Skip to main content
ACS Medicinal Chemistry Letters logoLink to ACS Medicinal Chemistry Letters
. 2024 Jul 25;15(8):1376–1385. doi: 10.1021/acsmedchemlett.4c00283

Fluorescent Peptides Sequester Redox Copper to Mitigate Oxidative Stress, Amyloid Toxicity, and Neuroinflammation

Sabyasachi Mandal , Yelisetty Venkata Suseela , Sourav Samanta , Bertrand Vileno , Peter Faller ‡,*, Thimmaiah Govindaraju †,*
PMCID: PMC11318102  PMID: 39140073

Abstract

graphic file with name ml4c00283_0008.jpg

Alzheimer’s disease is a progressive neurodegenerative disorder that significantly contributes to dementia. The lack of effective therapeutic interventions presents a significant challenge to global health. We have developed a set of short peptides (PNGln) conjugated with a dual-functional fluorophoric amino acid (NGln). The lead peptide, P2NGln, displays a high affinity for Cu2+, maintaining the metal ion in a redox-inactive state. This mitigates the cytotoxicity generated by reactive oxygen species (ROS), which are produced by Cu2+ under the reductive conditions of Asc and Aβ16 or Aβ42. Furthermore, P2NGln inhibits both Cu-dependent and -independent fibrillation of Aβ42, along with the subsequent toxicity induced by Aβ42. In addition, P2NGln exhibits inhibitory effects on the production of lipopolysaccharide (LPS)-induced ROS and reactive nitrogen species (RNS) in microglial cells. In vitro and cellular studies indicate that P2NGln could significantly reduce Aβ–Cu2+-induced ROS production, amyloid toxicity, and neuroinflammation, offering an innovative strategy against Alzheimer’s disease.

Keywords: Fluorescent peptides, Copper, Oxidative stress, Amyloid toxicity, Neuroinflammation


Alzheimer’s disease (AD) is a chronic neurological disorder characterized by the progressive decline of cognitive functions, accompanied by psychiatric symptoms, posing a significant challenge to global health.1,2 Currently, AD affects an estimated 50 million individuals, with projections suggesting a surge to 152 million by 2050. The pathology of AD is marked by two primary features: the buildup of extracellular amyloid beta (Aβ) plaques and intracellular neurofibrillary tangles (NFTs) composed of abnormal tau protein.24 The complex interplay between copper and the regulation of amyloid precursor protein (APP) in AD has been a subject of investigation.5,6 Studies indicate that Cu2+ may increase APP expression, potentially leading to a rise in Aβ peptide production in the brain.7 This overproduction is especially noted in ATP-7A-deficient fibroblast cells from mice and humans, where elevated Cu2+ levels stimulate Aβ peptide production.2,8,9 Aβ peptides undergo several conformational changes, eventually forming toxic oligomers, protofibrils, and fibrillar aggregates with an ordered β-sheet structure.10,11 The N-terminal region of Aβ (Aβ16: D1AEFRHDSGYEVHHQK16) contains binding sites for Cu2+, and this region does not directly contribute to amyloid aggregation. Structural elucidation studies have identified specific residues such as histidine (H6, H13, and H14) and aspartic acid (D1) to be responsible for the formation of the Aβ–Cu2+ complex.1215 This complexation accelerates Aβ aggregation, producing highly toxic species that cause neurodegenerative effects, including membrane and mitochondrial damage.1 The redox activity of the Aβ–Cu2+ complex can generate ROS catalytically in the presence of the physiological reducing agent ascorbate, damaging biomolecules (lipids, proteins, and DNA) and cellular components, leading to neuronal injury and death.16,17 Understanding and modulation of both Cu-dependent and -independent pathways of Aβ aggregation are vital for combating amyloid toxicity in AD.1820 Furthermore, the interplay between ROS and neuroinflammation is evident in AD, with activated microglia and astrocytes releasing ROS and RNS, exacerbating the oxidative stress environment in the brain.21,22 This oxidative stress contributes to increased mitochondrial ROS and mitochondrial damage.23

Small peptides have been used in the past in order to sequester and redox-silence Cu from Aβ to suppress the ROS formation.2428 In our previous study, we reported natural tripeptide-based peptidomimetic molecules that complex Cu2+ and keep it in a redox-dormant state to alleviate Cu-associated toxicity in AD.22,27,29 Inspired by these findings, we designed and synthesized a set of small peptides functionalized with a naphthalene monoimide (NMI)-based fluorophore (PNGln) to monitor and alleviate Cu-associated toxicity (Figure 1a). NMI is known for its anti-amyloidogenic and fluorescence properties, while the peptide fragment is designed for Cu2+ sequestration.27,30,31 Thus, the designed fluorescent peptides are anticipated to concurrently report and capture Cu2+, arrest its ROS-producing redox cycle, and inhibit amyloid aggregation. These peptides are promising candidates due to the ease of structural tunability to enhance cell penetration ability and pharmaceutical efficacy. A critical design consideration for these peptide conjugates (PNGln) was the preference to complex Cu2+ over Zn2+, as the latter’s 10-fold higher synaptic concentration and higher disponibility could hinder Cu2+ capture.

Figure 1.

Figure 1

(a) Structures of fluorescent peptides (PNGln): P1NGln (GGHKNGln), P2NGln (GHHKNGln), P3NGln (HKNGln), and P4NGln (GHKNGln). (b) Schematic illustration of tripartite activity of fluorescent peptides PNGln: sequestering Cu from Aβ–Cu2+ complex, ameliorating amyloid toxicity and oxidative stress, and suppressing neuroinflammation (created with BioRender).

The peptide conjugates (PNGln) displayed a marked preference for Cu2+ over Fe3+, Zn2+, and Al3+. The ability of P2NGln to sequester and form stable, redox-inactive complexes with Cu2+ led to a substantial reduction in ROS production triggered by the redox-active Aβ16–Cu2+ complex. Our cellular experiments showed that these peptide conjugates are nontoxic and significantly reduce ROS levels. Notably, the peptide P2NGln was most effective in sequestering Cu2+ from the Aβ16–Cu2+ complex, preventing Aβ fibrillation and thus reducing neuronal toxicity. P2NGln also penetrated cells and decreased neuroinflammation by reducing lipopolysaccharide (LPS)-induced ROS and RNS in N9 microglial cells, and it helped to restore mitochondrial function and lower mitochondrial ROS production.

In this Letter, we present glycine-histidine-lysine (GHK)-inspired fluorescent peptide conjugates (PNGln) as potent Cu2+ sequesterers capable of mitigating ROS, oxidative stress, Cu-dependent and -independent amyloid toxicity, and neuroinflammation (Figure 1b). Our findings underscore the potential therapeutic implications of these fluorescent peptide conjugates to ameliorate Cu-associated multifaceted toxicity in AD.

The accumulation of Cu in amyloid plaques suggests its role in exacerbating the pathogenesis of AD in the brain. Therefore, sequestering Cu2+ from the Aβ–Cu2+ complex emerges as a potential strategy to investigate the role of Cu2+ in amplifying Aβ toxicity.32,33 We designed four molecules by conjugating Cu-binding peptides with a fluorophore component. The fluorophore component consists of NMI-fluorophore-functionalized glutamine (NGln). The choice of NMI as the fluorophoric moiety stems from the fact that it can serve as a modulator of Aβ aggregation as well as reporter of metal ion binding in proximity. The tripeptide GHK, a natural peptide found in human plasma, saliva, and urine, has been recognized as a potent Cu-chelating ligand.34 GHK has a strong affinity for Cu2+ and readily forms the GHK–Cu complex.34,35 The formation of the GHK–Cu2+ complex is mediated by the α-amino group of Gly (G), the amide nitrogen of the Gly-His (G-H) peptide bond, and the imidazole side chain of His (H).36 The XXH motif is a better redox silencer and stronger ligand compared to XH. XHH, which contains both motifs, is as strong as XXH but binds Cu2+ also faster than XXH (more like XH);29,37,38 to complete the series, H at position 1 was also used. Based on the understandings from our previous studies, we considered HK, GHK, GHHK, and GGHK as Cu-binding peptides.22,27,36,37 Naphthalene anhydride was condensed with ethylenediamine, which was coupled to glutamine to obtain the fluorescent amino acid conjugate (NGln). NGln was conjugated to HK, GHK, GHHK, and GGHK. The N-terminal glycine and histidine residues primarily facilitate Cu chelation, while the C-terminal Lys does not participate in the Cu–GHK complexation. NGln was conjugated at the C-terminus of peptides to avoid any direct interference in the Cu chelation. Following our rational design strategy, we synthesized four PNGln conjugates, namely, P1NGln (GGHKNGln), P2NGln (GHHKNGln), P3NGln (HKNGln), and P4NGln (GHKNGln) (Figure 1a).

The PNGln conjugates were prepared in two stages: synthesis of (i) NGln by a solution-phase protocol (Scheme S1) and (ii) copper-binding peptides using a solid-phase protocol (Scheme S2). In the solid-phase synthesis of peptides, Rink amide resin beads were soaked in DCM and DMF, followed by treatment with piperidine at 20% in DMF to deprotect Fmoc protective groups of surface amines. NGln was attached to the resin using HBTU, HOBt, and DIPEA in DMF as coupling reagents. The Kaiser test confirmed the complete coupling of NGln to the resin, which was then treated with piperidine (20% in DMF) to release the free amine of the NGln unit. Subsequently, Lys, His, and Gly were added using the Fmoc-chemistry peptide synthesis protocols as shown in Scheme S2. Finally, the resin was treated with trifluoroacetic acid (TFA) to obtain P1NGln, P2NGln, P3NGln, and P4NGln. All the intermediates, NGln, and PNGln were characterized by NMR spectroscopy and mass spectrometry (LCMS, HRMS and MALDI-TOF) analysis.

The ability of peptides to bind metal ions was assessed by various spectroscopic techniques. The absorption spectra revealed an intense band at 450 nm, signifying the incorporation of the NGln moiety in the peptides. The PNGln showed a strong emission band in the green spectral region, with emission maximum (λmax) around 550 nm upon excitation at 450 nm. Numerous studies have established the role of various metal ions (Cu2+, Fe3+, Zn2+, and Al3+) to aggravate amyloid toxicity through metal-complexation-induced acceleration of Aβ aggregation.22,27,39 The selective sequestering of these metal ions from Aβ–metal complexes could provide valuable insights into their role in amyloid toxicity. We treated a homogeneous solution of PNGln at 50 μM independently with Cu2+, Zn2+, Fe3+, and Al3+ in a 1:1 stoichiometric ratio and recorded fluorescence emission (λex = 450 nm, λem = 550 nm). The fluorescence of PNGln was selectively reduced in the presence of Cu2+, while other metal ions showed minimal effect (Figures 2a and S1). Additionally, P3NGln showed a modest affinity toward Zn2+. Notably, P1NGln, P2NGln, and P4NGln showed excellent selectivity toward Cu2+. The His residue in the PNGln contains an imidazole group, which is known for its strong affinity for Cu2+. This imidazole group can readily form coordination bonds with Cu2+, facilitating its selective binding interaction. A possible explanation for the observed selectivity could be attributed to the ability of Cu2+ to deprotonate the amide via coordination at a neutral pH, a property not exhibited by Zn2+. Further, Fe3+ and Al3+, being hard acids, exhibit binding preferences for oxygen over nitrogen.40,41 The amino group of Gly and the amide group along with the imidazole group of His create a coordination environment that favors Cu binding. These findings collectively validate our design strategy, which aimed at achieving selective complexation of Cu2+ by the PNGln in the presence of other competing metal ions.

Figure 2.

Figure 2

Cu chelation by the PNGln. (a) Selectivity of P2NGln among Cu2+, Zn2+, Al3+, and Fe3+ from its fluorescence spectra (λex = 450 nm and λem = 550 nm). (b) Titration curve monitoring fluorescence of P2NGln with increasing concentration of Cu2+. (c) Frequency- and amplitude-normalized EPR spectra of Cu(II)–PNGln. Conditions: [Cu2+] = 450 μM (except for Cu(II)–(P3NGln)2, where [Cu2+] = 225 μM), [peptide] = 500 μM, HEPES 50 mM, pH 7.4, glycerol 10% v/v, and T = 100 K.

The titration of Cu2+ (0–100 μM) into the PNGln solution (50 μM) resulted in a marked decrease in emission intensity, which varied with the concentration of Cu2+ (Figures 2b and S1). This reduction in fluorescence intensity upon Cu2+ addition confirmed the formation of PNGln–Cu2+ complexes. Further, the titration experiment showed that PNGln and Cu2+ formed a complex predominantly in a 1:1 stoichiometric ratio, except for P3NGln, which showed 2:1 stoichiometric complexation, in agreement with the literature and electron paramagnetic resonance (EPR) data (Scheme S3 and Figure 2c). The EPR fingerprint of Cu(II)–P1NGln suggests a square-planar complex with a four-N coordination environment (consisting of NH2, N, N and NIm) (g ≈ 2.190, A ≈ 600 MHz), consistent with the expected Cu2+ binding to the ATCUN motif, which is a small metal-binding site found in the N-termini of many naturally occurring proteins (Scheme S3).38,42,43 The conjugation of the NMI fluorophore to peptides in PNGln did not affect their Cu2+ coordination ability. EPR analysis of Cu(II)–P4NGln complex (g ≈ 2.232, A ≈ 575 MHz) supported a three-N coordination mode of metal binding involving the N-terminal amine, one amidate, and the imidazole moiety of His2 (NH2, N, NIm), along with an additional ligand, typically H2O (Scheme S3).36 In Cu(II)–P2NGln, the major component observed is in line with the expected predominance of the four-N coordination of the ATCUN motif at pH 7.4.44 The fluorescence quenching observed for Cu–(P3NGln)2 complex (g ≈ 2.235, A ≈ 560 MHz) led to EPR spectra acquisition with this stoichiometry, suggesting coordination to three N and one O, similar to the reported crystal structure of Cu(II)–His2, where the equatorial coordination sphere is provided by the two amino groups (NH2), one imidazole (NIm), and one oxygen (Scheme S3).

After confirming the expected coordination sphere of PNGln for Cu2+, we sought to assess their ability to sequester Cu2+ from the Aβ16–Cu2+ complex. The extraction kinetics of Cu2+ from the Aβ16–Cu2+ complex by PNGln was monitored by recording the changes in NMI emission. The NMI fluorescence was found to decrease upon sequestration of Cu2+ from Aβ16 by PNGln, as shown in Figure 3a–d. PNGln showed variable binding kinetics, as evident from the fluorescence quenching within less than 2 min upon addition of Cu2+. This result suggests simultaneous binding of all Cu2+ in the solution to PNGln. P3NGln exhibited a rapid but partial quenching (∼65%) upon sequestration of Cu2+ from Aβ16–Cu2+ complex compared to that in the presence of Cu2+ alone (Figure 3c). P4NGln showed a quenching effect in the presence of Aβ16–Cu2+ complex comparable to that of Cu2+ alone after incubation for ∼15 min, indicating slower kinetics of sequestering Cu2+ from Aβ16–Cu2+ complex (Figure 3d).

Figure 3.

Figure 3

(a–d) Kinetic analysis of Cu2+ coordination by PNGln through monitoring NMI fluorescence: (a) P1NGln; (b) P2NGln; (c) P3NGln; (d) P4NGln. (e, f) MALDI-TOF mass analysis of Aβ16–Cu2+ complex in HEPES (10 mM, pH 7.4) buffer (e) and Aβ16–Cu2+ complex after the addition of P2NGln (f).

P2NGln exhibited quenching similar to that of Cu2+ alone after ∼10 min (Figure 3b). P1NGln displayed significantly slower kinetics and did not show complete quenching even after 20 min (Figure 3a). The overall rate of Cu2+ sequestration from Aβ16–Cu2+ complex by PNGln followed the order P3NGln > P2NGln > P4NGln > P1NGln. Circular dichroism (CD) analysis revealed that Aβ16 exhibits a negative Cotton effect at 200 nm, which suggests the presence of a random coil structure. Upon the addition of Cu2+, the intensity of this band diminished, thereby confirming the Aβ16–Cu2+ complexation. Peptides P1NGln, P2NGln, and P4NGln were able to restore the CD band, with P2NGln demonstrating the highest effectiveness (Figure S2a). On the other hand, P3NGln was less effective. The fluorescent peptides had a minimal impact on the CD signature of Aβ16 (Figure S2b,c). Overall, P2NGln showed the highest efficacy in sequestration of Cu2+ from the Aβ16–Cu2+ complex.

To substantiate the effectiveness of P2NGln in sequestering Cu2+, MALDI-TOF mass spectrometry analysis was performed. The Aβ16 complexation with Cu2+ resulted in the formation of [Aβ16 + Cu] as a major species with corresponding mass m/z = 2016.80. However, bimetallic and trimetallic complexes [Aβ16 + 2Cu] and [Aβ16 + 3Cu] with corresponding masses m/z = 2079.90 and 2141.93, respectively, were also observed, albeit with very low intensity (Figure 3e). Treatment of [Aβ16–Cu] complex with P2NGln revealed distinct peaks at m/z = 933.33 and 997.42, respectively, corresponding to [P2NGln+Cu] and bimetallic copper adduct [P2NGln+2Cu], while a significant decrease in the intensity of [Aβ16 + Cu] peaks was observed (Figure 3f). These findings confirm the potential of P2NGln to selectively sequester Cu2+ from the Aβ16–Cu2+ complex. The complexation of Aβ and Cu2+ has been reported with variable binding affinities of ∼105 to 109 M–1, depending on the methodologies employed and the competing ligands used in the determination.4547 Similarly, a number of short peptides including GHK have been reported to complex Cu with variable binding affinities.22,25,28,34,35,48,49 Small peptides are anticipated to exhibit higher binding affinities for Cu2+ than Aβ, making them potential candidates for sequestration of Cu2+ from the Aβ–Cu2+ complex.

We employed atomic force microscopy (AFM) to comprehensively study the ability of PNGln to sequester Cu2+, thereby inhibiting the Aβ aggregation to form toxic species. Treatment of Aβ42 and Aβ42–Cu2+ samples with P1NGln, P2NGln, or P4NGln resulted in a significant reduction in the formation of Aβ aggregates. P1NGln and P2NGln showed good anti-amyloid activity while P3NGln exhibited moderate activity, as revealed by the presence of fewer fibrils. This underscores the significant role of P1NGln, and P2NGln in inhibiting Aβ aggregation (Figure 4a). Notably, we demonstrated the amyloid aggregation inhibition property of the NMI moiety in our previous study.30 In this study, we treated Aβ42 with NGln to evaluate its anti-amyloidogenic properties. As expected, NGln effectively inhibited Aβ42 aggregation, while relatively lower activity was observed against Aβ42 + Cu2+ aggregation, possibly due to its inability to sequester Cu2+. The control study showed negligible activity of GHK tripeptide in inhibiting Cu-dependent and -independent amyloid aggregation (Figure S3). The AFM analysis confirmed the ability of P2NGln to effectively inhibit Cu2+-induced Aβ fibrillation. We employed a molecular docking study to understand the interaction of PNGln with Aβ42 fibril (PDB ID 2BEG) (Figures 4b and S4–S7). Among all PNGln, P2NGln displayed optimal binding affinity (ΔG = −6.6 kcal/mol) by establishing hydrogen bonds with residues A21, D23, and V36 of the Aβ42 fibril. Furthermore, the NMI moiety participated in a π–π stacking interaction with the F19 residue, while the imidazole of P2NGln formed a T-shaped π–π bond with residue F20 (Figure 4c).

Figure 4.

Figure 4

(a) AFM images of Aβ42 and Aβ42 + Cu2+ fibrils with peptide treatment after incubation for 48 h. (b) Molecular docking of P2NGln with Aβ42 fibril (PDB entry 2BEG). (c) Observed interactions between P2NGln and Aβ42 fibril. (d) Asc consumption (measured at λ = 265 nm) over time. Asc was added first and triggered by adding Cu2+ alone or PNGln–Cu2+ complexes after 10 min. (e) Asc consumption (measured at λ = 265 nm) over time. Asc was added first and triggered by the addition of Cu2+ or Aβ16–Cu2+ after 10 min, followed by the addition of PNGln.

We explored the potential of PNGln in mitigating the Aβ16–Cu2+-induced ROS production. Aβ16–Cu2+ in the presence of ascorbate (Asc), a physiologically relevant reducing agent, undergoes a redox process and is able to catalyze the production of ROS.50 To assess the impact of PNGln on the ROS production induced by Cu2+ and Aβ16–Cu2+, we monitored the consumption of Asc by UV–vis spectroscopy at 265 nm (εAsc = 14,500 M–1 cm–1) (Figures 4d,e and S8a). PNGln was premixed with Cu2+ in HEPES buffer (50 mM, pH 7.4) and added after Asc to initiate the redox reaction. Notably, Cu2+ was found to be highly redox-active in the presence of rapidly reducing conditions of Asc, culminating in complete reduction within 5 min. In contrast, P1NGln–Cu2+ and P2NGln–Cu2+ exhibited negligible oxidation of Asc, aligning with the redox behavior typically associated with the ATCUN–Cu2+ complex.29,38 P4NGln–Cu2+ displayed relatively slow oxidation of Asc, whereas P3NGln–Cu2+ showed faster kinetics, reaching completion after ∼10 min (Figure 4d).

To evaluate the impact of PNGln on the extraction of Cu2+ and inhibition of Aβ16–Cu2+-induced ROS production, PNGln were introduced during Asc consumption, which was triggered by the addition of either Cu2+ or Aβ16–Cu2+ complex (Figure 4e). Next, PNGln were preincubated with Asc solution followed by the addition of Aβ16–Cu2+ complex (Figure S8a). The results show that P3NGln, at two different stoichiometric ratios with Cu2+ (1:1 and 1:2), was unable to attenuate Asc consumption. Given that P3NGln–Cu2+ alone exhibited Asc reduction comparable to that of Aβ16–Cu2+, it was challenging to ascertain the precise site of Cu binding under these conditions (Figure 4e). P4NGln displayed a decrease in Asc consumption compared with P3NGln but did not reach a significant saturation plateau. This finding aligns with the capacity of P4NGln to exclusively bind Cu2+ within a three-N coordination pattern. This interaction gives rise to a marginally redox-active P4NGln–Cu2+ complex, as observed in only Cu2+ (Figure 4e). In contrast, P1NGln successfully prevented Asc reduction when added before or during the Cu–Asc redox cycle (Figures 4e and S8a). This finding aligns with the formation of a redox-inert P1NGln–Cu2+ complex, as illustrated in Figure 3c, where Cu2+ is bound to the four-N motif, effectively resisting reduction by Asc. Nevertheless, the initial change in the slope required to reach the saturation line was not instantaneous, indicating the gradual kinetics involved in the abstraction of Cu2+ from Aβ16–Cu2+ and its subsequent complex formation. This observation correlates with the delayed removal of Cu2+ from Aβ, as observed in the fluorescence measurements conducted earlier (Figure 3a–d). In both experiments (Figures 4e and S8a), P2NGln demonstrated the highest efficacy in limiting Asc consumption. This implies a rapid uptake of Cu2+ from the redox-active Aβ16–Cu2+ by the peptide followed by robust Cu2+ stabilization, making it nonreducible by Asc.

To evaluate the Cu2+ chelation and inhibition of Aβ16–Cu2+-induced ROS production by PNGln, we monitored the fluorescence kinetics of 7-HO-CCA reacting with HO. P3NGln exhibited minimal HO production due to P3NGln–Cu2+ complexation, which shows relatively less redox inhibition. In contrast, P1NGln and P2NGln completely inhibited HO, corroborated by the Asc consumption assay, which is attributed to the formation of redox-dormant PNGln–Cu2+ complexes (Figure S8b).

Aβ aggregates impair neuronal function by increasing the frequency of action potentials and membrane depolarization, resulting in reduced cell viability. This cytotoxic effect is mediated by the ROS production, which in turn is triggered by the interaction of Aβ and Cu2+ within the cells.51 First, the MTT assay was performed by treating the SH-SY5Y cells with PNGln (10–100 μM), and cytotoxicity was assessed after 24 h. The PNGln were nontoxic to cells, with observed viability of over 90% at 100 μM (Figures 5a and S10a). Next, we assessed rescue of cells from toxicity induced by Cu2+ and Asc, which showed the potential of P2NGln at 10 μM to restore over 90% cell viability (Figure 5b). Similarly, P2NGln at the same concentration exhibited remarkable efficacy in scavenging over 90% of ROS produced by Aβ16–Cu2+ and Asc (Figure 5c). 2′,7′-Dichlorofluorescein diacetate (DCFDA) assay was performed to elucidate the mechanism of inhibition of ROS production by PNGln. P2NGln emerged as the most potent inhibitor of ROS production, confirming its superior antioxidant capacity (Figure S10a). In all the assays, P2NGln showed higher efficacy in modulating Aβ16–Cu2+-induced ROS production and amyloid toxicity compared to the other peptide conjugates, including GHK control (P4NGln) (Figure S10b–e). The in vitro ability of PNGln in inhibiting Aβ aggregation led us to perform an in cellulo study to evaluate its effectiveness in mitigating Aβ-induced toxicity under both metal-independent and metal-dependent conditions. SH-SY5Y cells were incubated with Aβ42 fibrils (10 μM) alone and in combination with PNGln (10 μM) in the absence or presence of Cu2+ (10 μM) for 24 h. Cell viability was assessed using the MTT assay. Aβ42 treatment significantly reduced cell viability to ∼55% and 48% in the absence and presence of Cu2+, respectively. Interestingly, PNGln treatment significantly improved cell viability from metal-independent cellular toxicity to ∼80% (Figure 5d). The efficacy of all four PNGln in mitigating Aβ-fibril-induced toxicity underscores the anti-amyloidogenic potential of the NGlu moiety. P1NGln, P2NGln, P3NGln, and P4NGln exhibited rescue of cells from toxicity induced by Aβ42 fibrils in the presence of Cu2+, achieving viabilities of 85%, 92%, 82%, and 86%, respectively. Among them, P2NGln showed the highest efficacy in mitigating metal-dependent Aβ toxicity due to its superior ability to sequester Cu2+ from the Aβ–Cu2+ complex (Figure 5e). Furthermore, P2NGln exhibited a concentration-dependent rescue of cells from toxicity induced by Aβ42 oligomers in both the absence and presence of Cu2+ (Figure S10f,g). These in cellulo studies highlight the potential of P2NGln to effectively protect neuronal cells from both metal-independent and metal-dependent Aβ42 toxicity. A flow cytometry study was carried out to validate the effectiveness of P2NGln in suppressing ROS production in SH-SY5Y cells treated with Aβ16, Cu2+, and Asc using DCFDA dye (Figure 5f–j). The data showed concentration-dependent inhibition of ROS production by P2NGln (Figure 5k,l), which corroborated the results from other experiments. All these results demonstrate the remarkable ability of P2NGln to sequester and keep Cu2+ in a redox-dormant state to suppress ROS production under cellular conditions.

Figure 5.

Figure 5

Cellular studies of PNGln. (a) Cell viability assay of PNGln (100 μM) in SH-SY5Y cells after 24 h of incubation. (b, c) Effect of PNGln on cells treated with (b) Cu2+ and Asc or (c) Aβ16–Cu2+ and Asc. (d) Cellular rescue from Aβ42-induced toxicity by PNGln. (e) Cellular rescue from Aβ42–Cu2+-induced toxicity by PNGln. (f–j) Flow cytometry analysis of SH-SY5Y cells incubated with Aβ16 and Cu2+–Asc followed by treatment with varying concentrations of P2NGln and staining with DCFDA. The percentage written in the P1 quadrant represents the percentage of the population present in that quadrant with higher DCFDA-based fluorescence signal than the whole population. (k) Histogram of SH-SY5Y cells in count versus DCFDA intensity. (l) Quantification of the DCFDA fluorescence intensity after the treatment of cells under Aβ16–Cu2+ and Asc with various concentrations of P2NGln. (The experiment was repeated with n = 3, and statistical analysis was performed using a simple one-way ANOVA; *, p = 0.05).

Anti-inflammatory properties of PNGln were investigated by monitoring ROS levels and nitrite secretion in N9 microglial cells. The cellular uptake of P2NGln was studied by confocal microscopy by monitoring its intrinsic fluorescence property. N9 cells were incubated with P2NGln (5 and 10 μM) in RPMI 1640 medium at 37 °C for 1 h under standard cell culture conditions. The live-cell confocal imaging revealed the localization of P2NGln within the cytoplasm of N9 cells, which indicates its successful uptake across the cell membrane (Figure 6a). N9 cells were stimulated with lipopolysaccharide (LPS), which induced the secretion of proinflammatory cytokines and the activation of NADPH oxidase (NOX), leading to ROS production. The concentration-dependent cytotoxicity assessment revealed that PNGln were nontoxic in N9 microglia cells up to 50 μM (Figure S10b). DCFDA assay was employed to measure the ROS levels in real time after treating the cells with LPS at a concentration of 1 μg/mL for 24 h. PNGln were highly effective in ameliorating the ROS level in the N9 cells (Figure 6b). Treatment with P2NGln significantly inhibited the LPS-induced ROS production in a concentration-dependent manner, as shown in Figure 6c. The anti-inflammatory activity of P2NGln was further confirmed by the Griess assay, which quantifies the nitric oxide (NO) levels as nitrite (NO2). The inflammation in microglial N9 cells was assessed by measuring the nitrite levels in the medium. The nitrite levels in the LPS-treated cells were considered 100%, and in comparison, the untreated cells showed 50%. The LPS-treated cells incubated with PNGln (10 μM) exhibited decrease in nitrite production to 70%, 56%, 85%, and 89% for P1NGln, P2NGln, P3NGln, and PNGln, respectively (Figure 6d). Notably, P2NGln showed the most potent effect, with an IC50 of 6.45 μM, and inhibited nitrite production in a concentration-dependent manner (Figure 6e). Thus, P2NGln effectively modulates inflammatory response by inhibiting LPS-induced microglial activation.

Figure 6.

Figure 6

Effect of P2NGln in modulating inflammation in microglia cells. (a) Cellular uptake of P2NGln (blue: Hoechst dye was used for nuclear staining; green: P2NGln). (b) DCFDA assay in LPS-induced N9 cells treated with PNGlu (10 μM). (c) Concentration-dependent inhibitory effect of P2NGln on ROS production. (d) Effect of PNGln in inhibiting nitrite production. (e) Plot of nitrite secretion versus concentration of P2NGln and IC50 value of P2NGln inhibiting nitrite secretion. (The experiment was repeated with n = 3, and statistical analysis was performed using a simple one-way ANOVA; *, p = 0.05).

Neuroinflammation instigates the production of ROS, which in turn amplifies mitochondrial ROS and initiates mitochondrial fission.52 To assess the effectiveness of P2NGln in mitigating the morphological changes in mitochondria caused by neuroinflammation, the control and LPS-treated cells were stained with MitoTracker Orange and visualized through live-cell fluorescence microscopy. The control cells exhibited elongated and interconnected mitochondrial networks, while exposure to LPS (1 μg/mL) led to mitochondrial fragmentation. Remarkably, treatment with P2NGln (5 and 10 μM) maintained the tubular structures of the mitochondria (Figure 7a). Subsequently, the mitochondrial ROS production was evaluated using MitoSOX Red. The introduction of LPS caused an increase in mitochondrial ROS, which in turn led to an elevation in MitoSOX intensity. Intriguingly, P2NGln was capable of reducing the mitochondrial generation of superoxide anion in a concentration-dependent fashion (Figure 7b–d). These findings underscore the inherent anti-inflammatory attributes of P2NGln within the neurological microenvironment.

Figure 7.

Figure 7

Effect of P2NGln on mitochondrial damage and mitochondrial ROS in N9 cells after inducing neuroinflammation using LPS for 24 h with or without P2NGln. (a) Monitoring mitochondrial fragmentation and its inhibition by P2NGln. Mitochondria are stained with MitoTracker Orange. (b) Monitoring mitochondrial ROS using MitoSOX and its inhibition by P2NGln. (c) Quantification of mitochondrial ROS. (d) Quantification of relative fluorescence intensity of P2NGln showing the amount of P2NGln inside the cells after treatment. (Over 200 cells were quantified; the data between two groups were analyzed by the independent-sample t test, and one-way ANOVA was used to compare the means of three groups; *, p = 0.05).

The multifaceted nature of AD etiology presents a formidable challenge, with the alleviation of oxidative stress, amyloid toxicity, and neuroinflammation forming the cornerstone of therapeutic strategies. Peptides, because of their tunability and biocompatibility, are emerging as potential therapeutic agents for AD. We developed fluorophore-conjugated short peptides through rational design strategies to achieve selective sequestration of Cu2+ and modulation of oxidative stress, Cu-dependent and -independent amyloid toxicity, and neuroinflammation. These fluorescent peptides exhibited superior chelation capabilities for Cu2+ ions and selectively sequestered Cu2+ from the Aβ16–Cu2+ complex, demonstrating a high specificity for Cu2+ over other metal ions. Notably, P2NGln emerged as a potent suppressor of ROS production induced by Aβ16–Cu2+. The cellular studies validated the biocompatibility of these peptides and their efficacy in preserving cell viability. P2NGln was particularly effective in curtailing ROS generation under oxidative stress conditions and mitigating ROS production in microglial cells under inflammatory conditions. This peptide was able to penetrate cells, inhibit nitrite production, rescue mitochondrial fragmentation, and suppress mitochondrial ROS, thereby indicating its anti-inflammatory activity. In conclusion, our findings highlight the therapeutic potential of these fluorescent peptide conjugates in addressing various facets of AD pathology. Further research and optimization could potentially pave the way for the development of innovative therapeutic agents for AD.

Acknowledgments

The authors thank JNCASR, DST (CEFIPRA Grant IFCPAR/CEFIPRA-62T10-3) for the funding. S.M. thanks CSIR for the student fellowship.

Glossary

Abbreviations

AD

Alzheimer’s disease

amyloid β

APP

amyloid precursor protein

NMI

naphthalene monoimide

ATCUN

amino terminal copper(II) and nickel(II) binding

CD

circular dichroism

NMR

nuclear magnetic resonance

LCMS

liquid chromatography–mass spectrometry

HRMS

high-resolution mass spectrometry

MALDI-TOF

matrix-assisted laser desorption ionization time-of-flight mass spectrometry

EPR

electron paramagnetic resonance

AFM

atomic force microscopy

ROS

reactive oxygen species

HEPES

4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid

Asc

sodium ascorbate

LPS

lipopolysaccharide

DCFDA

2′-7′-dichlorodihydrofluorescein diacetate

MTT

3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide

CCA

coumarin carboxylic acid

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmedchemlett.4c00283.

  • Synthetic scheme of fluorescent peptides, characterization data, absorbance and fluorescence spectra, CD spectra, molecular docking figures of the peptides, and AFM images of the Aβ42 fibrils with GHK and fluorescent probe NGln (PDF)

Author Contributions

T.G. conceptualized the project. S.M., Y.V.S., S.S., and T.G. designed the experiments. S.S. and S.M. synthesized the fluorescent peptides. S.M. and Y.V.S. performed in vitro studies. B.V. performed EPR studies. S.M. performed Aβ42 expression and purification, AFM microscopy, in silico studies, and cellular studies. All authors contributed to curating and analyzing the data. S.M., Y.V.S., and T.G. wrote the manuscript, and others gave input.

The authors declare no competing financial interest.

Special Issue

Published as part of ACS Medicinal Chemistry Lettersvirtual special issue “Celebrating the 25th Anniversary of the Chemical Research Society of India”.

Supplementary Material

References

  1. Aguzzi A.; O’Connor T. Protein Aggregation Diseases: Pathogenicity and Therapeutic Perspectives. Nat. Rev. Drug Discov. 2010, 9 (3), 237–248. 10.1038/nrd3050. [DOI] [PubMed] [Google Scholar]
  2. Alzheimer’s Disease: Recent Findings in Pathophysiology, Diagnostic and Therapeutic Modalities; Govindaraju T., Ed.; Royal Society of Chemistry, 2022. [Google Scholar]
  3. Ross C. A.; Poirier M. A. Protein Aggregation and Neurodegenerative Disease. Nat. Med. 2004, 10 (7), S10–S17. 10.1038/nm1066. [DOI] [PubMed] [Google Scholar]
  4. Ramesh M.; Govindaraju T. Multipronged Diagnostic and Therapeutic Strategies for Alzheimer’s Disease. Chem. Sci. 2022, 13 (46), 13657–13689. 10.1039/D2SC03932J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Gaggelli E.; Kozlowski H.; Valensin D.; Valensin G. Copper Homeostasis and Neurodegenerative Disorders (Alzheimer’s, Prion, and Parkinson’s Diseases and Amyotrophic Lateral Sclerosis). Chem. Rev. 2006, 106 (6), 1995–2044. 10.1021/cr040410w. [DOI] [PubMed] [Google Scholar]
  6. O’Brien R. J.; Wong P. C. Amyloid Precursor Protein Processing and Alzheimer’s Disease. Annu. Rev. Neurosci. 2011, 34 (1), 185–204. 10.1146/annurev-neuro-061010-113613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Noda Y.; Asada M.; Kubota M.; Maesako M.; Watanabe K.; Uemura M.; Kihara T.; Shimohama S.; Takahashi R.; Kinoshita A.; Uemura K. Copper Enhances APP Dimerization and Promotes Aβ Production. Neurosci. Lett. 2013, 547, 10–15. 10.1016/j.neulet.2013.04.057. [DOI] [PubMed] [Google Scholar]
  8. Kong G. K. W.; Miles L. A.; Crespi G. A. N.; Morton C. J.; Ng H. L.; Barnham K. J.; McKinstry W. J.; Cappai R.; Parker M. W. Copper Binding to the Alzheimer’s Disease Amyloid Precursor Protein. Eur. Biophys. J. 2008, 37 (3), 269–279. 10.1007/s00249-007-0234-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Metsla K.; Kirss S.; Laks K.; Sildnik G.; Palgi M.; Palumaa T.; Tõugu V.; Palumaa P. α-Lipoic Acid Has the Potential to Normalize Copper Metabolism, Which Is Dysregulated in Alzheimer’s Disease. J. Alzheimer’s Dis. 2022, 85, 715–728. 10.3233/JAD-215026. [DOI] [PubMed] [Google Scholar]
  10. Alies B.; Hureau C.; Faller P. The Role of Metal Ions in Amyloid Formation: General Principles from Model Peptides. Metallomics 2013, 5 (3), 183–192. 10.1039/c3mt20219d. [DOI] [PubMed] [Google Scholar]
  11. Drew S. C.; Barnham K. J. The Heterogeneous Nature of Cu2+ Interactions with Alzheimer’s Amyloid-β Peptide. Acc. Chem. Res. 2011, 44 (11), 1146–1155. 10.1021/ar200014u. [DOI] [PubMed] [Google Scholar]
  12. Syme C. D.; Nadal R. C.; Rigby S. E. J.; Viles J. H. Copper Binding to the Amyloid-β (Aβ) Peptide Associated with Alzheimer’s Disease: Folding, Coordination Geometry, pH Dependence, Stoichiometry, and Affinity of Aβ-(1–28): Insights From a Range of Complementary Spectroscopic Techniques. J. Biol. Chem. 2004, 279 (18), 18169–18177. 10.1074/jbc.M313572200. [DOI] [PubMed] [Google Scholar]
  13. Faller P.; Hureau C.; La Penna G. Metal Ions and Intrinsically Disordered Proteins and Peptides: From Cu/Zn Amyloid-β to General Principles. Acc. Chem. Res. 2014, 47 (8), 2252–2259. 10.1021/ar400293h. [DOI] [PubMed] [Google Scholar]
  14. Hureau C.; Coppel Y.; Dorlet P.; Solari P. L.; Sayen S.; Guillon E.; Sabater L.; Faller P. Deprotonation of the Asp1-Ala2 Peptide Bond Induces Modification of the Dynamic Copper(II) Environment in the Amyloid-β Peptide near Physiological pH. Angew. Chem., Int. Ed. 2009, 48 (50), 9522–9525. 10.1002/anie.200904512. [DOI] [PubMed] [Google Scholar]
  15. Kowalik-Jankowska T.; Ruta-Dolejsz M.; Wiśniewska K.; Łankiewicz L. Cu(II) Interaction with N-Terminal Fragments of Human and Mouse β-Amyloid Peptide. J. Inorg. Biochem. 2001, 86 (2), 535–545. 10.1016/S0162-0134(01)00226-4. [DOI] [PubMed] [Google Scholar]
  16. Rajasekhar K.; Chakrabarti M.; Govindaraju T. Function and Toxicity of Amyloid Beta and Recent Therapeutic Interventions Targeting Amyloid Beta in Alzheimer’s Disease. Chem. Commun. 2015, 51 (70), 13434–13450. 10.1039/C5CC05264E. [DOI] [PubMed] [Google Scholar]
  17. Cheignon C.; Tomas M.; Bonnefont-Rousselot D.; Faller P.; Hureau C.; Collin F. Oxidative Stress and the Amyloid Beta Peptide in Alzheimer’s Disease. Redox Biol. 2018, 14, 450–464. 10.1016/j.redox.2017.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Huang Y.; Mucke L. Alzheimer Mechanisms and Therapeutic Strategies. Cell 2012, 148 (6), 1204–1222. 10.1016/j.cell.2012.02.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Kinney J. W.; Bemiller S. M.; Murtishaw A. S.; Leisgang A. M.; Salazar A. M.; Lamb B. T. Inflammation as a Central Mechanism in Alzheimer’s Disease. Alzheimers Dementia 2018, 4 (1), 575–590. 10.1016/j.trci.2018.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Torres L. L.; Quaglio N. B.; de Souza G. T.; Garcia R. T.; Dati L. M. M.; Moreira W. L.; de Melo Loureiro A. P.; de souza-Talarico J. N.; Smid J.; Porto C. S.; de Campos Bottino C. M.; Nitrini R.; de Moraes Barros S. B.; Camarini R.; Marcourakis T. Peripheral Oxidative Stress Biomarkers in Mild Cognitive Impairment and Alzheimer’s Disease. J. Alzheimer’s Dis. 2011, 26, 59–68. 10.3233/JAD-2011-110284. [DOI] [PubMed] [Google Scholar]
  21. Mendiola A. S.; Ryu J. K.; Bardehle S.; Meyer-Franke A.; Ang K. K.-H.; Wilson C.; Baeten K. M.; Hanspers K.; Merlini M.; Thomas S.; Petersen M. A.; Williams A.; Thomas R.; Rafalski V. A.; Meza-Acevedo R.; Tognatta R.; Yan Z.; Pfaff S. J.; Machado M. R.; Bedard C.; Rios Coronado P. E.; Jiang X.; Wang J.; Pleiss M. A.; Green A. J.; Zamvil S. S.; Pico A. R.; Bruneau B. G.; Arkin M. R.; Akassoglou K. Transcriptional Profiling and Therapeutic Targeting of Oxidative Stress in Neuroinflammation. Nat. Immunol. 2020, 21 (5), 513–524. 10.1038/s41590-020-0654-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Samanta S.; Rajasekhar K.; Babagond V.; Govindaraju T. Small Molecule Inhibits Metal-Dependent and -Independent Multifaceted Toxicity of Alzheimer’s Disease. ACS Chem. Neurosci. 2019, 10 (8), 3611–3621. 10.1021/acschemneuro.9b00216. [DOI] [PubMed] [Google Scholar]
  23. Cheignon C.; Tomas M.; Bonnefont-Rousselot D.; Faller P.; Hureau C.; Collin F. Oxidative Stress and the Amyloid Beta Peptide in Alzheimer’s Disease. Redox Biol. 2018, 14, 450–464. 10.1016/j.redox.2017.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Folk D. S.; Franz K. J. A Prochelator Activated by β-Secretase Inhibits Aβ Aggregation and Suppresses Copper-Induced Reactive Oxygen Species Formation. J. Am. Chem. Soc. 2010, 132 (14), 4994–4995. 10.1021/ja100943r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Hu X.; Zhang Q.; Wang W.; Yuan Z.; Zhu X.; Chen B.; Chen X. Tripeptide GGH as the Inhibitor of Copper-Amyloid-β-Mediated Redox Reaction and Toxicity. ACS Chem. Neurosci. 2016, 7 (9), 1255–1263. 10.1021/acschemneuro.6b00145. [DOI] [PubMed] [Google Scholar]
  26. Okafor M.; Gonzalez P.; Ronot P.; El Masoudi I.; Boos A.; Ory S.; Chasserot-Golaz S.; Gasman S.; Raibaut L.; Hureau C.; Vitale N.; Faller P. Development of Cu(II)-Specific Peptide Shuttles Capable of Preventing Cu-Amyloid Beta Toxicity and Importing Bioavailable Cu into Cells. Chem. Sci. 2022, 13 (40), 11829–11840. 10.1039/D2SC02593K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Rajasekhar K.; Madhu C.; Govindaraju T. Natural Tripeptide-Based Inhibitor of Multifaceted Amyloid β Toxicity. ACS Chem. Neurosci. 2016, 7 (9), 1300–1310. 10.1021/acschemneuro.6b00175. [DOI] [PubMed] [Google Scholar]
  28. Caballero A. B.; Iranzo O.; Hautier A.; Sabaté R.; Gamez P. Peptidic Scaffolds To Reduce the Interaction of Cu(II) Ions with β-Amyloid Protein. Inorg. Chem. 2020, 59 (1), 837–846. 10.1021/acs.inorgchem.9b03099. [DOI] [PubMed] [Google Scholar]
  29. Gonzalez P.; Sabater L.; Mathieu E.; Faller P.; Hureau C. Why the Ala-His-His Peptide Is an Appropriate Scaffold to Remove and Redox Silence Copper Ions from the Alzheimer’s-Related Aβ Peptide. Biomolecules 2022, 12 (10), 1327. 10.3390/biom12101327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Samanta S.; Rajasekhar K.; Ramesh M.; Murugan N. A.; Alam S.; Shah D.; Clement J. P.; Govindaraju T. Naphthalene Monoimide Derivative Ameliorates Amyloid Burden and Cognitive Decline in a Transgenic Mouse Model of Alzheimer’s Disease. Adv. Ther. 2021, 4 (4), 2000225. 10.1002/adtp.202000225. [DOI] [Google Scholar]
  31. Jensen M.; Canning A.; Chiha S.; Bouquerel P.; Pedersen J. T.; Østergaard J.; Cuvillier O.; Sasaki I.; Hureau C.; Faller P. Inhibition of Cu-Amyloid-β by Using Bifunctional Peptides with β-Sheet Breaker and Chelator Moieties. Chem.—Eur. J. 2012, 18 (16), 4836–4839. 10.1002/chem.201103546. [DOI] [PubMed] [Google Scholar]
  32. Sun L.; Cho H.-J.; Sen S.; Arango A. S.; Huynh T. T.; Huang Y.; Bandara N.; Rogers B. E.; Tajkhorshid E.; Mirica L. M. Amphiphilic Distyrylbenzene Derivatives as Potential Therapeutic and Imaging Agents for Soluble and Insoluble Amyloid β Aggregates in Alzheimer’s Disease. J. Am. Chem. Soc. 2021, 143 (27), 10462–10476. 10.1021/jacs.1c05470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Paul A.; Viswanathan G. K.; Mahapatra S.; Balboni G.; Pacifico S.; Gazit E.; Segal D. Antagonistic Activity of Naphthoquinone-Based Hybrids toward Amyloids Associated with Alzheimer’s Disease and Type-2 Diabetes. ACS Chem. Neurosci. 2019, 10 (8), 3510–3520. 10.1021/acschemneuro.9b00123. [DOI] [PubMed] [Google Scholar]
  34. Pickart L.; Freedman J. H.; Loker W. J.; Peisach J.; Perkins C. M.; Stenkamp R. E.; Weinstein B. Growth-Modulating Plasma Tripeptide May Function by Facilitating Copper Uptake into Cells. Nature 1980, 288 (5792), 715–717. 10.1038/288715a0. [DOI] [PubMed] [Google Scholar]
  35. Laussac J. P.; Haran R.; Sarkar B. NMR and EPR Investigation of the Interaction of Copper(II) and Glycyl-l-Histidyl-l-Lysine, a Growth-Modulating Tripeptide from Plasma. Biochem. J. 1983, 209 (2), 533–539. 10.1042/bj2090533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Hureau C.; Eury H.; Guillot R.; Bijani C.; Sayen S.; Solari P.-L.; Guillon E.; Faller P.; Dorlet P. X-Ray and Solution Structures of CuIIGHK and CuIIDAHK Complexes: Influence on Their Redox Properties. Chem.—Eur. J. 2011, 17 (36), 10151–10160. 10.1002/chem.201100751. [DOI] [PubMed] [Google Scholar]
  37. Gonzalez P.; Bossak-Ahmad K.; Vileno B.; Wezynfeld N. E.; El Khoury Y.; Hellwig P.; Hureau C.; Bal W.; Faller P. Triggering Cu-Coordination Change in Cu(II)-Ala-His-His by External Ligands. Chem. Commun. 2019, 55 (56), 8110–8113. 10.1039/C9CC03174J. [DOI] [PubMed] [Google Scholar]
  38. Santoro A.; Walke G.; Vileno B.; Kulkarni P. P.; Raibaut L.; Faller P. Low Catalytic Activity of the Cu(II)-Binding Motif (Xxx-Zzz-His; ATCUN) in Reactive Oxygen Species Production and Inhibition by the Cu(I)-Chelator BCS. Chem. Commun. 2018, 54 (84), 11945–11948. 10.1039/C8CC06040A. [DOI] [PubMed] [Google Scholar]
  39. Liu Y.; Nguyen M.; Robert A.; Meunier B. Metal Ions in Alzheimer’s Disease: A Key Role or Not?. Acc. Chem. Res. 2019, 52 (7), 2026–2035. 10.1021/acs.accounts.9b00248. [DOI] [PubMed] [Google Scholar]
  40. Kozłowski H.; Bal W.; Dyba M.; Kowalik-Jankowska T. Specific Structure-Stability Relations in Metallopeptides. Coord. Chem. Rev. 1999, 184 (1), 319–346. 10.1016/S0010-8545(98)00261-6. [DOI] [Google Scholar]
  41. Sóvágó I.; Kállay C.; Várnagy K. Peptides as Complexing Agents: Factors Influencing the Structure and Thermodynamic Stability of Peptide Complexes. Coord. Chem. Rev. 2012, 256 (19), 2225–2233. 10.1016/j.ccr.2012.02.026. [DOI] [Google Scholar]
  42. 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 (3), 123–130. 10.1021/ar9501535. [DOI] [Google Scholar]
  43. Peisach J.; Blumberg W. E. Structural Implications Derived from the Analysis of Electron Paramagnetic Resonance Spectra of Natural and Artificial Copper Proteins. Arch. Biochem. Biophys. 1974, 165 (2), 691–708. 10.1016/0003-9861(74)90298-7. [DOI] [PubMed] [Google Scholar]
  44. Rakhit G.; Sarkar B. Electron Spin Resonance Study of the Copper(II) Complexes of Human and Dog Serum Albumins and Some Peptide Analogs. J. Inorg. Biochem. 1981, 15 (3), 233–241. 10.1016/S0162-0134(00)80158-0. [DOI] [PubMed] [Google Scholar]
  45. Garzon-Rodriguez W.; Yatsimirsky A. K.; Glabe C. G. Binding of Zn(II), Cu(II), and Fe(II) Ions to Alzheimer’s Aβ Peptide Studied by Fluorescence. Bioorg. Med. Chem. Lett. 1999, 9 (15), 2243–2248. 10.1016/S0960-894X(99)00357-1. [DOI] [PubMed] [Google Scholar]
  46. Atwood C. S.; Scarpa R. C.; Huang X.; Moir R. D.; Jones W. D.; Fairlie D. P.; Tanzi R. E.; Bush A. I. Characterization of Copper Interactions with Alzheimer Amyloid β Peptides: Identification of an Attomolar-Affinity Copper Binding Site on Amyloid β1–42. J. Neurochem. 2000, 75 (3), 1219–1233. 10.1046/j.1471-4159.2000.0751219.x. [DOI] [PubMed] [Google Scholar]
  47. Hong L.; Bush W. D.; Hatcher L. Q.; Simon J. Determining Thermodynamic Parameters from Isothermal Calorimetric Isotherms of the Binding of Macromolecules to Metal Cations Originally Chelated by a Weak Ligand. J. Phys. Chem. B 2008, 112 (2), 604–611. 10.1021/jp075747r. [DOI] [PubMed] [Google Scholar]
  48. Trapaidze A.; Hureau C.; Bal W.; Winterhalter M.; Faller P. Thermodynamic Study of Cu2+ Binding to the DAHK and GHK Peptides by Isothermal Titration Calorimetry (ITC) with the Weaker Competitor Glycine. J. Biol. Inorg. Chem. 2012, 17 (1), 37–47. 10.1007/s00775-011-0824-5. [DOI] [PubMed] [Google Scholar]
  49. 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 (6), 712–718. 10.1039/B212881K. [DOI] [PubMed] [Google Scholar]
  50. Hureau C.; Faller P. Aβ-Mediated ROS Production by Cu Ions: Structural Insights, Mechanisms and Relevance to Alzheimer’s Disease. Biochimie 2009, 91 (10), 1212–1217. 10.1016/j.biochi.2009.03.013. [DOI] [PubMed] [Google Scholar]
  51. Lorenzo A.; Yankner B. A. Amyloid Fibril Toxicity in Alzheimer’s Disease and Diabetes. Ann. N.Y. Acad. Sci. 1996, 777 (1), 89–95. 10.1111/j.1749-6632.1996.tb34406.x. [DOI] [PubMed] [Google Scholar]
  52. Suliman H. B.; Carraway M. S.; Welty-Wolf K. E.; Whorton A. R.; Piantadosi C. A. Lipopolysaccharide Stimulates Mitochondrial Biogenesis via Activation of Nuclear Respiratory Factor-1. J. Biol. Chem. 2003, 278 (42), 41510–41518. 10.1074/jbc.M304719200. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from ACS Medicinal Chemistry Letters are provided here courtesy of American Chemical Society

RESOURCES