Next Article in Journal
Recent Advances in Electrochemical Sensors for the Detection of Anti-Inflammatory and Antibiotic Drugs: A Comprehensive Review
Previous Article in Journal
A Biofuel Cell for Electricity Generation from Biomass-Derived Cellobiose
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structure-Based Understanding of Cu2+ Coordination in Fluorescent Proteins for Metal Biosensor Applications—A Review

College of General Education, Kookmin University, Seoul 02707, Republic of Korea
Biosensors 2025, 15(10), 675; https://doi.org/10.3390/bios15100675
Submission received: 2 September 2025 / Revised: 29 September 2025 / Accepted: 4 October 2025 / Published: 7 October 2025
(This article belongs to the Special Issue Fluorescent Probes: Design and Biological Applications)

Abstract

Copper ions play essential roles in biological systems, but they can cause toxicity following dysregulation or excessive accumulation. In addition, environmental overexposure to Cu2+ can lead to serious agricultural and ecological issues. Accurate detection of Cu2+ is therefore critical in both medical diagnostics and environmental monitoring. Fluorescent proteins (FPs), which are widely used in molecular and cell biology, have been suggested as attractive modalities for metal ion detection owing to their biocompatibility and specific responsiveness to metal ions. The fluorescence emission of FPs is efficiently quenched by Cu2+ in a reversible manner, suggesting the potential to develop Cu2+-responsive biosensors. To develop highly sensitive and selective Cu2+ biosensors based on FPs, an understanding of Cu2+ binding to FPs is crucial, along with FP engineering guided by structural analysis. In this study, the molecular properties of FPs and their fluorescence responses to metal ions were reviewed. The crystal structures of FPs complexed with Cu2+ were analyzed, revealing both specific and nonspecific Cu2+ binding modes. This structural analysis provides insights into the potential of engineering FPs to enhance sensitivity and selectivity for Cu2+ detection.

1. Introduction

Transition metal ions, such as copper, zinc, and iron, are essential for various biological processes in all living organisms, and play an important role in metabolism [1]. They function as cofactors for enzymes, assist in protein folding and substrate binding, participate in redox reactions, maintain protein structural stability, and regulate cellular signaling [2,3]. Among various transition metal ions, Cu2+ serves as a catalytic cofactor in several critical biological functions, such as cytochrome c oxidase for electron transfer during mitochondrial respiration [4], ceruloplasmin for iron transport and antioxidant activity [5], superoxide dismutase 1 for protection against oxidative stress [6], tyrosinase for melanin biosynthesis, and lysyl oxidase for tissue stability [7].
In humans, disrupted copper homeostasis has been implicated in various diseases [8]. For example, Wilson’s disease and Menkes disease are genetic disorders associated with copper imbalance [9,10], and increasing evidence suggests a link between copper accumulation and neurodegenerative diseases, such as Alzheimer’s disease and Parkinson’s disease [11,12]. These findings highlight the importance of the accurate and sensitive detection of copper ions in biological systems.
Beyond biological systems, copper contamination in the environment is a growing concern [13]. The widespread use of copper-based compounds in agriculture, industry, and construction has contributed to the release of copper into water and soil systems [14]. Elevated copper concentrations can disrupt aquatic ecosystems, affect the physiology of aquatic organisms, and alter microbial communities in soil [15].
Therefore, monitoring copper levels in environmental samples is essential for ecological risk assessment and pollution control [16,17]. Various analytical techniques, such as atomic absorption spectroscopy (AAS) [18], inductively coupled plasma mass spectrometry (ICP-MS) [19], complexonometric assays [20], and electrochemical assays [21], have been developed to detect metal ions, including copper (Figure 1). Among these methods, the limit of detection (LOD) of ICP-MS is typically in the range of a few to several tens of nanomolar [22,23]. In electrochemical assays, the LOD of Cu2+ strongly depends on the type of sensor. For example, a Ag nanoparticle-based sensor can detect Cu2+ in the range of 1.0–1000 nM with a detection limit of 0.48 nM [24]. A ZnO–graphene photoelectrochemical sensor exhibits a linear detection range from 0.2 µM to 1 mM with a detection limit of 0.03 µM [25]. Using screen-printed electrodes modified with Fe3O4-embedded carbon nanofibers and multilayer carbon nanotubes, detection limits of 32 nM and 1.04 µM, respectively, have been reported [26]. Accordingly, these methods are highly sensitive and accurate, but they often require sophisticated instruments, expensive equipment, and labor-intensive sample preparation [27]. As such, they are not ideal for real-time monitoring or on-site applications. To measure metal ion levels, various biomolecules such as peptides [28], enzymes [29], antibodies [30], nucleic acids [31], DNAzymes [32], and whole cells [33] have been developed as metal biosensors. However, these biomolecules often face limitations concerning stability, cost, and detection efficiency. In addition, Cu2+ detection using organic fluorophores [34], aggregation-induced emission luminogens (AIEgens) [35], and metal–organic frameworks (MOFs) [36,37] has been developed and shown to provide higher sensitivity with lower limits of detection. However, these methods are less suitable for cellular compatibility and are not genetically encodable.
Fluorescence-based detection methods have gained attention because of their high sensitivity, rapid response, and compatibility with in vivo and real-time analysis [38,39,40]. Among fluorescent probes, fluorescent proteins (FPs) are widely used as optical probes in molecular and cell biology studies, as they offer distinct advantages such as a lack of toxicity, the ability to be genetically encoded and expressed in living cells, and the capability for spatiotemporal monitoring of target molecules [41,42]. For efficient FP applications, studies on the discovery of new fluorescent proteins and their engineering are ongoing [43,44]. In addition, FPs have been utilized as versatile biosensors for detecting various factors such as pH and chloride ions [45,46,47,48]. FPs undergo fluorescence quenching in the presence of certain transition metal ions, with Cu2+ causing particularly pronounced quenching effects [49,50]. This metal-induced quenching property has attracted considerable interest in the use of FPs as biosensor probes for Cu2+ detection (Figure 1). The Cu2+-induced fluorescence quenching of FPs has been proposed as a strategy for developing biosensors that respond to changes in copper concentrations [51,52], enable the imaging of copper fluctuations [53], support in vitro diagnostics [51], and permit intracellular copper detection [54]. To develop FPs with high selectivity and sensitivity toward Cu2+, protein engineering is essential to minimize the nonspecific binding of other metal ions. For rational protein engineering, a clear understanding of the mechanism by which Cu2+ interacts with FPs is crucial, including the identification of potential metal-binding sites. Elucidating the mechanism of Cu2+-induced quenching will facilitate the rational design of FP-based biosensors for copper detection in biological or environmental samples. However, the detailed structural basis of Cu2+ binding to FPs remains poorly understood.
This review introduces the fluorescence quenching of FPs by transition metal ions. The crystal structures of Cu2+-bound FPs were analyzed to identify potential specific and nonspecific Cu2+-binding sites. These findings provide insights into the mechanism by which Cu2+ interacts with FPs and offer guidance for engineering FPs with improved sensitivity and selectivity for use in Cu2+-responsive biosensors.

2. Fluorescent Proteins

Green FP (GFP) was first discovered in the jellyfish Aequorea victoria [55], and it has subsequently been discovered in other marine species, including corals and sea anemones [56,57]. FPs emit intrinsic fluorescence when excited by specific wavelengths of light. The optical properties of FPs can be affected by the chromophore sequence, surrounding environment, and overall protein conformation [57]. To date, various optical properties of FPs, such as photoactivation, photoswitching, photoconversion, and large Stokes shifts, have been discovered or developed through protein engineering and selectively utilized depending on the research purpose [58,59,60]. In particular, FPs are commonly employed in cellular imaging to monitor gene expression [61], protein localization and trafficking [62], and cellular structures in vivo [63]. Additionally, their ability to participate in fluorescence resonance energy transfer (FRET) enables real-time studies of molecular interactions and conformational changes [64,65].
FPs derived from A. victoria, corals, sea anemones, and other species share a highly conserved structure. Generally, FPs consist of 11 β-sheets and two α-helices, forming a single β-barrel fold with a nearly cylindrical shape [56]. The tripeptide corresponding to the chromophore is formed through autocatalytic posttranslational modification processes, including cyclization, dehydration, and oxidation, without the need for external cofactors [57]. The conjugated π-system of the chromophore plays a key role in light absorption and fluorescence emission. The chromophore is located inside the β-barrel, positioned near its center and connected by two α-helices. This rigid β-barrel structure shields the chromophore from water access, thereby enhancing quantum yield and photostability [57]. Mutation of the β-barrel or chromophore region leads to shifts in excitation/emission wavelengths, fluorescence intensity, and photostability [56,57].
The fluorescence intensity of FPs can vary in response to environmental factors such as pH, temperature, and the presence of metal ions [66,67,68]. These environment-sensitive fluorescence changes make FPs useful as biosensors. In particular, pH-dependent fluorescence changes have been widely applied in various imaging experiments [69]. Changes in pH can induce conformational alterations in the FP chromophore [70] or modify the hydrogen-bonding network surrounding the chromophore [71]. Exposure to specific transition metal ions can lead to the quenching of FP fluorescence, suggesting their potential use in the development of metal ion biosensors [68]. To investigate the fluorescence quenching effect of FPs by metal ions, spectroscopic analyses have been conducted for various FPs, including GFP [72], DsRed [51,73,74], Dronpa [75], AmCyan [76], ZsGreen [77], ZsYellow [78], and DendFP [52]. In addition, protein engineering has been employed to enhance metal binding in FP variants, such as BFPms1 [79] and iq-mEmerald [80]. These FPs exhibit significant fluorescence quenching in the presence of metal ions such as Cu2+, Fe2+, and Fe3+. Spectroscopic analyses suggested that that the fluorescence quenching of FPs induced by metal ions can occur through several mechanisms, such as static quenching [81], energy transfer [72], electron transfer [82], and structural perturbation [83].

3. Fluorescence Quenching and Reversibility of FP Fluorescence by Cu2+

Among various metal ions capable of quenching fluorescence, Cu2+ is particularly effective at quenching the fluorescence of various FPs [53,74], highlighting the potential of FPs as sensing probes for Cu2+ quantification. For example, the red FP DsRed and its variants (DsRed-Express, DsRed-Monomer, DsRed2, and Rmu13) selectively bind Cu2+ even in the presence of other divalent cations, leading to fluorescence quenching [73,74,84]. This quenching effect has been employed to develop in vitro biosensing systems for Cu2+ detection by correlating the extent of the fluorescence reduction with the Cu2+ concentration [74]. Furthermore, this copper-binding property of DsRed has potential applications in intracellular Cu2+ detection through fluorescence-based approaches [51]. Following the initial studies on the potential of DsRed as a Cu2+ biosensor, spectroscopic and structural investigations of various FPs in response to metal ions have been conducted to elucidate the mechanisms of fluorescence changes and to discover FPs with enhanced sensitivity to Cu2+. To date, various spectroscopic properties of FP in response to Cu2+-induced fluorescence quenching have been reported, including quenching efficiency, binding constants, and limits of detection (Table 1).
Spectroscopic analysis of ZsYellow revealed that Cu2+ caused the most pronounced quenching effect among the tested metal ions, reducing fluorescence intensity by 81.4% [78]. In addition, Mn2+, Co2+, Ni2+, Zn2+, and Cd2+ reduced fluorescence by approximately 21.1%, 54.0%, 35.2%, 25.9%, and 32.1%, respectively. These results indicate that ZsYellow exhibits high sensitivity to Cu2+ and significant quenching responses to other transition metal ions. Structural analysis further revealed that Cu2+ binds nonspecifically to ZsYellow, and the binding affinity is relatively weak.
For AmCyan and mOrange2, Cu2+ induced significant quenching, reducing their fluorescence intensities by 90% and 89%, respectively [76]. In addition, AmCyan exhibited quenching in the presence of Co2+ and Zn2+, with fluorescence reductions of approximately 40% and 50%, respectively, whereas mOrange2 was much less affected by these ions (reduction of approximately 7%). Time-resolved fluorescence measurements after Cu2+ addition illustrated that the fluorescence of AmCyan gradually decreased over 5 min, followed by partial recovery, whereas that of mOrange2 steadily declined without recovery throughout the measurement period [76]. The dissociation constant (Kd) for Cu2+ was estimated to be 56.10 µM for AmCyan and 21.46 µM for mOrange2, suggesting differences in Cu2+-dependent binding and quenching dynamics between the two FPs.
Visible fluorescence quenching was observed for DendFP upon the addition of Fe2+, Fe3+, and Cu2+ (Figure 2a) [52]. Spectroscopic data confirmed that Co2+, Ni2+, Zn2+, and Cd2+ also reduced the fluorescence intensity of DendFP (Figure 2b), indicating that this FP is susceptible to fluorescence quenching by several metal ions. Binding affinity analysis revealed that the Kd values of Fe2+, Fe3+, and Cu2+ were 24.59, 41.66, and 137.18 µM, respectively, indicating that Fe2+ and Fe3+ exhibited stronger binding and quenching effects than Cu2+.
Collectively, many FPs exhibit high sensitivity to Cu2+, highlighting their potential utility as Cu2+ biosensor probes. Notably, mOrange2 displayed strong fluorescence quenching, specifically by Cu2+, with minimal interference from other metal ions, suggesting favorable selectivity. Contrarily, DendFP exhibited effective quenching by Cu2+ but even greater quenching and binding by Fe2+ and Fe3+, indicating that without protein engineering, DendFP might not be suitable for Cu2+-selective sensing applications. Therefore, comprehensive screening against a broad panel of metal ions is essential for assessing the selectivity and suitability of FPs as Cu2+ biosensor probes.
To enhance the practical utility of Cu2+ biosensors, the FP probe must be reusable. In FP-based Cu2+ biosensors, this reusability requires the recovery of fluorescence emission following Cu2+ quenching. This reversibility has enabled the development of cost-effective and regenerable FP-based metal ion biosensors. To achieve this, the removal of Cu2+ ions bound to FPs is critical. Typically, chelating agents, such as EDTA or EGTA, are used to restore fluorescence by stripping Cu2+ from FP. The efficiency of fluorescence recovery depends on the chelating agent concentration. Moreover, the reversibility of fluorescence quenching differs depending on the type of FP used. For example, the fluorescence of AmCyan and mOrange2 was strongly quenched upon Cu2+ binding to lower than 20% of their wild-type fluorescence intensity. Treatment with 50 mM EDTA restored their fluorescence intensity to levels exceeding those of the wild-type FPs [76]. In the case of ZsYellow, approximately 90% of the quenched fluorescence was recovered after treatment with 50 mM EDTA [78], whereas DendFP displayed less than 50% recovery under the same conditions [52]. Although the spectroscopic properties of these reversible processes have been extensively studied, the molecular mechanisms underlying the differences in the recovery of FP fluorescence remain unclear. In addition, dialysis has been used to recover fluorescence emission from Cu2+-quenched FPs [85]. However, the recovery efficiency by dialysis varies among FPs. For instance, Cu2+-quenched mTFP1 recovered approximately 90% of its fluorescence after dialysis, whereas the engineered variant mTFP* exhibited even higher fluorescence recovery [85]. These results suggest that the reversibility efficiency of Cu2+-induced fluorescence quenching can be improved through protein engineering.

4. Structural Analysis of Cu2+ Binding to the FP

To understand the interaction between Cu2+ and FP, FP structures deposited in the Protein Data Bank (PDB) were investigated. The search revealed that only four FP structures (mTFPCHH, Dronpa, split-GFP, and GHK-GFP) contained Cu2+ bound to the protein surface. Accordingly, these FPs were selected for further analysis to provide insights into engineering FPs with improved selectivity and sensitivity for Cu2+ detection.

4.1. Quenchable Cu2+-Bound mTFPCHH

Artificial metalloenzymes (ArMs) can be used in biotechnological applications because they integrate the diverse catalytic potential of transition metals with the selective substrate recognition of enzymes [86,87]. Fischer et al. engineered the cyan FP mTFP from Clavularia to create a de novo functionalized fluorescent ArM scaffold named mTFP*, resulting in efficient Diels–Alder and Friedel–Crafts alkylase activities [85]. mTFP* was generated by substituting six surface-exposed histidine and methionine residues in mTFP1 (H30Y, M118L, H128Y, H177Y, H178Y, and H209Y) to minimize nonspecific interactions with metal ions. The nonspecific metal binding affinities of mTFP1 and mTFP* were assessed by incubating the proteins with Cu2+, Ni2+, Rh3+, and Pd2+ for up to 24 h, followed by attempts to reverse fluorescence quenching through dialysis. For both ArMs, Cu2+ induced relatively stronger fluorescence quenching than the other metal ions (Ni2+, Rh3+, and Pd2+). After dialysis, the fluorescence emission of mTFP* incubated with Cu2+ was fully recovered, whereas less than 80% recovery was observed for mTFP1. The binding affinity between the FPs and Cu2+ was further analyzed using time-resolved metal FRET (tmFRET) titration experiments. Titration fitting of mTFP* yielded Stern–Volmer constants of 0.09 ± 0.01 mM−1 at pH 6.0 and 1.12 ± 0.14 mM−1 at pH 7.5. Conversely, mTFP1 exhibited at least two distinct Kd values to achieve a satisfactory curve fit (R > 0.95): a site-specific Kd of 0.02 ± 0.01 μM and a background Kd of 330 ± 151 μM at pH 6.0; and values of 2.14 ± 0.29 and 268 ± 125 μM, respectively, at pH 7.5. These results indicate that substituting surface-exposed histidine and methionine residues in the FP can reduce the binding of nonspecific metal ions in mTFP*. However, the binding affinity of metal ions for mTFP* remained significantly low (Kd > 1 mM). Fischer et al. further engineered the mTFP* cavity to introduce a specific transition metal ion binding site by substituting I197C–Y200H–Y204H (mTFPCHH) and I197E–Y200H–Y204H (mTFPEHH) [85]. K d s i t e of Cu2+ for mTFPCHH was 0.56 ± 0.06 μM at pH 6.0, versus 0.05 ± 0.01 μM at pH 7.5. For mTFPEHH, the corresponding values were 8.2 ± 1.5 and 4.5 ± 2.7 μM, respectively. Liquid chromatography–electrospray ionization mass spectrometry and ICP–MS analyses confirmed the formation of a 1:1 complex between the engineered proteins (mTFPCHH or mTFPEHH) and Cu2+. By contrast, no Cu2+ coordination was detected for mTFP* using these methods.
The crystal structure of Cu2+-bound mTFPCHH (PDB code: 4R6D) was determined at 1.55 Å resolution. In the deposited Cu2+-bound mTFPCHH structure in the PDB, two Cu2+ ions were modeled (named Cu1 and Cu2; Figure 3a). Cu1 was located in the cleft of mTFPCHH, which is near the α-helix connected to the chromophore of mTFPCHH (Figure 3a). Cu1 was tetragonally coordinated by the NE2 atoms of His200 and His204, along with two water molecules, with coordination distances of 2.13, 2.12, 2.35, and 2.77 Å, respectively (Figure 3b). The distance between Cu1 and the CA2 atom of the chromophore in mTFPCHH was 16.39 Å (Figure 3a). The distance between the chromophore and cleft fell within the appropriate range for tmFRET studies, namely 1.05–2.30 nm [88]. This indicates that metal ions can approach the chromophore within this range and quench its fluorescence via tmFRET [85]. Accordingly, Cu1 located in the engineered region of mTFPCHH represents a quenchable Cu2+-binding site.
Cu2 was positioned at the N-terminus (Figure 3a), which was modeled in the deposited PDB but was not discussed in the previous study [85]. Cu2 was coordinated by the backbone nitrogen and oxygen atoms of Gly2 and the NZ atom of Lys5 with coordination distances of 2.29, 2.12, and 2.18 Å, respectively (Figure 3c). The distance between Cu2 and the CA2 atom of the chromophore was 24.5 Å (Figure 3a), suggesting little to no effect on fluorescence quenching, as it lies outside the effective range for tmFRET studies. These findings indicate that Cu2+ at the N-terminus can interact nonspecifically with mTFPCHH. Although the Cu2 binding site is not directly involved in fluorescence quenching, the N-terminal region modification could be required to avoid nonspecific Cu2+ binding and enable precise quantification of Cu2+ interactions with mTFPCHH.
Accordingly, the Cu1 site is critical for the fluorescence quenching effect in mTFPCHH. To understand the mechanism by which Cu1 binding affects the structural changes of mTFPCHH, the crystal structure of Cu2+-bound mTFPCHH was compared with apo-mTFPCHH (PDB code: 6QSL) determined at 1.6 Å resolution. The superimposition of mTFPCHH and mTFPCHH–Cu2+ revealed a root mean square deviation of 0.288 Å for all Cα atoms. The superimposed structures revealed that the side chains of His200 and His204 in mTFPCHH shifted toward the inside of the β-barrel cleft upon Cu2+ coordination. The distance between the ND1 atom of His204 in apo-mTFPCHH and Cu2+-bound mTFPCHH was 1.64 Å, and the distance between the NE2 atom of His200 in the two structures was 3.31 Å (Figure 3d). These positional and conformational changes in His200 and His204 upon Cu2+ binding affected the conformation of the loop between Leu198 and His204 (Figure 3d). In addition, Cu2+ binding to the cleft of mTFPCHH reduced the flexibility of this loop (Figure 3e). The B-factor values of the loop between Leu198 and His204 from apo-mTFPCHH and Cu2+-bound mTFPCHH were 24.73 and 15.32 Å, respectively. These results indicate that Cu2+ binding to mTFPCHH affects both the flexibility of the loop regions and the engineered metal-binding site.

4.2. Quenchable Cu2+-Bound Dronpa

Dronpa is a photoactivatable FP capable of reversible on/off switching between fluorescent and nonfluorescent states using different wavelengths of light [89,90]. This unique property makes this FP highly valuable for superresolution microscopy and dynamic cellular process tracking. Kim et al. performed spectroscopic and structural analyses to investigate whether Dronpa is suitable as a metal biosensor probe [75]. To identify the metal ions that induce fluorescence quenching, purified Dronpa was incubated with various divalent metal ion solutions (Mg2+, Ca2+, Mn2+, Co2+, Ni2+, Cu2+, or Zn2+). The fluorescence emission of Dronpa was significantly reduced by 86% in the presence of 50 μM Cu2+. Additionally, 50 μM Co2+ and Ni2+ reduced the fluorescence emission of Dronpa by 26% and 6.3%, respectively. Contrarily, the fluorescence emission of Dronpa was enhanced in the presence of Mg2+, Ca2+, Mn2+, and Zn2+. Notably, Cu2+-induced fluorescence quenching of Dronpa was highly reversible with approximately 95% recovery.
To determine the quenchable metal ion binding site, the crystal structure of Dronpa soaked with Cu2+ was determined at 2.84 Å resolution (PDB code: 5HZT). The crystal belongs to the triclinic P1 space group and contains 24 molecules in the asymmetric unit. Two Cu2+ ions (designated Cu1 and Cu2) were bound to the Dronpa molecule (Figure 4a). Cu1 was located on the surface at the midpoint of the β-barrel fold, positioned perpendicular to the Dronpa chromophore. Superimposition of the 24 Dronpa molecules revealed slight positional variation of the Cu1 and C2 sites (Figure 4b). The Cu1 ion was commonly coordinated by His210 and His212, and the side chains of two Lys37 residues from nearby Dronpa molecules also interacted with Cu1. However, the side-chain conformations of the metal-interacting histidine residues were not identical, indicating that Cu1 is not tightly bound (Figure 4b). Among all Cu1 sites, the electron density map for chain H was relatively clearer than in the other Cu2+-bound Dronpa molecules. In chain H, the Cu1 ion was distorted tetrahedrally coordinated by the NE2 atoms of His210 and His212 and two water molecules at distances of 2.28, 2.40, 2.26, and 3.41 Å, respectively (Figure 4c). For the 24 molecules in the asymmetric unit, the distances between Cu1 and the NE2 atoms of His210 and His212 in the 24 Dronpa molecules ranged from 2.10 to 2.95 Å and 2.11 to 3.01 Å, respectively. The distance between Cu1 and the CA2 atom of the chromophore ranged from 14.62 to 15.29 Å, and the shortest distance between Cu1 and the tyrosine ring of the Dronpa chromophore ranged from 12.67 to 13.68 Å. These values suggest that the Cu1 site, coordinated by His210 and His212, is responsible for quenching, as supported by previous FRET studies [88].
A second Cu2+ site (Cu2) was located at the bottom of the β-barrel. The Cu2+-binding site was not analyzed and described in the original manuscript [75]. Cu2 was observed in all Dronpa molecules in the asymmetric unit and was coordinated by the NE2 atom of His220, located at the edge of the β-strand (Figure 4c). Superimposition of the 24 Dronpa molecules illustrated that the side-chain conformation of His220 was conserved, although the position of Cu2 varied slightly, indicating weak coordination. The distance between Cu2 and the CA2 atom of the chromophore ranged from 19.27 to 20.26 Å (Figure 4a). Although this Cu2+-binding site has not been previously studied in detail, the distance suggests that Cu2 also contributes to fluorescence quenching, which warrants further investigation.
To understand Cu2+-induced structural changes at the binding site, the side-chain conformations of His210 and His212 were compared between the Cu2+-bound and native Dronpa structures. Superimposition demonstrated that the His210 and His212 side chains point away from the Cu2+-binding site in the native Dronpa structure. The average rotation angles of the His210 and His212 side chains were approximately 90 and 115°, respectively (Figure 4d). These findings suggest that Cu2+ binding can induce conformational changes in the metal-coordinating residues.
Compared with the findings for wild-type Dronpa, Co2+ and Ni2+ ions also reduced the fluorescence emission of Dronpa by 26% and 6.3%, respectively. The crystal structures of Co2+- and Ni2+-bound Dronpa were determined at a resolution of 1.90 and 2.15 Å, respectively. Two Co2+ ions and two Ni2+ ions were bound at two nearly identical sites on the FP surface, respectively (Figure 4e,i). Superimposition of the 24 molecules in the asymmetric unit of the Dronpa–Co and Dronpa–Ni structures revealed slight variations in metal ion positions, suggesting that Co2+ and Ni2+ are not rigidly bound (Figure 4f,j). The first metal binding site (Co1/Ni1) was octahedrally coordinated by His194, His212, and water molecules (Figure 4g,k). However, the coordination angles and distances slightly differed [75]. The second metal binding site (Co2/Ni2) was coordinated by His200 and was located near the N-terminal α-helix of the β-barrel (Figure 4g,k). Superimposition of the 24 molecules in the asymmetric unit of the Dronpa–Co and Dronpa–Ni structures revealed slight variations in the positions of the metal ions. The distances of Co1/Ni1 and Co2/Ni2 to the CA2 atom of the Dronpa chromophore were approximately 13.75–14.15/13.76–13.89 and 19.67–20.15/19.72–20.06 Å, respectively. These results suggest that Co1 are more likely responsible for the observed fluorescence reduction for Co2+. Compared with the findings for the native structure, the side chains of His194 and His212 exhibited conformational changes upon Co2+ or Ni2+ binding (Figure 4h,l), whereas no such changes were noted for His210 side chains. These findings indicate that although the Co2+ and Ni2+ binding sites and coordination geometries are similar, their fluorescence quenching effects differ significantly, underscoring the importance of the specific metal ion involved.
The positions and coordination geometries of Cu1 versus Co1/Ni1 were notably different even though all three ions were coordinated by His212, suggesting that the preferred binding position depends on the identity of the coordinating histidine residues [75]. By contrast, Cu2, Co2, and Ni2 were all coordinated by His200 alone, indicating a lack of binding selectivity at this site.

4.3. Cu2+ Bound to the Oligomeric Interface of Split-GFP

The spatial organization of multiple proteins has potential applications in synthetic biology, including metabolic pathway optimization [91], spatial arrangement of signaling molecules [92], and self-assembling protein architecture construction [93]. The development of efficient systems for precise protein spatial organization could also significantly advance research in protein crystallization [94]. Leibly et al. selected split-GFP, consisting of a GFP core (strands 1–9) and a GFP hairpin (strands 10–11), as a scaffold and engineered 11 distinct oligomeric forms by introducing specific disulfide bonds or metal-binding sites [95]. They explored the potential of forming GFP dimers and higher-order oligomers by designing metal-binding half-sites on the GFP surface, with metal ions such as Cu2+, Zn2+, or Ni2+ playing a stabilizing role by bridging multiple GFP units. Ultimately, this study did not focus on the fluorescence quenching of FPs by metal ion; therefore, no spectroscopic analysis was performed to assess the fluorescence quenching effect of metal ion bridging on GFP. However, the structural characterization of Cu2+ binding to FPs could provide valuable insights for the development of FP-based Cu2+ biosensors. The detailed structural properties of Cu2+ binding at the interface of the GFP oligomer were analyzed in this study.
In this study, 33 crystal structures of oligomeric states of split-GFP were determined, and they can be broadly categorized into four types: disulfide-mediated, metal ion-mediated, disulfide and metal contacts, and Cu2+ dimers. Only the five structures (PDB code: 4W6T, 4W7D, 4W7E, 4W7F, and 4W7R) that exclusively exhibited Cu2+-mediated oligomerization, excluding those involving both disulfide and Cu2+ interactions, were analyzed in this study.
For split-GFP-4W6T, two Cu2+ ions bound to the NE2 atoms of the engineered residues E115H and T118H at distances of 2.20 and 1.94 Å, respectively (Figure 5a). These two Cu2+ ions facilitated intermolecular interactions between adjacent split-GFP molecules within the crystal lattice (Figure 5b). The Cu2+ bound to E115H is also coordinated by the NE2 atom of His25, the OE1 and OE2 atoms of Glu132 from neighboring molecules in the crystal lattice, and two water molecules at distances of 1.99, 1.97, 2.93, 2.34, and 2.91 Å, respectively (Figure 5c). The Cu2+ bound to T118H was additionally coordinated by the OD2 atom of Asp133 from neighboring molecules in the crystal lattice and two water molecules at distances of 2.18, 1.83, and 2.95 Å, respectively (Figure 5c). The distances from the Cu2+ ions interacting with E115H and T118H to the CA2 atom of the split-GFP chromophore were 23.31 and 20.19 Å, respectively.
For split-GFP-4W7D, two split-GFP molecules were present in the asymmetric unit, each exhibiting distinct Cu2+ binding patterns. In chain A, three Cu2+ ions were modeled (Figure 5d), whereas no Cu2+ ions were modeled in chain B. However, in chain B, an interaction between a Cu2+ ion and a symmetry-related molecule was observed (Figure 5e). The distances from Cu1, Cu2, and Cu3 to the CA2 atom of the chromophore were 22.87, 26.98, and 28.81 Å, respectively (Figure 5d). In chain A, the first Cu2+ (Cu1) was coordinated by the NE2 atom of His21, the ND1 atom of His26, the OE2 atom of Glu6 from a neighboring molecule in the crystal lattice, and two water molecules with coordination distances of 2.27, 2.04, 2.06, 2.65, and 2.86 Å, respectively (Figure 5f). The second Cu2+ ion (Cu2) interacted with the NE2 atom of His26 and one water molecule at distances of 1.97 and 2.82 Å, respectively (Figure 5f). The third Cu2+ ion (Cu3) was coordinated by the NZ atom of Lys3, the NE2 atom of His21 from a neighboring molecule, and two water molecules at distances of 2.50, 2.06, 2.55, and 2.65 Å, respectively (Figure 5g). For split-GFP-4W7E, one Cu2+ was bound to the FP surface (Figure 5h). The distance between Cu2+ and the CA2 atom of the chromophore was 21.26 Å. Cu2+ was coordinated by the OD1 and OD2 atoms of Asp19, NE2 atom of His21, and N1 atom of an imidazole molecule derived from the crystallization solution (Figure 5i). The coordination distances were 3.21, 2.84, 2.61, and 2.96 Å, respectively. Cu2+ was additionally located near a neighboring molecule, but it did not interact with any neighboring residues in the crystal lattice (Figure 5i). Four FP molecules were present in the asymmetric unit of split-GFP-4W7R, and two Cu2+ molecules were located at the interface between two split-GFP molecules, exhibiting identical coordination patterns (Figure 5j). The distances between the Cu2+ ions and the CA2 atoms of the chromophores from the four split-GFP molecules ranged from 20.35 to 20.43 Å (Figure 5j). Each Cu2+ ion was coordinated by the NE2 atoms of His124 and His126 from two separate split-GFP molecules, with distances of 2.04–2.13 Å and 2.05–2.10 Å, respectively, in the four molecules of the asymmetric unit (Figure 5k). For split-GFP-4W7F, one Cu2+ ion was bound to the FP surface (Figure 5l). The distance between Cu2+ and the CA2 atom of the chromophore was 20.00 Å. The Cu2+ ion mediated an interaction with a neighboring molecule (Figure 5m). This ion was coordinated by the NE2 atoms of His124 and His126, both at distances of 2.21 Å. Additionally, this Cu2+ interacted with the OE1 and OE2 atoms of Glu5 from a neighboring molecule in the crystal lattice with coordination distances of 2.04 and 2.08 Å, respectively (Figure 5n).
Taken together, the split-GFP structures demonstrated that even for the same protein, Cu2+ can bind at various positions, exhibiting distinct binding modes through interactions with different combinations of amino acid residues. From the perspective of Cu2+ biosensor probe development, these results suggest that Cu2+ can form stable yet nonspecific interactions with various amino acid residues, particularly histidine, on the surface of FPs.

4.4. Cu2+-Binding GHK Tripeptide Fused to GFP

The phase problem remains a major bottleneck in protein crystallography because X-ray diffraction experiments only provide the amplitudes of scattered X-rays but not their phases [96]. Various experimental phasing methods have been developed to overcome this limitation, including molecular replacement, single- and multi-wavelength anomalous dispersion, and single- and multiple isomorphous replacement (SIR and MIR) [96].
Mehr et al. reported the high-affinity Cu2+-binding tripeptide GHK genetically fused to the N-terminus of a GFP variant and the fused peptide MBP–FG [97]. This study demonstrated that incorporation of the GHK–Cu2+ complex both facilitated crystallization and enabled successful experimental phasing using copper single-wavelength anomalous dispersion. Because the purpose of their study was unrelated to the effect of Cu2+ on fluorescence, it did not investigate the influence of Cu2+ binding to the GHK–GFP variant on the fluorescence intensity of the FP. Nevertheless, the interaction of the Cu2+ ion bound to the GHK tag at the N-terminus of GFP could offer structural insights for the future development of Cu2+ biosensors. Therefore, in the present study, the coordination of Cu2+ ions within the crystal structure of GHK-tagged GFP was analyzed in greater detail than in the previous report.
The crystal structures of GHK–GFP complexed with Cu2+ were determined in two different crystal forms: P6522 (PDB codes: 6QUJ and 6QUI) and P212121 (PDB code: 6QUH). The two GHK–GFP structures in the P6522 crystal form contained two molecules in the asymmetric unit, each exhibited a Cu2+ ion bound to the GHK tripeptide at the GFP N-terminus (Figure 6a). The Cu2+ ion binds to the GHK peptide of GFP and also interacts with neighboring GFP molecules in the crystal lattice (Figure 6b). For GHK-GFP-6QUJ, Cu2+ ion was coordinated by the N atom of glycine, N atom of histidine, and ND1 atom of histidine from the GHK tripeptide in all molecules of the P6522 form, and bond distances of 2.04/2.06 Å, 2.07/2.06 Å, and 2.07/2.03 Å for chain A/B, respectively (Figure 6c). In addition, Cu2+ was coordinated by the NE2 atom of His25 from a neighboring molecule in the crystal lattice and a water molecule at distances of 1.98–2.19 and 2.38–2.60 Å, respectively (Figure 6c). Although the exact coordination distances varied slightly, the square–pyramidal coordination geometry of the Cu2+ ion by the GHK tripeptide, His25, and a water molecule was essentially identical in all four GHK–GFP structures. The Cu2+ binding in GHK-GFP-6QUI were similar with GHK-GFP-6QUJ.
In the P212121 crystal form, the structure contained two molecules in the asymmetric unit, each with one Cu2+ ion bound to the GHK–GFP molecule (Figure 6d). In chains B/E (based on PDB assignment), Cu2+ was coordinated by the N atom of glycine, N atom of histidine, and ND1 atom of histidine from the GHK tripeptide at distances of 2.03/2.03, 2.04/2.03, and 2.03/2.03 Å, respectively (Figure 6e). The Cu2+ ion was also coordinated by residues from neighboring molecules, although the specific interactions differed because of variations in crystal packing, likely attributable to the different conformations of the Cu2+-bound GHK. In chain B, the GHK-bound Cu2+ was further coordinated by the OD1 and OD2 atoms of Asp74 from a neighboring molecule at distances of 2.00 Å and 2.84 Å, respectively (Figure 6e). In chain E, the Cu2+ is coordinated by the OE1 and OE2 atoms of Glu170 from a neighboring molecule, the main-chain carbonyl group of His137, and a water molecule at distances of 3.13 Å, 1.84 Å, and 2.49 Å, respectively (Figure 6e).
A second Cu2+-binding site was observed on the opposite side of the GHK tripeptide region on GFP. In chain B, this solvent-exposed second Cu2+ molecule had an occupancy of 0.29, and it was coordinated by the NE2 atom of His25, the OE1/OE2 atoms of Glu130, and a water molecule at distances of 2.03 2.12/2.13, and 2.41 Å, respectively (Figure 6d). In chain E, the second Cu2+ site had full occupancy (1.00), and it was coordinated by the NE2 atom of His25, and the OE1/OE2 atoms of Glu130 at distances of 2.01, and 2.19/2.76 Å, respectively (Figure 6f). Accordingly, the occupancy and coordination of the second Cu2+ site differed significantly depending on the crystal packing.

5. Discussion

FPs are widely used to monitor molecular functions in molecular and cell biology, and they also serve as biosensors that respond to environmental changes. In particular, metal ion-induced fluorescence quenching represents a useful strategy for detecting specific metal ions, such as Cu2+, with high sensitivity and in real time. Numerous studies have focused on the application of FPs as biosensors for detecting metal ions. Among the various metal ions that induce quenching, this review focused on Cu2+-induced fluorescence quenching and the associated metal-binding characteristics of FPs. Spectroscopic analyses have revealed that many FPs commonly exhibit high sensitivity to Cu2+-induced quenching, although their responses to other metal ions vary.
In real-world samples, multiple metal ions coexist with Cu2+, making selective quenching by Cu2+ a key requirement for biosensor development. For this reason, FPs that exhibit highly sensitive and selective quenching by Cu2+, such as mOrange2, are promising biosensor candidates. Conversely, FPs such as DendFP and ZsGreen exhibit stronger quenching in the presence of Fe2+ and Fe3+ than in the presence of Cu2+. Although these FPs might be suitable for detecting iron, their susceptibility to Cu2+ interference limits their utility as selective Fe ion sensors. Similarly, FPs such as ZsYellow and Dronpa also exhibit reduced fluorescence in the presence of Co2+ and Zn2+, which could lead to inaccurate Cu2+ quantification. Therefore, the discovery of novel FPs that selectively respond to Cu2+ or the engineering of proteins to bind Cu2+ specifically is critical.
The fluorescence recovery of Cu2+-quenched FPs by chelating reagents or dialysis has been reported, providing preliminary insights into the potential reusability of FP-based biosensors. However, the speed and reliability of this reversibility have not been thoroughly evaluated for real-time or continuous monitoring applications. From an applied perspective, further in-depth studies are required to establish efficient and rapid recovery processes to enable continuous Cu2+ sensing.
In this study, four different crystal structures of Cu2+-bound FPs were analyzed. Cu2+ was found to bind not only near the chromophore but also at the N-terminal or flexible regions of the proteins (Figure 7a). Cu2+ interacted with the imidazole ring of histidine, the carboxyl groups of aspartate or glutamate residues, or nonspecifically with the N-terminal backbone. Cu2+ generally prefers 4–6 coordination [98,99]. Meanwhile, the Cu2+ bound to FPs exhibited diverse coordination geometries. For example, in mTFPCHH and Dronpa, Cu2+ displayed a tetrahedral coordination involving histidine residues and water molecules (Figure 7b). In split-GFP (PDB: 4W7R), Cu2+ was coordinated in a square planar geometry by four histidine residues (Figure 7b). In contrast to such stable coordination geometries, in some cases, such as in Dronpa, Cu2+ interacted with only a single histidine residue, or with main-chain atoms, making it difficult to clearly define a specific coordination geometry (Figure 7c). Overall, Cu2+ can bind to multiple regions of FPs and adopt diverse coordination geometries through interactions with amino acid side chains and water molecules. Therefore, precise identification of Cu2+ binding sites and coordination modes requires experimental structural determination.
Structural analysis of Cu2+-bound FPs showed that Cu2+ commonly interacts with surface-exposed histidine or aspartic acid residues, but the binding specificity varies among FPs. For example, in mTFPCHH and Dronpa, despite the presence of multiple histidine and aspartic acid residues on the FP surface, Cu2+ binds at specific residues. In contrast, in split-GFP and GHK-GFP, Cu2+ interacts with different histidine or aspartic acid residues in a more nonspecific manner. Collectively, these results indicate that the binding specificity of Cu2+ differs among FPs. In the case of mTFP*, engineering efforts involved substituting surface histidine and methionine residues to prevent nonspecific metal binding [85]. Titration experiments confirmed the reduction in nonspecific interactions and illustrated that fluorescence was effectively recovered after dialysis. This strategy highlights the effectiveness of removing potential metal-binding residues to improve FP recovery and reduce background interference, ultimately enhancing the accuracy of metal quantification. Regarding Dronpa, the Cu2+-binding sites were distinct from those of Co2+ and Ni2+ [75], suggesting that selective Cu2+ quenching can be achieved by mutating Co2+- or Ni2+-binding residues while preserving Cu2+-binding ability. In split-GFP, Cu2+ bound at multiple locations within the same protein, potentially inducing protein aggregation [95]. Cu2+ tends to coordinate with surface histidine residues, thereby promoting interactions between neighboring FPs and leading to aggregation. Such aggregation might hinder accurate Cu2+ quantification, indicating the need to replace surface residues to prevent nonspecific Cu2+ binding for metal biosensor applications. Concerning GFP fused with the GHK tripeptide, the design was intended for specific Cu2+ binding [97]. This result illustrated that the presence of potential metal ion binding sites, such as the N-terminal backbone nitrogen, could permit nonspecific Cu2+ interactions. Although the chromophore is distant from this site and quenching might not occur, nonspecific binding could lead to inaccurate Cu2+ quantification. Thus, truncating the N- or C-terminal regions or substituting metal-preferred residues in FP-based probes could help eliminate nonspecific interactions. Structural data illustrated that Cu2+ binds to histidine and aspartate residues on the FP surface. To eliminate nonspecific Cu2+ interactions, surface histidine residues not involved in targeted sensing should be replaced.
To identify the quenchable Cu2+-binding site on FPs, experimental approaches such as co-crystallization of FPs with Cu2+ or soaking FP crystals into the crystallization solutions containing Cu2+ are commonly employed. These methods have been successfully used to obtain the crystal structures of Cu2+-bound Dronpa, mTFP, split-GFP, and GHK–GFP. However, co-crystallization and soaking with quenchable metal ions were unsuccessful in the case of DendFP because of protein precipitation or crystal damage. This suggests that DendFP has higher sensitivity to structural perturbation by quenchable Cu2+ or that its surface residues involved in quenchable Cu2+ binding cause aggregation or interfere with crystal packing. Given these challenges, a possible approach for future studies would be to start from extremely low concentrations and gradually increase the concentration of quenchable Cu2+ during the soaking process. This stepwise method could help minimize crystal damage and facilitate the identification of quenchable Cu2+-binding sites in FP.
Meanwhile, in the crystallographic study of ZsYellow with quenchable Cu2+, the color of ZsYellow crystals successfully changed to a quenched state after soaking in a crystallization solution containing Cu2+. However, Cu2+ was not observed in the electron density map, suggesting that Cu2+ bound nonspecifically to ZsYellow. This result is consistent with that of spectroscopic analysis, which indicated a low binding affinity of Cu2+ for ZsYellow. These findings suggest that although Cu2+ can reduce or quench the fluorescence emission of FPs, this might not be observed in the crystal structure if the binding affinity between the protein and Cu2+ is low because of weak or nonspecific interactions. Therefore, to identify Cu2+-binding sites in FPs, it is important to first evaluate the binding affinity between the FP and Cu2+, which can help assess the feasibility of crystallographic experiments.
When structural information is not available, nonspecific Cu2+ binding could be reduced by substituting surface-exposed histidine and aspartic acid residues on the β-barrel fold with other amino acids. However, such mutations may alter the fluorescence properties of the FP. Therefore, accurate identification of nonspecific Cu2+ binding sites and targeted substitution of the involved residues essentially require experimentally determined Cu2+-bound FP structures. When experimental structures are not available, metal ion docking to FPs can serve as an alternative approach [100]. Recent studies have shown that, although such models do not provide precise coordination geometries, some Cu2+ docking models can reliably predict potential binding sites when compared with experimentally determined Cu2+-bound FP structures [100].
Overall, FP-based Cu2+ biosensors are still at the developmental stage; nevertheless, they hold promise for diverse future applications, including in vivo imaging platforms for biomedical diagnostics, as well as environmental and food monitoring.

Funding

This work was supported by the National Research Foundation of Korea (NRF) (NRF-2021R1I1A1A01050838) and by PAL-XFEL, which is supported by MSIT and POSTECH (XFEL2025-01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

I would like to thank the Global Science experimental Data hub Center (GSDC) at the Korea Institute of Science and Technology Information (KISTI) for providing computing resources and technical support. During the revision of this manuscript, the author used ChatGPT-5 for minor corrections and adjustments of the English language. The author has reviewed and edited the output and takes full responsibility for the content of this publication.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Kapoor, S.; Brazete, D.; Pereira, I.C.; Bhatia, G.; Kaur, M.; Santos, L.F.; Banerjee, D.; Goel, A.; Ferreira, J.M.F. Impact of transition metal ions on the structure and bioactivity of alkali-free bioactive glasses. J. Non-Cryst. Solids 2019, 506, 98–108. [Google Scholar] [CrossRef]
  2. Kostenkova, K.; Scalese, G.; Gambino, D.; Crans, D.C. Highlighting the roles of transition metals and speciation in chemical biology. Curr. Opin. Chem. Biol. 2022, 69, 102155. [Google Scholar] [CrossRef]
  3. Thiele, D.J.; Gitlin, J.D. Assembling the pieces. Nat. Chem. Biol. 2008, 4, 145–147. [Google Scholar] [CrossRef]
  4. Hüttemann, M.; Helling, S.; Sanderson, T.H.; Sinkler, C.; Samavati, L.; Mahapatra, G.; Varughese, A.; Lu, G.; Liu, J.; Ramzan, R.; et al. Regulation of mitochondrial respiration and apoptosis through cell signaling: Cytochrome c oxidase and cytochrome c in ischemia/reperfusion injury and inflammation. Biochim. Biophys. Acta-Bioenerg. 2012, 1817, 598–609. [Google Scholar] [CrossRef]
  5. Wiggins, J.E.; Goyal, M.; Wharram, B.L.; Wiggins, R.C. Antioxidant Ceruloplasmin Is Expressed by Glomerular Parietal Epithelial Cells and Secreted into Urine in Association with Glomerular Aging and High-Calorie Diet. J. Am. Soc. Nephrol. 2006, 17, 1382–1387. [Google Scholar] [CrossRef]
  6. Harris, E.D. Copper as a Cofactor and Regulator of Copper, Zinc Superoxide Dismutase. J. Nutr. 1992, 122, 636–640. [Google Scholar] [CrossRef]
  7. Postma, G.C.; Nicastro, C.N.; Valdez, L.B.; Rukavina Mikusic, I.A.; Grecco, A.; Minatel, L. Decrease lysyl oxidase activity in hearts of copper-deficient bovines. J. Trace Elem. Med. Biol. 2021, 65, 126715. [Google Scholar] [CrossRef] [PubMed]
  8. Kumar, N. Copper Deficiency Myelopathy (Human Swayback). Mayo Clin. Proc. 2006, 81, 1371–1384. [Google Scholar] [CrossRef]
  9. Menkes, J.H. Menkes disease and Wilson disease: Two sides of the same copper coin Part II: Wilson disease. Eur. J. Paediatr. Neurol. 1999, 3, 245–253. [Google Scholar] [CrossRef] [PubMed]
  10. de Bie, P.; Muller, P.; Wijmenga, C.; Klomp, L.W.J. Molecular pathogenesis of Wilson and Menkes disease: Correlation of mutations with molecular defects and disease phenotypes. J. Med. Genet. 2007, 44, 673–688. [Google Scholar] [CrossRef] [PubMed]
  11. Scolari Grotto, F.; Glaser, V. Are high copper levels related to Alzheimer’s and Parkinson’s diseases? A systematic review and meta-analysis of articles published between 2011 and 2022. BioMetals 2023, 37, 3–22. [Google Scholar] [CrossRef]
  12. Barnham, K.J.; Bush, A.I. Metals in Alzheimer’s and Parkinson’s Diseases. Curr. Opin. Chem. Biol. 2008, 12, 222–228. [Google Scholar] [CrossRef] [PubMed]
  13. Izydorczyk, G.; Mikula, K.; Skrzypczak, D.; Moustakas, K.; Witek-Krowiak, A.; Chojnacka, K. Potential environmental pollution from copper metallurgy and methods of management. Environ. Res. 2021, 197, 111050. [Google Scholar] [CrossRef] [PubMed]
  14. Rehman, M.; Liu, L.; Wang, Q.; Saleem, M.H.; Bashir, S.; Ullah, S.; Peng, D. Copper environmental toxicology, recent advances, and future outlook: A review. Environ. Sci. Pollut. Res. 2019, 26, 18003–18016. [Google Scholar] [CrossRef]
  15. Mahamoud Ahmed, A.; Tardy, V.; Bonnineau, C.; Billard, P.; Pesce, S.; Lyautey, E. Changes in sediment microbial diversity following chronic copper-exposure induce community copper-tolerance without increasing sensitivity to arsenic. J. Hazard. Mater. 2020, 391, 122197. [Google Scholar] [CrossRef]
  16. Shrivas, K. Monitoring of copper level in water and soil samples by using liquid–liquid extraction. Environ. Monit. Assess. 2009, 168, 315–319. [Google Scholar] [CrossRef]
  17. Umbría-Salinas, K.; Valero, A.; Martins, S.E.; Wallner-Kersanach, M. Copper ecological risk assessment using DGT technique and PNEC: A case study in the Brazilian coast. J. Hazard. Mater. 2021, 403, 123918. [Google Scholar] [CrossRef]
  18. Narin, I. Determination of trace metal ions by AAS in natural water samples after preconcentration of pyrocatechol violet complexes on an activated carbon column. Talanta 2000, 52, 1041–1046. [Google Scholar] [CrossRef]
  19. Bulska, E.; Wagner, B. Quantitative aspects of inductively coupled plasma mass spectrometry. Philos. Trans. R. Soc. A 2016, 374, 20150369. [Google Scholar] [CrossRef]
  20. Chen, D.; Darabedian, N.; Li, Z.; Kai, T.; Jiang, D.; Zhou, F. An improved Bathocuproine assay for accurate valence identification and quantification of copper bound by biomolecules. Anal. Biochem. 2016, 497, 27–35. [Google Scholar] [CrossRef]
  21. Sulthana, S.F.; Iqbal, U.M.; Suseela, S.B.; Anbazhagan, R.; Chinthaginjala, R.; Chitathuru, D.; Ahmad, I.; Kim, T.-h. Electrochemical Sensors for Heavy Metal Ion Detection in Aqueous Medium: A Systematic Review. ACS Omega 2024, 9, 25493–25512. [Google Scholar] [CrossRef]
  22. Wainwright, P.; Wadey, D.; Cook, P. An inductively coupled plasma mass spectrometry method for relative free copper determination and generation of a paediatric reference interval. Ann. Clin. Biochem. 2017, 55, 485–490. [Google Scholar] [CrossRef] [PubMed]
  23. Ratnasekhar, C.; Laur, N.; Kinscherf, R.; Pomytkin, K.; Kaiser, L.; Knes, O.; Deigner, H.-P. ICP-MS trace element analysis in serum and whole blood. PLoS ONE 2020, 15, e0233357. [Google Scholar] [CrossRef]
  24. Cui, L.; Wu, J.; Li, J.; Ge, Y.; Ju, H. Electrochemical detection of Cu2+ through Ag nanoparticle assembly regulated by copper-catalyzed oxidation of cysteamine. Biosens. Bioelectron. 2014, 55, 272–277. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, C.; Dai, J.; Guo, S.; Sun, R.; Zhang, C.; Zhao, X.; Zhou, L.; Zhang, F.; Li, N.; Wang, M.; et al. Efficient photoelectrochemical sensor of Cu2+ based on ZnO-graphene nanocomposite sensitized with hexagonal CdS by calcination method. J. Electroanal. Chem. 2021, 893, 115330. [Google Scholar] [CrossRef]
  26. Dinu, A.; Bounegru, A.V.; Iticescu, C.; Georgescu, L.P.; Apetrei, C. Electrochemical Detection of Cd2+, Pb2+, Cu2+ and Hg2+ with Sensors Based on Carbonaceous Nanomaterials and Fe3O4 Nanoparticles. Nanomaterials 2024, 14, 702. [Google Scholar] [CrossRef]
  27. Zheng, X.; Cheng, W.; Ji, C.; Zhang, J.; Yin, M. Detection of metal ions in biological systems: A review. Rev. Anal. Chem. 2020, 39, 231–246. [Google Scholar] [CrossRef]
  28. Forzani, E.S.; Zhang, H.; Chen, W.; Tao, N. Detection of Heavy Metal Ions in Drinking Water Using a High-Resolution Differential Surface Plasmon Resonance Sensor. Environ. Sci. Technol. 2004, 39, 1257–1262. [Google Scholar] [CrossRef]
  29. Malitesta, C.; Guascito, M.R. Heavy metal determination by biosensors based on enzyme immobilised by electropolymerisation. Biosens. Bioelectron. 2005, 20, 1643–1647. [Google Scholar] [CrossRef]
  30. Zhu, X.; Xu, L.; Lou, Y.; Yu, H.; Li, X.; Blake, D.A.; Liu, F. Preparation of Specific Monoclonal Antibodies (MAbs) against Heavy Metals: MAbs That Recognize Chelated Cadmium Ions. J. Agric. Food Chem. 2007, 55, 7648–7653. [Google Scholar] [CrossRef]
  31. Lee, J.-S.; Han, M.S.; Mirkin, C.A. Colorimetric Detection of Mercuric Ion (Hg2+) in Aqueous Media using DNA-Functionalized Gold Nanoparticles. Angew. Chem. Int. Ed. 2007, 46, 4093–4096. [Google Scholar] [CrossRef]
  32. Li, T.; Shi, L.; Wang, E.; Dong, S. Silver-Ion-Mediated DNAzyme Switch for the Ultrasensitive and Selective Colorimetric Detection of Aqueous Ag+ and Cysteine. Chem. Eur. J. 2009, 15, 3347–3350. [Google Scholar] [CrossRef] [PubMed]
  33. Lehmann, M.; Riedel, K.; Adler, K.; Kunze, G. Amperometric measurement of copper ions with a deputy substrate using a novel Saccharomyces cerevisiae sensor. Biosens. Bioelectron. 2000, 15, 211–219. [Google Scholar] [CrossRef]
  34. Yan, L.; Bao, K.; Xu, X.; Li, L.; Wu, X. Recent progress in fluorescent probes for Cu2+ based on small organic molecules. J. Mol. Struct. 2024, 1316, 139100. [Google Scholar] [CrossRef]
  35. Hsiao, W.W.-W.; Pham, U.K.; Le, T.-N.; Lam, X.M.; Chiang, W.-H. Advances in aggregation-induced emission luminogens for biomedicine: From luminescence mechanisms to diagnostic applications. Biosens. Bioelectron. 2025, 270, 116942. [Google Scholar] [CrossRef]
  36. Saedi, Z.; Roushani, M.; Khaleghian-Moghadam, R.; Darabi, A. Selective and sensitive Detection of Cu2+ in aqueous solution based on cation exchange by Metal−Organic framework TMU-16 as a fluorescent sensor. J. Lumin. 2022, 251, 119165. [Google Scholar] [CrossRef]
  37. Fonseca, R.R.F.d.; Santos, G.F.S.d.; Rodrigues, J.G.A.; Ferreira, R.d.Q.; Luz, P.P. A functional FeIII-based NH2-MOF as an electrochemical sensor for Cu2+ detection in ethanol fuel. Ionics 2024, 30, 2397–2407. [Google Scholar] [CrossRef]
  38. Sargazi, S.; Fatima, I.; Hassan Kiani, M.; Mohammadzadeh, V.; Arshad, R.; Bilal, M.; Rahdar, A.; Díez-Pascual, A.M.; Behzadmehr, R. Fluorescent-based nanosensors for selective detection of a wide range of biological macromolecules: A comprehensive review. Int. J. Biol. Macromol. 2022, 206, 115–147. [Google Scholar] [CrossRef]
  39. Fang, X.; Zheng, Y.; Duan, Y.; Liu, Y.; Zhong, W. Recent Advances in Design of Fluorescence-Based Assays for High-Throughput Screening. Anal. Chem. 2018, 91, 482–504. [Google Scholar] [CrossRef]
  40. Jensen, G.C.; Janis, M.K.; Nguyen, H.N.; David, O.W.; Zastrow, M.L. Fluorescent Protein-Based Sensors for Detecting Essential Metal Ions across the Tree of Life. ACS Sens. 2024, 9, 1622–1643. [Google Scholar] [CrossRef]
  41. Wang, M.; Da, Y.; Tian, Y. Fluorescent proteins and genetically encoded biosensors. Chem. Soc. Rev. 2023, 52, 1189–1214. [Google Scholar] [CrossRef]
  42. Stepanenko, O.V.; Verkhusha, V.V.; Kuznetsova, I.M.; Uversky, V.N.; Turoverov, K.K. Fluorescent proteins as biomarkers and biosensors: Throwing color lights on molecular and cellular processes. Curr. Protein Pept. Sci. 2008, 9, 338–369. [Google Scholar] [CrossRef] [PubMed]
  43. Frei, M.S.; Mehta, S.; Zhang, J. Next-Generation Genetically Encoded Fluorescent Biosensors Illuminate Cell Signaling and Metabolism. Annu. Rev. Biophys. 2024, 53, 275–297. [Google Scholar] [CrossRef] [PubMed]
  44. Nam, K.H.; Xu, Y. Structural Analysis of the Large Stokes Shift Red Fluorescent Protein tKeima. Molecules 2024, 29, 2579. [Google Scholar] [CrossRef] [PubMed]
  45. Tian, F.; Xu, G.; Zhou, S.; Chen, S.; He, D. Principles and applications of green fluorescent protein-based biosensors: A mini-review. Analyst 2023, 148, 2882–2891. [Google Scholar] [CrossRef]
  46. Li, S.-A.; Meng, X.-Y.; Zhang, Y.-J.; Chen, C.-L.; Jiao, Y.-X.; Zhu, Y.-Q.; Liu, P.-P.; Sun, W. Progress in pH-Sensitive sensors: Essential tools for organelle pH detection, spotlighting mitochondrion and diverse applications. Front. Pharmacol. 2024, 14, 1339518. [Google Scholar] [CrossRef]
  47. Chin, M.Y.; Patwardhan, A.R.; Ang, K.-H.; Wang, A.L.; Alquezar, C.; Welch, M.; Nguyen, P.T.; Grabe, M.; Molofsky, A.V.; Arkin, M.R.; et al. Genetically Encoded, pH-Sensitive mTFP1 Biosensor for Probing Lysosomal pH. ACS Sens. 2021, 6, 2168–2180. [Google Scholar] [CrossRef]
  48. Tutol, J.N.; Ong, W.S.Y.; Phelps, S.M.; Peng, W.; Goenawan, H.; Dodani, S.C. Engineering the ChlorON Series: Turn-On Fluorescent Protein Sensors for Imaging Labile Chloride in Living Cells. ACS Cent. Sci. 2023, 10, 77–86. [Google Scholar] [CrossRef]
  49. Yu, T.-G.; Lee, J.; Yoon, J.; Choi, J.M.; Kim, D.-G.; Heo, W.D.; Song, J.-J.; Kim, H.-S. Engineering of a Fluorescent Protein for a Sensing of an Intrinsically Disordered Protein through Transition in the Chromophore State. JACS Au 2023, 3, 3055–3065. [Google Scholar] [CrossRef]
  50. Torres-Ocampo, A.P.; Palmer, A.E. Genetically encoded fluorescent sensors for metals in biology. Curr. Opin. Chem. Biol. 2023, 74, 102284. [Google Scholar] [CrossRef]
  51. Rahimi, Y.; Goulding, A.; Shrestha, S.; Mirpuri, S.; Deo, S.K. Mechanism of copper induced fluorescence quenching of red fluorescent protein, DsRed. Biochem. Biophys. Res. Commun. 2008, 370, 57–61. [Google Scholar] [CrossRef]
  52. Kim, I.J.; Xu, Y.; Nam, K.H. Metal-Induced Fluorescence Quenching of Photoconvertible Fluorescent Protein DendFP. Molecules 2022, 27, 2922. [Google Scholar] [CrossRef]
  53. Choi, Y.-A.; Keem, J.O.; Kim, C.Y.; Yoon, H.R.; Heo, W.D.; Chung, B.H.; Jung, Y. A novel copper-chelating strategy for fluorescent proteins to image dynamic copper fluctuations on live cell surfaces. Chem. Sci. 2015, 6, 1301–1307. [Google Scholar] [CrossRef] [PubMed]
  54. Taki, M.; Iyoshi, S.; Ojida, A.; Hamachi, I.; Yamamoto, Y. Development of Highly Sensitive Fluorescent Probes for Detection of Intracellular Copper(I) in Living Systems. J. Am. Chem. Soc. 2010, 132, 5938–5939. [Google Scholar] [CrossRef] [PubMed]
  55. Shimomura, O.; Johnson, F.H.; Saiga, Y. Extraction, Purification and Properties of Aequorin, a Bioluminescent Protein from the Luminous Hydromedusan, Aequorea. J. Cell. Comp. Physiol. 1962, 59, 223–239. [Google Scholar] [CrossRef] [PubMed]
  56. Zimmer, M. Green fluorescent protein (GFP): Applications, structure, and related photophysical behavior. Chem. Rev. 2002, 102, 759–781. [Google Scholar] [CrossRef]
  57. Tsien, R.Y. The green fluorescent protein. Annu. Rev. Biochem. 1998, 67, 509–544. [Google Scholar] [CrossRef]
  58. Piatkevich, K.D.; Verkhusha, V.V. Advances in engineering of fluorescent proteins and photoactivatable proteins with red emission. Curr. Opin. Chem. Biol. 2010, 14, 23–29. [Google Scholar] [CrossRef]
  59. Seo, P.-W.; Kim, G.-J.; Kim, J.-S. A short guide on blue fluorescent proteins: Limits and perspectives. Appl. Microbiol. Biotechnol. 2024, 108, 208. [Google Scholar] [CrossRef]
  60. Day, R.N.; Davidson, M.W. The fluorescent protein palette: Tools for cellular imaging. Chem. Soc. Rev. 2009, 38, 2887–2921. [Google Scholar] [CrossRef]
  61. Cheng, L.; Fu, J.; Tsukamoto, A.; Hawley, R.G. Use of green fluorescent protein variants to monitor gene transfer and expression in mammalian cells. Nat. Biotechnol. 1996, 14, 606–609. [Google Scholar] [CrossRef]
  62. Ogawa, H.; Inouye, S.; Tsuji, F.I.; Yasuda, K.; Umesono, K. Localization, trafficking, and temperature-dependence of the Aequorea green fluorescent protein in cultured vertebrate cells. Proc. Natl. Acad. Sci. USA 1995, 92, 11899–11903. [Google Scholar] [CrossRef] [PubMed]
  63. Mohr, M.A.; Argast, P.; Pantazis, P. Labeling cellular structures in vivo using confined primed conversion of photoconvertible fluorescent proteins. Nat. Protoc. 2016, 11, 2419–2431. [Google Scholar] [CrossRef] [PubMed]
  64. Piston, D.W.; Kremers, G.-J. Fluorescent protein FRET: The good, the bad and the ugly. Trends Biochem. Sci. 2007, 32, 407–414. [Google Scholar] [CrossRef] [PubMed]
  65. Kenworthy, A.K. Imaging Protein-Protein Interactions Using Fluorescence Resonance Energy Transfer Microscopy. Methods 2001, 24, 289–296. [Google Scholar] [CrossRef]
  66. dos Santos, N.V.; Saponi, C.F.; Ryan, T.M.; Primo, F.L.; Greaves, T.L.; Pereira, J.F.B. Reversible and irreversible fluorescence activity of the Enhanced Green Fluorescent Protein in pH: Insights for the development of pH-biosensors. Int. J. Biol. Macromol. 2020, 164, 3474–3484. [Google Scholar] [CrossRef]
  67. Zhang, C.; Liu, M.-S.; Xing, X.-H. Temperature influence on fluorescence intensity and enzyme activity of the fusion protein of GFP and hyperthermophilic xylanase. Appl. Microbiol. Biotechnol. 2009, 84, 511–517. [Google Scholar] [CrossRef]
  68. Nam, K.H. Fluorescent Protein-Based Metal Biosensors. Chemosensors 2023, 11, 216. [Google Scholar] [CrossRef]
  69. Sun, N.; Malide, D.; Liu, J.; Rovira, I.I.; Combs, C.A.; Finkel, T. A fluorescence-based imaging method to measure in vitro and in vivo mitophagy using mt-Keima. Nat. Protoc. 2017, 12, 1576–1587. [Google Scholar] [CrossRef]
  70. Xu, Y.; Seo, Y.G.; Kim, I.J.; Nam, K.H. pH-Induced Conformational Change of the Chromophore of the Large Stokes Shift Fluorescent Protein tKeima. Molecules 2025, 30, 1623. [Google Scholar] [CrossRef]
  71. Bae, J.-E.; Kim, I.J.; Nam, K.H. Disruption of the hydrogen bonding network determines the pH-induced non-fluorescent state of the fluorescent protein ZsYellow by protonation of Glu221. Biochem. Biophys. Res. Commun. 2017, 493, 562–567. [Google Scholar] [CrossRef] [PubMed]
  72. Richmond, T.A.; Takahashi, T.T.; Shimkhada, R.; Bernsdorf, J. Engineered Metal Binding Sites on Green Fluorescence Protein. Biochem. Biophys. Res. Commun. 2000, 268, 462–465. [Google Scholar] [CrossRef] [PubMed]
  73. Eli, P.; Chakrabartty, A. Variants of DsRed fluorescent protein: Development of a copper sensor. Protein Sci. 2006, 15, 2442–2447. [Google Scholar] [CrossRef]
  74. Sumner, J.P.; Westerberg, N.M.; Stoddard, A.K.; Hurst, T.K.; Cramer, M.; Thompson, R.B.; Fierke, C.A.; Kopelman, R. DsRed as a highly sensitive, selective, and reversible fluorescence-based biosensor for both Cu+ and Cu2+ ions. Biosens. Bioelectron. 2006, 21, 1302–1308. [Google Scholar] [CrossRef] [PubMed]
  75. Kim, I.J.; Kim, S.; Park, J.; Eom, I.; Kim, S.; Kim, J.H.; Ha, S.C.; Kim, Y.G.; Hwang, K.Y.; Nam, K.H. Crystal structures of Dronpa complexed with quenchable metal ions provide insight into metal biosensor development. FEBS Lett. 2016, 590, 2982–2990. [Google Scholar] [CrossRef]
  76. Bae, J.E.; Kim, I.J.; Nam, K.H. Spectroscopic Analysis of the Cu2+-Induced Fluorescence Quenching of Fluorescent Proteins AmCyan and mOrange2. Mol. Biotechnol. 2018, 60, 485–491. [Google Scholar] [CrossRef]
  77. Kim, I.J.; Xu, Y.; Nam, K.H. Spectroscopic Analysis of Fe Ion-Induced Fluorescence Quenching of the Green Fluorescent Protein ZsGreen. J. Fluoresc. 2021, 31, 307–314. [Google Scholar] [CrossRef]
  78. Kim, I.J.; Xu, Y.; Nam, K.H. Spectroscopic and Structural Analysis of Cu2+-Induced Fluorescence Quenching of ZsYellow. Biosensors 2020, 10, 29. [Google Scholar] [CrossRef]
  79. Barondeau, D.P.; Kassmann, C.J.; Tainer, J.A.; Getzoff, E.D. Structural chemistry of a green fluorescent protein Zn biosensor. J. Am. Chem. Soc. 2002, 124, 3522–3524. [Google Scholar] [CrossRef]
  80. Yu, X.; Strub, M.P.; Barnard, T.J.; Noinaj, N.; Piszczek, G.; Buchanan, S.K.; Taraska, J.W. An engineered palette of metal ion quenchable fluorescent proteins. PLoS ONE 2014, 9, e95808. [Google Scholar] [CrossRef]
  81. Isarankura-Na-Ayudhya, C.; Tantimongcolwat, T.; Galla, H.J.; Prachayasittikul, V. Fluorescent protein-based optical biosensor for copper ion quantitation. Biol. Trace Elem. Res. 2010, 134, 352–363. [Google Scholar] [CrossRef]
  82. Ababou, A.; Bombarda, E. On the involvement of electron transfer reactions in the fluorescence decay kinetics heterogeneity of proteins. Protein Sci. 2008, 10, 2102–2113. [Google Scholar] [CrossRef]
  83. Mizuno, T.; Murao, K.; Tanabe, Y.; Oda, M.; Tanaka, T. Metal-ion-dependent GFP emission in vivo by combining a circularly permutated green fluorescent protein with an engineered metal-ion-binding coiled-coil. J. Am. Chem. Soc. 2007, 129, 11378–11383. [Google Scholar] [CrossRef] [PubMed]
  84. Rahimi, Y.; Shrestha, S.; Deo, S.K. Metal Affinity-Based Purification of a Red Fluorescent Protein. Chromatographia 2007, 65, 429–433. [Google Scholar] [CrossRef]
  85. Fischer, J.; Renn, D.; Quitterer, F.; Radhakrishnan, A.; Liu, M.; Makki, A.; Ghorpade, S.; Rueping, M.; Arold, S.T.; Groll, M.; et al. Robust and Versatile Host Protein for the Design and Evaluation of Artificial Metal Centers. ACS Catal. 2019, 9, 11371–11380. [Google Scholar] [CrossRef]
  86. Jeschek, M.; Panke, S.; Ward, T.R. Artificial Metalloenzymes on the Verge of New-to-Nature Metabolism. Trends Biotechnol. 2018, 36, 60–72. [Google Scholar] [CrossRef]
  87. Davis, H.J.; Ward, T.R. Artificial Metalloenzymes: Challenges and Opportunities. ACS Cent. Sci. 2019, 5, 1120–1136. [Google Scholar] [CrossRef]
  88. Yu, X.; Wu, X.; Bermejo, G.A.; Brooks, B.R.; Taraska, J.W. Accurate High-Throughput Structure Mapping and Prediction with Transition Metal Ion FRET. Structure 2013, 21, 9–19. [Google Scholar] [CrossRef]
  89. Ando, R.; Mizuno, H.; Miyawaki, A. Regulated Fast Nucleocytoplasmic Shuttling Observed by Reversible Protein Highlighting. Science 2004, 306, 1370–1373. [Google Scholar] [CrossRef]
  90. Habuchi, S.; Ando, R.; Dedecker, P.; Verheijen, W.; Mizuno, H.; Miyawaki, A.; Hofkens, J. Reversible single-molecule photoswitching in the GFP-like fluorescent protein Dronpa. Proc. Natl. Acad. Sci. USA 2005, 102, 9511–9516. [Google Scholar] [CrossRef]
  91. Lee, H.; DeLoache, W.C.; Dueber, J.E. Spatial organization of enzymes for metabolic engineering. Metab. Eng. 2012, 14, 242–251. [Google Scholar] [CrossRef]
  92. Good, M.C.; Zalatan, J.G.; Lim, W.A. Scaffold Proteins: Hubs for Controlling the Flow of Cellular Information. Science 2011, 332, 680–686. [Google Scholar] [CrossRef]
  93. Lai, Y.-T.; King, N.P.; Yeates, T.O. Principles for designing ordered protein assemblies. Trends Cell Biol. 2012, 22, 653–661. [Google Scholar] [CrossRef] [PubMed]
  94. Laganowsky, A.; Zhao, M.; Soriaga, A.B.; Sawaya, M.R.; Cascio, D.; Yeates, T.O. An approach to crystallizing proteins by metal-mediated synthetic symmetrization. Protein Sci. 2011, 20, 1876–1890. [Google Scholar] [CrossRef] [PubMed]
  95. Leibly, D.J.; Arbing, M.A.; Pashkov, I.; DeVore, N.; Waldo, G.S.; Terwilliger, T.C.; Yeates, T.O. A Suite of Engineered GFP Molecules for Oligomeric Scaffolding. Structure 2015, 23, 1754–1768. [Google Scholar] [CrossRef]
  96. Taylor, G.L. Introduction to phasing. Acta Crystallogr. D Struct. Biol. 2010, 66, 325–338. [Google Scholar] [CrossRef] [PubMed]
  97. Mehr, A.; Henneberg, F.; Chari, A.; Görlich, D.; Huyton, T. The copper(II)-binding tripeptide GHK, a valuable crystallization and phasing tag for macromolecular crystallography. Acta Crystallogr. D Struct. Biol. 2020, 76, 1222–1232. [Google Scholar] [CrossRef]
  98. Rubino, J.T.; Franz, K.J. Coordination chemistry of copper proteins: How nature handles a toxic cargo for essential function. J. Inorg. Biochem. 2012, 107, 129–143. [Google Scholar] [CrossRef]
  99. Ashraf, J.; Riaz, M.A. Biological potential of copper complexes: A review. Turk. J. Chem. 2022, 46, 595–623. [Google Scholar] [CrossRef]
  100. Nam, K.H. Evaluation of AlphaFold3 prediction for post-translational modification, oligomeric assembly, and quenchable metal binding of fluorescent proteins. J. Mol. Graph. Model. 2025, 142, 109169. [Google Scholar] [CrossRef]
Figure 1. Traditional metal ion analytical techniques and potential fluorescent protein (FP)-based metal ion biosensors. Some images were obtained from BioRender (https://biorender.com/: accessed on 8 March 2025).
Figure 1. Traditional metal ion analytical techniques and potential fluorescent protein (FP)-based metal ion biosensors. Some images were obtained from BioRender (https://biorender.com/: accessed on 8 March 2025).
Biosensors 15 00675 g001
Figure 2. Changes in the fluorescence emission intensity of DendFP (green state) in the presence of metal ions. (a) The fluorescence emission of DendFP was significantly quenched by Fe2+, Fe3+, and Cu2+. (b) Relative fluorescence emission of DendFP measured in the presence of various metal ions. The original figures were adapted and modified from a previous study [52].
Figure 2. Changes in the fluorescence emission intensity of DendFP (green state) in the presence of metal ions. (a) The fluorescence emission of DendFP was significantly quenched by Fe2+, Fe3+, and Cu2+. (b) Relative fluorescence emission of DendFP measured in the presence of various metal ions. The original figures were adapted and modified from a previous study [52].
Biosensors 15 00675 g002
Figure 3. Cu2+ binding to mTFPCHH (PDB code: 4R6D). (a) Cartoon representation presenting two Cu2+ ions bound to mTFPCHH. (b) Coordination of the Cu2+ ion at Cu1, which is located near the α-helix connected to the chromophore of mTFPCHH. (c) Coordination of the Cu2+ ion at Cu2, which is located in the N-terminal region of mTFP*. (d) Structural comparison of the Cu1 binding site between Cu2+-bound mTFPCHH and apo-mTFPCHH (PDB code: 6QSL). (e) B-factor putty representation of the loop between Leu198 and His204 in apo-mTFPCHH and Cu2+-bound mTFPCHH.
Figure 3. Cu2+ binding to mTFPCHH (PDB code: 4R6D). (a) Cartoon representation presenting two Cu2+ ions bound to mTFPCHH. (b) Coordination of the Cu2+ ion at Cu1, which is located near the α-helix connected to the chromophore of mTFPCHH. (c) Coordination of the Cu2+ ion at Cu2, which is located in the N-terminal region of mTFP*. (d) Structural comparison of the Cu1 binding site between Cu2+-bound mTFPCHH and apo-mTFPCHH (PDB code: 6QSL). (e) B-factor putty representation of the loop between Leu198 and His204 in apo-mTFPCHH and Cu2+-bound mTFPCHH.
Biosensors 15 00675 g003
Figure 4. Analysis of metal ion binding to Dronpa. (a) Cartoon representation of two Cu2+ bound to Dronpa (PDB code: 5HZT). (b) Superimposition of the Cu1 and Cu2 binding sites from 24 Dronpa molecules in the asymmetric unit. (c) Close-up view of representative Cu1 and Cu2 binding sites in Dronpa. (d) Superimposition of the wild-type and Cu2+-bound Dronpa structures. (e) Cartoon representation of two Co2+ bound to Dronpa (PDB code: 5HZS). (f) Superimposition of the Co1 and Co2 binding sites from 24 Dronpa molecules in the asymmetric unit. (g) Close-up view of representative Co1 and Co2 binding sites in Dronpa. (h) Superimposition of the wild-type and Co2+-bound Dronpa structures. (i) Cartoon representation of two Ni2+ bound to Dronpa (PDB code: 5HZU). (j) Superimposition of the Ni1 and Ni2 binding sites from 24 Dronpa molecules in the asymmetric unit. (k) Close-up view of representative Ni1 and Ni2 binding sites in Dronpa. (l) Superimposition of the wild-type and Ni2+-bound Dronpa structures. The conformational change in the histidine side chain upon metal ion binding is indicated by an arrow.
Figure 4. Analysis of metal ion binding to Dronpa. (a) Cartoon representation of two Cu2+ bound to Dronpa (PDB code: 5HZT). (b) Superimposition of the Cu1 and Cu2 binding sites from 24 Dronpa molecules in the asymmetric unit. (c) Close-up view of representative Cu1 and Cu2 binding sites in Dronpa. (d) Superimposition of the wild-type and Cu2+-bound Dronpa structures. (e) Cartoon representation of two Co2+ bound to Dronpa (PDB code: 5HZS). (f) Superimposition of the Co1 and Co2 binding sites from 24 Dronpa molecules in the asymmetric unit. (g) Close-up view of representative Co1 and Co2 binding sites in Dronpa. (h) Superimposition of the wild-type and Co2+-bound Dronpa structures. (i) Cartoon representation of two Ni2+ bound to Dronpa (PDB code: 5HZU). (j) Superimposition of the Ni1 and Ni2 binding sites from 24 Dronpa molecules in the asymmetric unit. (k) Close-up view of representative Ni1 and Ni2 binding sites in Dronpa. (l) Superimposition of the wild-type and Ni2+-bound Dronpa structures. The conformational change in the histidine side chain upon metal ion binding is indicated by an arrow.
Biosensors 15 00675 g004
Figure 5. Crystal structure of Cu2+-bound split-GFP. (a) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W6T). (b) Cu2+-mediated interaction between the split-GFP molecules (PDB code: 4W6T) within the crystal lattice. (c) Close-up view of the Cu2+-mediated interface between split-GFPs (PDB code: 4W6T). (d) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7D). (e) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7D). Close-up views of (f) the Cu1 and Cu2 sites and (g) the Cu3 site revealing the Cu2+-mediated interfaces between split-GFPs (PDB code: 4W7D). (h) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7E). (i) Close-up view of the Cu2+-binding site (PDB code: 4W7E). The Cu2+ ion did not interact with the neighboring molecules. IMD, imidazole. (j) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7R). (k) Close-up view of the Cu2+-binding site (PDB code: 4W7R). (l) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7F). (m) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7F). (n) Close-up view of the Cu2+-binding site (PDB code: 4W7F).
Figure 5. Crystal structure of Cu2+-bound split-GFP. (a) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W6T). (b) Cu2+-mediated interaction between the split-GFP molecules (PDB code: 4W6T) within the crystal lattice. (c) Close-up view of the Cu2+-mediated interface between split-GFPs (PDB code: 4W6T). (d) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7D). (e) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7D). Close-up views of (f) the Cu1 and Cu2 sites and (g) the Cu3 site revealing the Cu2+-mediated interfaces between split-GFPs (PDB code: 4W7D). (h) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7E). (i) Close-up view of the Cu2+-binding site (PDB code: 4W7E). The Cu2+ ion did not interact with the neighboring molecules. IMD, imidazole. (j) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7R). (k) Close-up view of the Cu2+-binding site (PDB code: 4W7R). (l) Cartoon representation of Cu2+-bound split-GFP (PDB code: 4W7F). (m) Cu2+-mediated interaction between split-GFP molecules within the crystal lattice (PDB code: 4W7F). (n) Close-up view of the Cu2+-binding site (PDB code: 4W7F).
Biosensors 15 00675 g005
Figure 6. Crystal structure of Cu2+ bound to the GHK tripeptide fused at the N-terminus of GFP. (a) Cu2+ binding to the GHK tripeptide in the GFP structure with the P6522 crystal form (PDB code: 6QUJ). (b) Interaction of Cu2+ bound to GHK-GFP with the neighboring molecule. (c) Close-up view of the Cu2+ coordination in GHK-GFP-P6522, including interaction with the neighboring molecule. (d) Cu2+ binding to the GHK tripeptide in the GFP structure with the P212121 crystal form (PDB code: 6QUH). (e) Close-up view of the Cu2+ coordination in GHK-GFP-P212121, including interaction with the neighboring molecule. (f) Close-up view of the Cu2 coordination in GHK-GFP-P212121.
Figure 6. Crystal structure of Cu2+ bound to the GHK tripeptide fused at the N-terminus of GFP. (a) Cu2+ binding to the GHK tripeptide in the GFP structure with the P6522 crystal form (PDB code: 6QUJ). (b) Interaction of Cu2+ bound to GHK-GFP with the neighboring molecule. (c) Close-up view of the Cu2+ coordination in GHK-GFP-P6522, including interaction with the neighboring molecule. (d) Cu2+ binding to the GHK tripeptide in the GFP structure with the P212121 crystal form (PDB code: 6QUH). (e) Close-up view of the Cu2+ coordination in GHK-GFP-P212121, including interaction with the neighboring molecule. (f) Close-up view of the Cu2 coordination in GHK-GFP-P212121.
Biosensors 15 00675 g006
Figure 7. Visual schematic summarizing Cu2+ binding in fluorescent proteins (FPs). (a) Superimposition of Cu2+ binding sites in mTFPCHH, Dronpa, split-GFP, and GHK-GFP. The bound Cu2+ ions in mTFPCHH, Dronpa, split-GFP, and GHK-GFP are shown as green, pink, blue, and yellow spheres, respectively. (b) Representative tetrahedral or square planar coordination geometries of Cu2+ observed in FPs. (c) Example of an unidentified Cu2+ coordination geometry in FPs.
Figure 7. Visual schematic summarizing Cu2+ binding in fluorescent proteins (FPs). (a) Superimposition of Cu2+ binding sites in mTFPCHH, Dronpa, split-GFP, and GHK-GFP. The bound Cu2+ ions in mTFPCHH, Dronpa, split-GFP, and GHK-GFP are shown as green, pink, blue, and yellow spheres, respectively. (b) Representative tetrahedral or square planar coordination geometries of Cu2+ observed in FPs. (c) Example of an unidentified Cu2+ coordination geometry in FPs.
Biosensors 15 00675 g007
Table 1. Comparison of the Cu2+-induced fluorescence quenching properties of FPs.
Table 1. Comparison of the Cu2+-induced fluorescence quenching properties of FPs.
FPQuenching
Efficiency (%)
Kd (μM)Recovery (EDTA)Limit of Detection (LOD)Reference
DsRed90.0 ± 100.54ND45 ± 2 nM[74]
drFP58378.014.80 ± 1.68>90% (1 mM)ND[73]
Rmu1366.010.90 ± 1.74NDND[73]
Dronpa86.0NDNDND[75]
AmCyan80.056.1089.4% (5 mM)ND[76]
mOrange289.021.46>100% (5 mM)ND[76]
ZsYellow81.4NDNDND[78]
ZsGreen77.268.2>90% (50 mM)ND[77]
DendFP98.8137.1844.2% (50 mM)3.2 μM[52]
ND: not discussed.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nam, K.H. Structure-Based Understanding of Cu2+ Coordination in Fluorescent Proteins for Metal Biosensor Applications—A Review. Biosensors 2025, 15, 675. https://doi.org/10.3390/bios15100675

AMA Style

Nam KH. Structure-Based Understanding of Cu2+ Coordination in Fluorescent Proteins for Metal Biosensor Applications—A Review. Biosensors. 2025; 15(10):675. https://doi.org/10.3390/bios15100675

Chicago/Turabian Style

Nam, Ki Hyun. 2025. "Structure-Based Understanding of Cu2+ Coordination in Fluorescent Proteins for Metal Biosensor Applications—A Review" Biosensors 15, no. 10: 675. https://doi.org/10.3390/bios15100675

APA Style

Nam, K. H. (2025). Structure-Based Understanding of Cu2+ Coordination in Fluorescent Proteins for Metal Biosensor Applications—A Review. Biosensors, 15(10), 675. https://doi.org/10.3390/bios15100675

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