Remelli, M.; Rowinska-Zyrek, M.; Valensin, D. Specific metal ion binding sites in unstructured regions of proteins. Coordin. Chem. Rev. 2013, 257, 2625. [Google Scholar] [CrossRef]
  3. Sóvágó, I.; Várnagy, K.; Lihi, N.; Grenács, Á. Coordinating properties of peptides containing histidyl residues. Coordin. Chem. Rev. 2016, 327, 43. [Google Scholar] [CrossRef]
  4. Weng, Y.C.; Fan, F.-R.F.; Bard, A.J. Combinatorial biomimetics. Optimization of a composition of copper(II) poly-L-histidine complex as an electrocatalyst for O2 reduction by scanning electrochemical microscopy. J. Am. Cheml. Soc. 2005, 127, 17576. [Google Scholar] [CrossRef]
  5. Nguyen, L.T.; Ho, W.F.; Yang, K.-L. Copper–tripeptides (cuzymes) with peroxidase-mimetic activity. RSC Adv. 2020, 10, 17408. [Google Scholar] [CrossRef]
  6. Gonzalez, P.; Bossak, K.; Stefaniak, E.; Hureau, C.; Raibaut, L.; Bal, W.; Faller, P. N-Terminal Cu-binding motifs (Xxx-Zzz-His, Xxx-His) and their derivatives: Chemistry, biology and medicinal applications. Chem. Eur. J. 2018, 24, 8029. [Google Scholar] [CrossRef]
  7. Portelinha, J.; Duay, S.S.; Yu, S.I.; Heilemann, K.; Libardo, M.D.J.; Juliano, S.A.; Klassen, J.L.; Angeles-Boza, A.M. Antimicrobial peptides and copper(II) ions: Novel therapeutic opportunities. Chem. Rev. 2021, 121, 2648. [Google Scholar] [CrossRef]
  8. Harford, C.; Sarkar, B. Amino terminal Cu(II)- and Ni(II)-binding (ATCUN) motif of proteins and peptides: Metal binding, DNA cleavage, and other properties. Acc. Chem. Res. 1997, 30, 123–130. [Google Scholar] [CrossRef]
  9. Frączyk, T. Cu(II)-Binding N-terminal sequences of human proteins. Chem. Biodivers. 2021, 18, e210004. [Google Scholar] [CrossRef]
  10. Deng, D.; Hao, Y.; Xue, J.; Liu, X.; Xu, X.; Liu, L. A colorimetric enzyme-linked immunosorbent assay with CuO nanoparticles as signal labels based on the growth of gold nanoparticles in situ. Nanomaterials 2019, 9, 4. [Google Scholar] [CrossRef] [Green Version]
  11. Brunner, J.; Kraemer, R. Copper(II)-quenched oligonucleotide probes for fluorescent DNA sensing. J. Am. Chem. Soc. 2004, 126, 13626. [Google Scholar] [CrossRef]
  12. Moro, A.J.; Santos, M.; Outis, M.; Mateus, P.; Pereira, P.M. Selective coordination of Cu2+ and subsequent anion detection based on a naphthalimide-triazine-(DPA)2 chemosensor. Biosensors 2020, 10, 129. [Google Scholar] [CrossRef]
  13. Monson, C.F.; Cong, X.; Robison, A.D.; Pace, H.P.; Liu, C.; Poyton, M.F.; Cremer, P.S. Phosphatidylserine reversibly binds Cu2+ with extremely high affinity. J. Am. Chem. Soc. 2012, 134, 7773–7779. [Google Scholar] [CrossRef]
  14. Sasakura, K.; Hanaoka, K.; Shibuya, N.; Mikami, Y.; Kimura, Y.; Komatsu, T.; Ueno, T.; Terai, T.; Kimura, H.; Nagano, T. Development of a highly selective fluorescence probe for hydrogen sulfide. J. Am. Chem. Soc. 2011, 133, 18003. [Google Scholar] [CrossRef]
  15. Wang, M.; Li, L.; Zhang, L.; Zhao, J.; Jiang, Z.; Wang, W. Peptide-derived biosensors and their applications in tumor immunology-related detection. Anal. Chem. 2022, 94, 431–441. [Google Scholar] [CrossRef]
  16. Zhang, H.; Ding, J.; Qin, W. Recent advances in peptide-based electrochemical biosensor. Prog. Chem. 2021, 33, 1756. [Google Scholar]
  17. Saw, P.E.; Xu, X.; Kim, S.; Jon, S. Biomedical applications of a novel class of high-affinity peptides. Acc. Chem. Res. 2021, 54, 3576–3592. [Google Scholar] [CrossRef]
  18. Sfragano, P.S.; Moro, G.; Polo, F.; Palchetti, I. The role of peptides in the sesign of electrochemical biosensors for clinical diagnostics. Biosensors 2021, 11, 246. [Google Scholar] [CrossRef]
  19. Vanova, V.; Mitrevska, K.; Milosavljevic, V.; Hynek, D.; Richtera, L.; Adam, V. Peptide-based electrochemical biosensors utilized for protein detection. Biosens. Bioelectron. 2021, 180, 113087. [Google Scholar] [CrossRef]
  20. Alies, B.; Renaglia, E.; Rozga, M.; Bal, W.; Faller, P.; Hureau, C. Cu(II) affinity for the Alzheimer’s peptide: Tyrosine fluorescence studies revisited. Anal. Chem. 2013, 85, 1501. [Google Scholar] [CrossRef]
  21. Yang, T.; Zhu, F.; Zhou, T.; Cao, J.; Xie, Y.; Zhang, M.; Wang, Y.; Cao, D.S.; Lin, Q.; Zhang, L. Label-free, water-soluble fluorescent peptide probe for a sensitive and selective determination of copper ions. Anal. Sci. 2017, 33, 191. [Google Scholar] [CrossRef] [Green Version]
  22. Makowska, J.; Zamojc, K.; Wyrzykowski, D.; Uber, D.; Wierzbicka, M.; Wiczk, W.; Chmurzynski, L. Binding of Cu(II) ions to peptides studied by fluorescence spectroscopy and isothermal titration calorimetry. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 153, 451. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Cai, Y.; He, Y.; Lin, Q.; Ren, J.; Cao, D.; Zhang, L. A label-free fluorescent peptide probe for sensitive and selective determination of copper and sulfide ions in aqueous systems. RSC Adv. 2021, 11, 7426. [Google Scholar] [CrossRef]
  24. Żamojć, K.; Kamrowski, D.; Zdrowowicz, M.; Wyrzykowski, D.; Wiczk, W.; Chmurzyński, L.; Makowska, J. A pentapeptide with tyrosine moiety as fluorescent chemosensor for selective nanomolar-level detection of copper(II) ions. Int. J. Mol. Sci. 2020, 21, 743. [Google Scholar] [CrossRef]
  25. White, B.R.; Holcombe, J.A. Fluorescent peptide sensor for the selective detection of Cu2+. Talanta 2007, 71, 2015. [Google Scholar] [CrossRef]
  26. Pang, X.; Gao, L.; Feng, H.; Li, X.; Kong, J.; Li, L. A peptide-based multifunctional fluorescent probe for Cu2+, Hg2+ and biothiols. New J. Chem. 2018, 42, 15770. [Google Scholar] [CrossRef]
  27. Pang, X.; Wang, L.; Gao, L.; Feng, H.; Kong, J.; Li, L. Multifunctional peptide-based fluorescent chemosensor for detection of Hg2+, Cu2+ and S2− ions. Luminescence 2019, 34, 585. [Google Scholar] [CrossRef]
  28. Wang, P.; Liu, L.; Zhou, P.; Wu, W.; Wu, J.; Liu, W.; Tang, Y. A peptide-based fluorescent chemosensor for multianalyte detection. Biosens. Bioelectron. 2015, 72, 80. [Google Scholar] [CrossRef]
  29. Wang, P.; Wu, J.; Su, P.; Xu, C.; Ge, Y.; Liu, D.; Liu, W.; Tang, Y. Fluorescence “on-off-on” peptide-based chemosensor for the selective detection of Cu2+ and S2− and its application in living cell bioimaging. Dalton Trans. 2016, 45, 16246. [Google Scholar] [CrossRef]
  30. Wang, P.; Wu, J. Highly selective and sensitive detection of Zn(II) and Cu(II) ions using a novel peptide fluorescent probe by two different mechanisms and its application in live cell imaging. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 208, 140. [Google Scholar] [CrossRef]
  31. An, Y.; Wang, P.; Yue, Z. A sequential and reversibility fluorescent pentapeptide probe for Cu(II) ions and hydrogen sulfide detections and its application in two different living cells imaging. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 216, 319. [Google Scholar] [CrossRef]
  32. Wang, P.; Wu, J.; Zhao, C. A water-soluble peptide fluorescent chemosensor for detection of cadmium(II) and copper(II) by two different response modes and its application in living LNcap cells. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 226, 117600. [Google Scholar] [CrossRef]
  33. Wang, P.; Xue, S.; Yang, X. A novel peptide-based fluorescent chemosensor for detection of zinc(II) and copper(II) through differential response and application in logic gate and bioimaging. Microchem. J. 2020, 158, 105147. [Google Scholar] [CrossRef]
  34. Hu, Y.; Quan, S.; Zhao, C.; Li, J.; Sun, X.; Xiao, J. An “on–off–on” fluorescent peptide probe for the specific detection of Cu2+ and S2− in living cells and zebrafish. New J. Chem. 2022, 46, 7663. [Google Scholar] [CrossRef]
  35. Zheng, Y.; Cao, X.; Orbulescu, J.; Konka, V.; Andreopoulos, F.M.; Pham, S.M.; Leblanc, R.M. Peptidyl fluorescent chemosensors for the detection of divalent copper. Anal. Chem. 2003, 75, 1706. [Google Scholar] [CrossRef]
  36. Zhang, X.; Zhang, G.; Wei, G.; Su, Z. One-pot, in-situ synthesis of 8-armed poly(ethylene glycol)-coated Ag nanoclusters as a fluorescent sensor for selective detection of Cu2+. Biosensors 2020, 10, 131. [Google Scholar] [CrossRef]
  37. Qian, D.; Wang, Z.; Xiao, Z.; Fang, C.-J. A fluorescent probe for the detection of Cu(II) in water and tumor cells. Inorg. Chem. Commun. 2021, 126, 108471. [Google Scholar] [CrossRef]
  38. Zhuang, H.; Jiang, X.; Wu, S.; Wang, S.; Pang, Y.; Huang, Y.; Yan, H. A novel polypeptidemodified fluorescent gold nanoclusters for copper ion detection. Sci. Rep. 2022, 12, 6624. [Google Scholar] [CrossRef]
  39. Falcone, E.; Gonzalez, P.; Lorusso, L.; Seneque, O.; Faller, P.; Raibaut, L. A terbium(III) luminescent ATCUN-based peptide sensor for selective and reversible detection of copper(II) in biological media. Chem. Commun. 2020, 56, 4797. [Google Scholar] [CrossRef]
  40. Falcone, E.; Vileno, B.; Hoang, M.; Raibaut, L.; Faller, P. A luminescent ATCUN peptide variant with enhanced properties for copper(II) sensing in biological media. J. Inorg. Biochem. 2021, 221, 111478. [Google Scholar] [CrossRef]
  41. Jung, K.H.; Oh, E.T.; Park, H.J.; Lee, K.H. Development of new peptide-based receptor of fluorescent probe with femtomolar affinity for Cu+ and detection of Cu+ in Golgi apparatus. Biosens. Bioelectron. 2016, 85, 437. [Google Scholar] [CrossRef]
  42. Li, Y.; Di, C.; Wu, J.; Si, J.; Chen, Y.; Zhang, H.; Ge, Y.; Liu, D.; Liu, W. A peptide-based fluorescent sensor for selective imaging of glutathione in living cells and zebrafish. Anal. Bioanal. Chem. 2020, 412, 481. [Google Scholar] [CrossRef]
  43. Wang, P.; Xue, S.; Zhou, D.; Guo, Z.; Wang, Q.; Guo, B.; Yang, X.; Wu, J. Peptide-based colorimetric and fluorescent dual-functional probe for sequential detection of copper(II) and cyanide ions and its application in real water samples, test strips and living cells. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 276, 121222. [Google Scholar] [CrossRef]
  44. Wang, P.; Wu, J.; Di, C.; Zhou, R.; Zhang, H.; Su, P.; Xu, C.; Zhou, P.; Ge, Y.; Liu, D.; et al. A novel peptide-based fluorescence chemosensor for selective imaging of hydrogen sulfide both in living cells and zebrafish. Biosens. Bioelectron. 2017, 92, 602. [Google Scholar] [CrossRef]
  45. Wang, P.; Sun, L.; Wu, J.; Yang, X.; Lin, P.; Wang, M. A dual-functional colorimetric and fluorescent peptide-based probe for sequential detection of Cu2+ and S2− in 100% aqueous buffered solutions and living cells. J. Hazard. Mater. 2021, 407, 124388. [Google Scholar] [CrossRef]
  46. Wang, P.; Xue, S.; Yang, X. Highly selective and sensitive detection of hydrogen sulfide in aqueous medium and live cells using peptide-based bioprobe to mimic the binding sites of the ceruloplasmin for Cu(II) ions. Biosens. Bioelectron. 2020, 163, 112283. [Google Scholar] [CrossRef]
  47. Hao, C.; Guo, X.; Lai, Q.; Li, Y.; Fan, B.; Zeng, G.; He, Z.; Wu, J. Peptide-based fluorescent chemical sensors for the specific detection of Cu2+ and S2−. Inorg. Chim. Acta 2020, 513, 119943. [Google Scholar] [CrossRef]
  48. Jung, K.H.; Lee, K.H. Efficient ensemble system based on the copper binding motif for highly sensitive and selective detection of cyanide ions in 100% aqueous solutions by fluorescent and colorimetric changes. Anal. Chem. 2015, 87, 9308. [Google Scholar] [CrossRef]
  49. Hao, C.; Li, Y.; Fan, B.; Zeng, G.; Zhang, D.; Bian, Z.; Wu, J. A new peptide-based chemosensor for selective imaging of copper ion and hydrogen sulfide in living cells. Microchem. J. 2020, 154, 104658. [Google Scholar] [CrossRef]
  50. Deng, D.; Hao, Y.; Yang, P.; Xia, N.; Yu, W.; Liu, X.; Liu, L. Single-labeled peptide substrates for detection of protease activity based on the inherent fluorescence quenching ability of Cu2+. Anal. Methods 2019, 11, 1248. [Google Scholar] [CrossRef]
  51. Song, N.; Zhang, Z.; Liu, P.; Yang, Y.-W.; Wang, L.; Wang, D.; Tang, B.Z. Nanomaterials with supramolecular assembly based on AIE luminogens for theranostic applications. Adv. Mater. 2020, 32, 2004208. [Google Scholar] [CrossRef] [PubMed]
  52. Gao, M.; Tang, B.Z. Fluorescent sensors based on aggregation-induced emission: Recent advances and perspectives. ACS Sens. 2017, 2, 1382. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, D.; Ji, S.; Li, H.; Hong, L.; Kong, D.; Qi, X.; Ding, D. Cellular membrane-anchored fluorescent probe with aggregation-induced emission characteristics for selective detection of Cu2+ ions. Faraday Discuss. 2017, 196, 377. [Google Scholar] [CrossRef] [PubMed]
  54. Gour, N.; Kshtriya, V.; Gupta, S.; Koshti, B.; Singh, R.; Patel, D.; Joshi, K.B. Synthesis and aggregation studies of a pyridothiazole-based AIEE probe and its application in sensing amyloid fibrillation. ACS Appl. Bio. Mater. 2019, 2, 4442. [Google Scholar] [CrossRef]
  55. Cai, Y.; Shi, Y.; Wang, H.; Wang, J.; Ding, D.; Wang, L.; Yang, Z. Environment-sensitive fluorescent supramolecular nanofibers for imaging applications. Anal. Chem. 2014, 86, 2193. [Google Scholar] [CrossRef]
  56. Kim, I.; Jeong, H.H.; Kim, Y.J.; Lee, N.E.; Huh, K.M.; Lee, C.S.; Kim, G.H.; Lee, E. A “Light-up” 1D supramolecular nanoprobe for silver ions based on assembly of pyrene-labeled peptide amphiphiles: Cell-imaging and antimicrobial activity. J. Mater. Chem. B 2014, 2, 6478. [Google Scholar] [CrossRef]
  57. Xia, N.; Wang, X.; Yu, J.; Wu, Y.; Cheng, S.; Xing, Y.; Liu, L. Design of electrochemical biosensors with peptide probes as the receptors of targets and the inducers of gold nanoparticles assembly on electrode surface. Sens. Actuat. B Chem. 2017, 239, 834. [Google Scholar] [CrossRef]
  58. Korkmaz, N.; Hwang, C.; Kessler, K.K.; Silina, Y.E.; Müller, L.; Park, J. A novel copper (II) binding peptide for a colorimetric biosensor system design. Talanta 2021, 232, 122439. [Google Scholar] [CrossRef]
  59. Wang, C.; Wang, K.; Wang, Z. Development of gold nanoparticle based colorimetric method for quantitatively studying the inhibitors of Cu2+/Zn2+ induced beta-amyloid peptide assembly. Anal. Chim. Acta 2015, 858, 42. [Google Scholar] [CrossRef]
  60. Zhou, Y.; Dong, H.; Liu, L.; Xu, M. Simple colorimetric setection of amyloid beta-peptide (1–40) based on aggregation of gold nanoparticles in the presence of copper ions. Small 2015, 11, 2144. [Google Scholar] [CrossRef]
  61. Pelin, J.N.B.D.; Edwards-Gayle, C.J.C.; Martinho, H.; Gerbelli, B.B.; Castelletto, V.; Hamley, I.W.; Alves, W.A. Self-assembled gold nanoparticles and amphiphile peptides: A colorimetric probe for copper(II) ion detection. Dalton Trans. 2020, 49, 16226. [Google Scholar] [CrossRef] [PubMed]
  62. Ghodake, G.S.; Shinde, S.K.; Saratale, R.G.; Kadam, A.A.; Saratale, G.D.; Syed, A.; Ameen, F.; Kim, D.Y. Colorimetric detection of Cu(2+) based on the formation of peptide-copper complexes on silver nanoparticle surfaces. Beilstein J. Nanotechnol. 2018, 9, 1414. [Google Scholar] [CrossRef] [PubMed]
  63. Hu, T.; Chen, C.; Huang, G.; Yang, X. Antibody modified-silver nanoparticles for colorimetric immuno sensing of Aβ(1–40/1–42) based on the interaction between β-amyloid and Cu2+. Sens. Actuat. B Chem. 2016, 234, 63. [Google Scholar] [CrossRef]
  64. Guarise, C.; Pasquato, L.; Filippis, V.D.; Scrimin, P. Gold nanoparticles-based protease assay. Proc. Natl. Acad. Sci. USA 2006, 103, 3978. [Google Scholar] [CrossRef]
  65. Liu, L.; Deng, D.; Wang, Y.; Song, K.; Shang, Z.; Wang, Q.; Xia, N.; Zhang, B. A colorimetric strategy for assay of protease activity based on gold nanoparticle growth controlled by ascorbic acid and Cu(II)-coordinated peptide. Sens. Actuat. B Chem. 2018, 266, 246. [Google Scholar] [CrossRef]
  66. Sousa, C.P.; Coutinho-Neto, M.D.; Liberato, M.S.; Kubota, L.T.; Alves, W.A. Self-assembly of peptide nanostructures onto an electrode surface for nonenzymatic oxygen sensing. J. Phys. Chem. C 2015, 119, 1038–1046. [Google Scholar] [CrossRef]
  67. Xia, N.; Huang, Y.; Zhao, Y.; Wang, F.; Liu, L.; Sun, Z. Electrochemical biosensors by in situ dissolution of self-assembled nanolabels into small monomers on electrode surface. Sens. Actuat. B Chem. 2020, 325, 128777. [Google Scholar] [CrossRef]
  68. Mitra, S.; Prakash, D.; Rajabimoghadam, K.; Wawrzak, Z.; Prasad, P.; Wu, T.; Misra, S.K.; Sharp, J.S.; Garcia-Bosch, I.; Chakraborty, S. De novo design of a self-assembled artificial copper peptide that activates and reduces peroxide. ACS Catal. 2021, 11, 10267. [Google Scholar] [CrossRef]
  69. Sun, T.; Xia, N.; Yuan, F.; Liu, X.; Chang, Y.; Liu, S.; Liu, L. A colorimetric method for determination of the prostate specific antigen based on enzyme-free cascaded signal amplification via peptide-copper(II) nanoparticles. Microchim. Acta 2020, 187, 116. [Google Scholar] [CrossRef]
  70. Wang, J.; Huang, R.; Qi, W.; Su, R.; Binks, B.P.; He, Z. Construction of a bioinspired laccase-mimicking nanozyme for the degradation and detection of phenolic pollutants. Appl. Catal. B Environ. 2019, 254, 452. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Tian, X.; Li, X. Supramolecular assemblies of histidine-containing peptides with switchable hydrolase and peroxidase activities through Cu(II) binding and co-assembling. J. Mater. Chem. B 2022, 10, 3716. [Google Scholar] [CrossRef] [PubMed]
  72. Yang, W.; Chow, E.; Willett, G.D.; Hibbert, D.B.; Gooding, J.J. Exploring the use of the tripeptide Gly-Gly-his as a selective recognition element for the fabrication of electrochemical copper sensors. Analyst 2003, 128, 712. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, W.R.; Hibbert, D.B.; Zhang, R.; Willett, G.D.; Gooding, J.J. Stepwise synthesis of Gly-Gly-His on gold surfaces modified with mixed self-assembled monolayers. Langmuir 2005, 21, 260. [Google Scholar] [CrossRef]
  74. Liu, G.; Nguyen, Q.T.; Chow, E.; Böcking, T.; Hibbert, D.B.; Gooding, J.J. Study of factors affecting the performance of voltammetric copper sensors based on Gly-Gly-His modified glassy carbon and gold electrodes. Electroanalysis 2006, 18, 1141. [Google Scholar] [CrossRef]
  75. Wawrzyniak, U.E.; Ciosek, P.; Zaborowski, M.; Liu, G.; Gooding, J.J. Gly-Gly-His immobilized on monolayer modified back-side contact miniaturized sensors for complexation of copper ions. Electroanalysis 2013, 25, 1461. [Google Scholar] [CrossRef]
  76. Sam, S.S.; Chazalviel, J.N.; Gouget-Laemmel, A.C.; Ozanam, F.F.; Etcheberry, A.A.; Gabouze, N.E. Peptide immobilisation on porous silicon surface for metal ions detection. Nanoscale Res. Lett. 2011, 6, 412. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, X.; Li, H.; Li, X.; Chen, Y.; Yin, Y.; Li, G. Electrochemical assay of melanoma biomarker in human blood. Electrochem. Commun. 2014, 39, 12. [Google Scholar] [CrossRef]
  78. Chen, H.; Zhang, J.; Liu, X.; Gao, Y.; Ye, Z.; Li, G. Colorimetric copper(II) ion sensor based on the conformational change of peptide immobilized onto the surface of gold nanoparticles. Anal. Methods 2014, 6, 2580. [Google Scholar] [CrossRef]
  79. Chen, H.; Jia, S.; Zhang, J.; Jang, M.; Chen, X.; Koh, K.; Wang, Z. Sensitive detection of copper(II) ions based on the conformational change of peptides by surface plasmon resonance spectroscopy. Anal. Methods 2015, 7, 8942. [Google Scholar] [CrossRef]
  80. Jiang, Y.; Chen, X.; Lan, L.; Pan, Y.; Zhu, G.; Miao, P. Gly–Gly–His tripeptide- and silver nanoparticle-assisted electrochemical evaluation of copper(ii) ions in aqueous environment. New J. Chem. 2018, 42, 14733. [Google Scholar] [CrossRef]
  81. Mervinetsky, E.; Alshanski, I.; Hamo, Y.; Sandonas, L.M.; Dianat, A.; Buchwald, J.; Gutierrez, R.; Cuniberti, G.; Hurevich, M.; Yitzchaik, S. Copper induced conformational changes of tripeptide monolayer based impedimetric biosensor. Sci. Rep. 2017, 7, 9498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ding, X.; Yang, K.-L. Quantitative serine protease assays based on formation of copper(II)–oligopeptide complexes. Analyst 2015, 140, 340. [Google Scholar] [CrossRef] [PubMed]
  83. Synhaivska, O.; Mermoud, Y.; Baghernejad, M.; Alshanski, I.; Hurevich, M.; Yitzchaik, S.; Wipf, M.; Calame, M. Detection of Cu2+ ions with GGH peptide realized with Si-nanoribbon ISFET. Sensors 2019, 19, 4022. [Google Scholar] [CrossRef] [PubMed]
  84. Deng, D.; Liu, L.; Bu, Y.; Liu, X.; Wang, X.; Zhang, B. Electrochemical sensing devices using ATCUN-Cu(II) complexes as electrocatalysts for water oxidation. Sens. Actuat. B Chem. 2018, 269, 189. [Google Scholar] [CrossRef]
  85. Deng, D.; Hao, Y.; Yang, S.; Han, Q.; Liu, L.; Xiang, Y.; Tu, F.; Xia, N. A signal-on electrochemical biosensor for evaluation of caspase-3 activity and cell apoptosis by the generation of molecular electrocatalysts on graphene electrode surface for water oxidation. Sens. Actuat. B Chem. 2019, 286, 415. [Google Scholar] [CrossRef]
  86. Xia, N.; Deng, D.; Yang, S.; Hao, Y.; Wang, L.; Liu, Y.; An, C.; Han, Q.; Liu, L. Electrochemical immunosensors with protease as the signal label for the generation of peptide-Cu(II) complexes as the electrocatalysts toward water oxidation. Sens. Actuat. B Chem. 2019, 291, 113. [Google Scholar] [CrossRef]
  87. Xia, N.; Liu, G.; Zhang, S.; Shang, Z.; Yang, Y.; Li, Y.; Liu, L. Oxidase-mimicking peptide-copper complexes and their applications in sandwich affinity biosensors. Anal. Chim. Acta 2022, 1214, 339965. [Google Scholar] [CrossRef]
  88. Feng, Y.; Liu, G.; Zhang, C.; Li, J.; Li, Y.; Liu, L. Fluorescent immunoassay with a copper polymer as the signal label for catalytic oxidation of O-phenylenediamine. Molecules 2022, 27, 3675. [Google Scholar] [CrossRef]
  89. Feng, Y.; Liu, G.; Zhang, F.; Liu, J.; La, M.; Xia, N. A general, label-free and homogeneous electrochemical strategy for probing of protease activity and screening of inhibitor. Micromachines 2022, 13, 803. [Google Scholar] [CrossRef]
Scheme 1. The functions of peptide–copper complexes in the design of different biosensors.
Scheme 1. The functions of peptide–copper complexes in the design of different biosensors.
Biosensors 12 00809 sch001
Figure 1. Proposed binding and detection mode of cyanide ions using the peptide-based ensemble. Reprinted with permission from [48]. Copyright 2015 American Chemical Society.
Figure 1. Proposed binding and detection mode of cyanide ions using the peptide-based ensemble. Reprinted with permission from [48]. Copyright 2015 American Chemical Society.
Biosensors 12 00809 g001
Figure 2. Schematic representation of the “signal-on” or “signal-off” detection of caspase-3 (A) and β-secretase (B). Reprinted with permission from [50]. Copyright 2019 Royal Society of Chemistry.
Figure 2. Schematic representation of the “signal-on” or “signal-off” detection of caspase-3 (A) and β-secretase (B). Reprinted with permission from [50]. Copyright 2019 Royal Society of Chemistry.
Biosensors 12 00809 g002
Figure 3. Schematic representation of the NBD-FFYEEGGH and NBD-FFFDEVDGGH nanofibers for the detection of Cu2+ and caspase-3. Reprinted with permission from [55]. Copyright 2014 American Chemical Society.
Figure 3. Schematic representation of the NBD-FFYEEGGH and NBD-FFFDEVDGGH nanofibers for the detection of Cu2+ and caspase-3. Reprinted with permission from [55]. Copyright 2014 American Chemical Society.
Biosensors 12 00809 g003
Figure 4. Schematic illustrations of the AuNP-based colorimetric mechanism for Aβ (1–40) detection upon addition of Cu2+. Reprinted with permission from [60]. Copyright 2015 John Wiley and Sons.
Figure 4. Schematic illustrations of the AuNP-based colorimetric mechanism for Aβ (1–40) detection upon addition of Cu2+. Reprinted with permission from [60]. Copyright 2015 John Wiley and Sons.
Biosensors 12 00809 g004
Figure 5. (A) Chemical structure of Cu2+ complexes with AA (1) and GGH (2, the minimalist version of ATCUN motif). (B) Schematic representation of the Cu2+/Cu+-initiated redox cycling for protease detection. Reprinted with permission from [65]. Copyright 2018 Elsevier B.V.
Figure 5. (A) Chemical structure of Cu2+ complexes with AA (1) and GGH (2, the minimalist version of ATCUN motif). (B) Schematic representation of the Cu2+/Cu+-initiated redox cycling for protease detection. Reprinted with permission from [65]. Copyright 2018 Elsevier B.V.
Biosensors 12 00809 g005
Figure 6. Schematic illustration for (A) the preparation of Cu–PNPs by the self-assembly of KbiotinGHFF/Cu2+ and (B) the colorimetric immunoassay of PSA by Cu2+-catalyzed oxidation of TMB. Reprinted with permission from [69]. Copyright 2020 Springer Nature.
Figure 6. Schematic illustration for (A) the preparation of Cu–PNPs by the self-assembly of KbiotinGHFF/Cu2+ and (B) the colorimetric immunoassay of PSA by Cu2+-catalyzed oxidation of TMB. Reprinted with permission from [69]. Copyright 2020 Springer Nature.
Biosensors 12 00809 g006
Figure 7. (A) Chemical structure of GGGG-Cu(II) and GGH-Cu(II) complexes. (B) CVs of Cu(II) complexes with GGH, GGHG, and GGH-NH2 peptides at the scan rate of 100 mV/s. CVs of GGH–Cu(II) (C) and GGHG–Cu(II) (D) at different scan rates. The concentrations of peptide and Cu(II) are 600 μM and 400 μM, respectively. Reprinted with permission from [84]. Copyright 2018 Elsevier B.V.
Figure 7. (A) Chemical structure of GGGG-Cu(II) and GGH-Cu(II) complexes. (B) CVs of Cu(II) complexes with GGH, GGHG, and GGH-NH2 peptides at the scan rate of 100 mV/s. CVs of GGH–Cu(II) (C) and GGHG–Cu(II) (D) at different scan rates. The concentrations of peptide and Cu(II) are 600 μM and 400 μM, respectively. Reprinted with permission from [84]. Copyright 2018 Elsevier B.V.
Biosensors 12 00809 g007
Figure 8. Schematic representation of the electrochemical (A) and fluorescent (B) immunoassays with AuNP/peptide-Cu2+ conjugates as the nanolabels for the electrocatalytic reduction of oxygen and the catalyzed oxidation of AA, respectively. The resulting oxidization product DHA was reacted with OPD to form fluorescent DFQ. Reprinted with permission from [87]. Copyright 2022 Elsevier B.V.
Figure 8. Schematic representation of the electrochemical (A) and fluorescent (B) immunoassays with AuNP/peptide-Cu2+ conjugates as the nanolabels for the electrocatalytic reduction of oxygen and the catalyzed oxidation of AA, respectively. The resulting oxidization product DHA was reacted with OPD to form fluorescent DFQ. Reprinted with permission from [87]. Copyright 2022 Elsevier B.V.
Biosensors 12 00809 g008
Table 1. An overview on the analytical performances of different fluorescent probes.
Table 1. An overview on the analytical performances of different fluorescent probes.
ProbeAnalyteDetection Limit (nM)Ref.
Dansyl-Gly-Gly-Asp-Gly-Gly-Asp-Gly-Gly-Asp-Gly-Gly-Asp-Gly-Gly-Trp-CONH2Cu2+500[25]
Dansyl-His-Pro-Gly-Trp-NH2Cu2+, Hg2+105, 37[26]
Dansyl-Gly-His-Gly-Gly-Trp-COOHCu2+, Hg2+, S2−85, 25[27]
Dansyl-His-Lys-His- DansylCu2+, Zn2+, S2−78, 82[28]
Dansyl-Ser-Pro-Gly-His-NH2Cu2+, S2−88, 75[29]
Dansyl-His-Pro-Gly-Glu-NH2Cu2+, Zn2+15, 4.9[30]
Dansyl-Ser-Pro-Gly-His-Gly-NH2Cu2+, H2S23.5, 17.2[31]
Dansyl-Gly-Cys-NH2Cu2+, Cd2+26.3, 14.5[32]
Dansyl-Ser-Pro-Gly-His-NH2Cu2+, Zn2+4.1, 13.2[33]
Dansyl-Asp-Asp-Gly-Glu-Glu-NH2Cu2+, S2−47, 57[34]
Dansyl-Leu-Leu-CysCu+, Cu2+, Zn2+Not reported[41]
FITC-Ahx-His-Lys-His-NH2Cu2+, CN1.5, 12.7[43]
FITC-Ahx-Ser-Pro-Gly-His-NH2Cu2+, H2S31[44]
FITC-Ahx-Ser-Ser-His-NH2Cu2+, S2−76.7, 27.2[45]
FITC-Ahx-Gly-His-Lys-NH2Cu2+, H2S14.7[46]
FITC-Ahx-His-Glu-Phe-Cys-NH2Cu2+, S2−32, 53[47]
FITC-Ahx-His-Glu-Phe-His-NH2Cu2+, S2−54, 68[49]
NBD-Ser-Ser-HisCu2+, CNNot reported[48]
Table 2. Analytical performances of colorimetric biosensors based on peptide–copper complexes.
Table 2. Analytical performances of colorimetric biosensors based on peptide–copper complexes.
Signal ReportersAnalyteLinear RangeDetection LimitRef.
AuNPsAβ(1–40)Not reported0.6 nM[60]
AuNPsCu2+Not reported190 nM[61]
AuNPsCu2+0.08~1.44 μM0.16 nM[62]
Ab-AgNPsCu2+0.00025~0.125 μM0.086 nM[63]
AuNPsbeta-secretase1~75 nM1 nM[65]
Cu–PNPsPSA0.001–1 ng/mL1 pg/mL[69]
CH–CuepinephrineNot reported0.31 μg/mL[70]
Table 3. Analytical performances of electrochemical biosensors with peptide–Cu2+ complexes as the signal reporters.
Table 3. Analytical performances of electrochemical biosensors with peptide–Cu2+ complexes as the signal reporters.
Signal TagsAnalyteLinear RangeDetection LimitRef.
AuNPs/DNA/DCH–Cu2+DNA0.1~2.5 pM0.1 pM[84]
ATCUN–Cu2+caspase-30.0005~2 ng/mL2 pg/mL[85]
Ab2-CNTs-trypsinPSA0.01~2 ng/mL10 pg/mL[86]
Ab2/AuNP/peptidePSA0.001~0.5 ng/mL0.4 pg/mL[87]
Ab2-biotin-poly-(KH–Cu)20PSA0.01~2 ng/mL8 pg/mL[88]
GGH–Cu2+ACE1~50 mU/mL1 mU/mL[89]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, G.; Xia, N.; Tian, L.; Sun, Z.; Liu, L. Progress in the Development of Biosensors Based on Peptide–Copper Coordination Interaction. Biosensors 2022, 12, 809. https://doi.org/10.3390/bios12100809

AMA Style

Liu G, Xia N, Tian L, Sun Z, Liu L. Progress in the Development of Biosensors Based on Peptide–Copper Coordination Interaction. Biosensors. 2022; 12(10):809. https://doi.org/10.3390/bios12100809

Chicago/Turabian Style

Liu, Gang, Ning Xia, Linxu Tian, Zhifang Sun, and Lin Liu. 2022. "Progress in the Development of Biosensors Based on Peptide–Copper Coordination Interaction" Biosensors 12, no. 10: 809. https://doi.org/10.3390/bios12100809

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