pubs.acs.org/JACS Article Unraveling the Stoichiometric Interactions and Synergism between Ligand-Protected Gold Nanoparticles and Proteins Bihan Zhang,◆ María Francisca Matus,◆ Qiaofeng Yao,* Xiaorong Song, Zhennan Wu, Wenping Hu,* Hannu Häkkinen,* and Jianping Xie* Downloaded via UNIV ELECTRONIC SCI & TECHLGY CHINA on February 23, 2025 at 06:41:55 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles. Cite This: https://doi.org/10.1021/jacs.4c09879 ACCESS Metrics & More Read Online Article Recommendations sı Supporting Information * ABSTRACT: Nanomaterials that engage in well-defined and tunable interactions with proteins are pivotal for the development of advanced applications. Achieving a precise molecular-level understanding of nano-bio interactions is essential for establishing these interactions. However, such an understanding remains challenging and elusive. Here, we identified stoichiometric interactions of water-soluble gold nanoparticles (Au NPs) with bovine serum albumin (BSA), unraveling their synergism in manipulating emission of nano-bio conjugates in the second near-infrared (NIR-II) regime. Using Au25(p-MBS)18 (p-MBS = paramercaptobenzenesulfonic acid) as paradigm particles, we achieved precise binding of Au NPs to BSA with definitive molar ratios of 1:1 and 2:1, which is unambiguously evidenced by high-resolution mass spectrometry and transmission electron microscopy. Molecular dynamics simulations identified well-defined binding sites, mediated by electrostatic interactions and hydrogen bonds between the p-MBS moieties on the Au25(p-MBS)18 surface and BSA. Particularly, positively charged residues on BSA were found to be pivotal. By careful control of the molar ratio of Au25(pMBS)18 to BSA, atomically precise [Au25(p-MBS)18]x−BSA conjugates (x = 1 or 2) could be formed. Through a comprehensive spectroscopy study, an electron transfer process and synergistic effect were manifested in the Au25(p-MBS)18−BSA conjugates, leading to drastically enhanced emission in the NIR-II window. This work offers insights into the precise engineering of nanomaterial−protein interactions and opens new avenues for the development of next-generation nano-bio conjugates for nanotheranostics. ■ INTRODUCTION Synthesizing nanomaterials with tunable interactions with biological matter has emerged as one of the exciting frontiers in nanoscience with several potential applications in biosensing,1,2 therapeutics,3−6 diagnostics,7−9 and targeted drug delivery.10−13 On the surface of nanomaterials, adsorption of proteins is a common phenomenon determining the therapeutic effect, targeting effect, biocompatibility and cellular uptake of nanomaterials in vivo.14−18 Documented effort has been made to reveal the interactions between nanomaterials and proteins.19−21 However, the lack of clear molecular-level details of nanomaterial surfaces and the random adsorption of proteins on these surfaces make it challenging to achieve a molecular-level understanding of nanomaterial-protein interactions. This ambiguity hinders the development of precise nano-bio conjugates for advanced applications. Ligand-stabilized, atomically precise gold nanoclusters (Au NCs) are an emerging subclass of metal nanoparticles.22,23 These clusters can, in most cases, be synthesized with a precise chemical composition, e.g., Aum(SR)n where m and n denote the number of Au atoms and thiolate ligands (SR), respectively, and the crystal structure of a large number of these precise compounds is also known. Stabilizing the clusters © XXXX American Chemical Society with hydrophilic ligands creates charging patterns and potential sites for hydrogen bonding on the nanocluster surface, increasing the diversity of favorable noncovalent interactions with proteins. The size of the metal core of Au NCs is generally below 2 nm, creating strong quantum confinement effects and molecular-like properties such as enhanced photoluminescence24,25 due to discrete electronic structures26−28 and good biocompatibilities,29 facilitating a series of successful implementations in bioimaging9,30 and biomedicine9,29−33 It has also been shown that the electronic structure of Au NCs is sensitive to their atomic structure and environment, which can help to unveil the mechanism of the interactions between proteins and Au NCs.34−36 Herein, we employ atomically precise Au25(p-MBS)18 (pMBS = para-mercaptobenzenesulfonic acid) as a model nanoparticle to unravel the stoichiometric interactions between Received: July 20, 2024 Revised: October 9, 2024 Accepted: December 17, 2024 A https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Figure 1. Formation of Au NCs−BSA conjugates. (a) Model structure of Au25(p-MBS)18 (top) and crystal structure of BSA (bottom, PDB ID: 4F5S) with the domains (I, II, and III) and subdomains (A and B) depicted as ribbons in different colors. UV−vis absorption (black line), photoluminescence emission (red line), and excitation (blue dotted line) spectra of (b) Au25(p-MBS)18 (1100 nm emission under 470 nm excitation) and (c) BSA (350 nm emission under 290 nm excitation). (d) Photoluminescence emission spectra of 0.01 mM Au25(p-MBS)18 in the presence of different molar ratios of BSA (from 0 to 10 molar ratios of BSA:Au25(p-MBS)18) with 470 nm excitation. (e) BSA:Au25(p-MBS)18 molar ratio-dependent photoluminescence intensity of Au25(p-MBS)18. (f) DLS graph of Au25(p-MBS)18 at different BSA:Au25(p-MBS)18 molar ratios. Error bars represent the standard deviation of triplicate independent measurements. (g) PAGE result of Au25(p-MBS)18 + BSA at different BSA:Au25(p-MBS)18 molar ratios. Coomassie blue was used to stain the gel for the visualization of BSA at the bottom panel. All of the experiments were conducted in 0.01 M PBS solution at pH 7.4 and room temperature. naturally occurring proteins and water-soluble nanoparticles. The selected Au25(p-MBS)18 NCs exhibit a second nearinfrared window (NIR-II) photoluminescence, high negative surface charge at neutral pH conditions in water, as well as distinct stabilities.37 Bovine serum albumin (BSA)38 is used as the protein model, given its abundance in blood plasma39 (Figure 1A). Because of their high surface charge density and well-organized SR ligands on the surface, Au25(p-MBS)18 NCs can be preferentially bonded to the specific sites of BSA. Comprehensive matrix-assisted laser desorption ionizationtime-of-flight (MALDI-TOF) and transmission electron microscopy (TEM) analyses imply that Au25(p-MBS)18 NCs can react with BSA with a definitive molar ratio of 1:1 or 2:1 to form atomically precise [Au25(p-MBS)18]x−BSA conjugates (x = 1 or 2), suggesting two distinct binding sites in BSA. The nature of such two distinct binding sites is resolved by performing large-scale molecular dynamics (MD) simulations of nanocluster−protein conjugates in water up to 0.2microsecond time scale. The Au NCs bind to the protein via a combination of electrostatic interactions and hydrogen bonds between the p-MBS ligands on the cluster surface and residues from BSA domains I and II. More importantly, detailed steadystate photoluminescence and femtosecond transient absorption (TA) spectroscopy measurements manifest an electron transfer mechanism via such stoichiometric interactions, giving rise to significantly enhanced emission in the NIR-II regime. Besides the electronic and optical property manipulation of Au25(pMBS)18 by BSA, the stability of BSA can be reinforced by the anchoring effects of Au25(p-MBS)18 in the [Au25(p-MBS)18]x− BSA conjugates (x = 1 or 2). Such identified stoichiometric and synergistic Au NCs-proteins interaction paves the way for the creation of advanced nano-bio conjugates with broad applicability in various fields. ■ RESULTS Formation of Au NCs-BSA Conjugates. The discrete electronic structure of Au25(p-MBS)18 enables the NIR-II fluorescence emission40 at 1100 nm under a broad excitation region (300−900 nm) (Figure 1b). We chose 470 nm as the B https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Table 1. Lifetimes of Au25(p-MBS)18 and [Au25(p-MBS)18]1−BSA Conjugates Au25(p-MBS)18 [Au25(p-MBS)18]1−BSA [Au25(p-MBS)18]1−BSA 80 °C [Au25(p-MBS)18]1−BSA 8 M Urea τ1 (ns) res% τ2 (ns) res% τave (ns) 32.6 34.2 48.7 35.3 77.7 13.8 7.6 4.5 78.2 124.1 159.7 125.3 22.3 86.2 92.4 95.5 42.8 111.7 151.3 121.3 excitation wavelength that can excite the strongest fluorescence emission at 1100 nm. According to the relationship between the optical properties and electronic structures of Au25 NCs,26 470 nm excitation includes both the intraband and interband transitions related with both Au atoms in Au13 core and S atoms in ligands, which correlate with the ligand-to-metal charge-transfer (LMCT) process in the NIR-II emission of Au25 NCs.41 Although it was widely accepted that the NIR-II emission of Au25 NCs originates from the relaxation of core excited states,40 two lifetimes, 32.6 and 78.2 ns (Table 1), were simulated from the time-resolved spectroscopy of Au25(pMBS)18 (Figure S1). These two lifetimes may represent the singlet and triplet excited states due to the heavy atoms facilitating the intersystem crossing (ISC) process.42 The relatively shorter lifetime of the triplet excited states here may be due to the heavy NIR-II quenching effect of H2O and O2 in aqueous solution.43 This conclusion is further supported by the observed quenching of Au25(p-MBS)18 emission in the presence of O2 (Figure S2). The pristine fluorescence of BSA at 350 nm shows simpler mechanisms compared to Au25(p-MBS)18 and originates from tryptophan residues (Trp134 and Trp213)44 upon excited at 290 nm (Figure 1c). Upon mixing Au25(p-MBS)18 and BSA in phosphatebuffered saline (PBS) solution at room temperature (see supplemental experimental procedures for details), the visible fluorescence of BSA was quenched (Figure S3). In contrast, the NIR-II fluorescence of Au25(p-MBS)18 was enhanced saliently (up to 4.5 times) and remained nearly unchanged for two hours (Figures 1d and Figure S4), indicating the energy or electron transfer between Au25(p-MBS)18 and BSA. The BSA concentration-dependent emission enhancement of Au25(pMBS)18 (Figure 1e) is reminiscent of the Langmuir adsorption curve. It is worthy pointing out that the emission enhancement saturates after the feeding BSA:Au25(p-MBS)18 molar ratio of ∼2:1 and the curve fits well with the Hill equation for a multibinding model34 (Figure S5), implying the formation of stable and stoichiometric Au NCs−BSA conjugates. Further substantiation of such conjugate formation across varying BSA:Au25(p-MBS)18 ratios is provided by dynamic light scattering (DLS) analysis (Figure 1f). It reveals that the emergence of Au25 NCs−BSA conjugates with a hydrophilic diameter of ∼7 nm occurs at a feeding BSA:Au25 NCs ratio between 1:2 and 2:1. Beyond this ratio, further increasing the BSA dosage would result in excess BSA and lower the mean collective sizes measured by DLS. The formation of Au25 NCs−BSA conjugates at different molar ratios was also examined through polyacrylamide gel electrophoresis (PAGE). As shown in Figure 1g, two distinct bands were observed in the PAGE gels, assignable to free Au25(p-MBS)18 and Au25 NCs−BSA conjugates, respectively. With the increasing molar ratio of BSA:Au25(p-MBS)18, the band corresponding to free Au25(p-MBS)18 fades while the band of Au25 NCs−BSA conjugates intensifies. Notably, the free Au25(p-MBS)18 band dominates in the PAGE gels until a feeding BSA:Au25(p-MBS)18 ratio of 1:1, which disappears entirely at a feeding BSA:Au25(p-MBS)18 ratio of 2:1. The trace amount of free Au25(p-MBS)18 observed at the 1:1 ratio may be attributed to the partial denaturation of BSA in the PAGE separation process. This, in conjunction with the DLS results, strongly supports the conclusion that the enhancement of nanocluster emission in the NIR-II regime stems from the formation of stoichiometric Au NCs−BSA conjugates. BSA Possesses Two Binding Sites for Au NCs. In order to shed fundamental light on the binding sites in the formed Au NCs−BSA conjugates, we performed MALDI-TOF mass spectrometry and negative-stain TEM analysis (Figure 2). Figure 2. BSA possesses two binding sites for Au 25 (pMBS)18nanoclusters. (a) MALDI-TOF mass spectra of Au25(pMBS)18 + BSA at different BSA:Au25(p-MBS)18 molar ratios. (b) Representative negative-stain TEM micrographs of Au NCs−BSA conjugates with 1:1 binding stoichiometry. (c) Representative negative-stain TEM micrograph of Au NCs−BSA conjugates with 2:1 binding stoichiometry. There are only three sets of peaks discernible in the MALDITOF spectra (Figure 2a), where the peaks centered at m/z = 67,000, 75,256, and 83,434 correspond to free BSA, BSA bound with one Au25(p-MBS)18 and BSA bound with two Au25(p-MBS)18 (i.e., [Au25(p-MBS)18]1−BSA, and [Au25(pMBS)18]2−BSA, respectively). Moreover, with the concentration increase of Au25(p-MBS)18, the peaks of [Au25(pMBS)18]x−BSA with higher x value becomes dominant. When the concentration of Au25(p-MBS)18 is less than that of BSA, the spectra displayed peaks only for free BSA and [Au25(pMBS)18]1−BSA. Peaks indicative of [Au25(p-MBS)18]2−BSA emerge only when the concentration of Au25(p-MBS)18 exceeds that of BSA, which aligns with the negative cooperativity of Au25(p-MBS)18−BSA calculated from the Hill equation in the BSA concentration-dependent PL spectra (Figure S5). Notably, no peaks were detected beyond a massto-charge ratio (m/z) of 90,000 (Figure S6), suggesting the absence of additional binding sites on BSA or the formation of the “protein corona”, i.e., a layer of BSA proteins around nanoclusters.15 The above mass spectrometry analysis strongly indicates up to two binding sites existing in BSA for cluster C https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Figure 3. Computational modeling of the specific binding sites for Au25(p-MBS)18on BSA using 1:1 Au NC−BSA conjugates. (a and b) Representative snapshots from the 200 ns MD trajectory of [Au25(p-MBS)18]1−BSA conjugate showing the identified binding sites on BSA in blue (BSite1) and green (BSite2) surfaces. (c and d) Representative zoom-in snapshots from the 200 ns MD trajectory showing the H-bonds (black dashed lines) formed between the ligand layer of Au25(p-MBS)18 and residues from (c) BSite1 and (d) BSite2 on BSA. (e) Number of H-bonds formed between Au25(p-MBS)18 and any region of BSA (cyan line) or specific residues that compose BSite1 (blue line) as a function of the simulation time. (f) Number of H-bonds formed between Au25(p-MBS)18 and any region of BSA (cyan line) or specific residues that compose BSite2 (green line) as a function of the simulation time. Color code for Au25(p-MBS)18: gold atoms are shown in large mustard spheres, sulfur atoms at the metal−ligand interface in small yellow spheres, p-MBS ligands in sticks with carbon atoms in gray, sulfur atoms in yellow, oxygen atoms in red, and hydrogen atoms in white. Color code for BSA: Full structure of BSA is shown in pastel cyan ribbons and binding sites residues in sticks, with carbon atoms in pastel cyan, oxygen atoms in red, nitrogen atoms in blue, and hydrogen atoms in white. anchoring, which is in good agreement with the aforementioned emission enhancement and DLS result. The PAGE results, which clearly show free Au 25 (p-MBS) 18 at a BSA:Au25(p-MBS)18 ratio of 1:2, may be attributed to the biased affinities of these two binding sites (vide infra). The Au25 NCs binding to the relatively weaker site becomes less durable in the PAGE conditions, rendering their desorption from the [Au25(p-MBS)18]2−BSA conjugates. This hypothesis is also corroborated by the detection of [Au25(p-MBS)18]1− BSA in MALDI-TOF spectra at a feeding BSA:Au25(p-MBS)18 ratio of 1:2 (Figure 2a). Negative-stain TEM imaging (using uranyl acetate dihydrate as the contrast agent) was performed to visualize the Au25 NCs within individual protein. In the TEM micrographs (Figures 2b-c), BSA appears bright and white, while Au25(p-MBS)18 and uranyl acetate appear dark due to the heavy metal atoms. A distinct dark spot is evident within the BSA structure at a 1:1 molar ratio of Au25(p-MBS)18 to BSA (Figures 2b). Upon elevating the molar ratio to 2:1, dual dark spots are observed (Figure 2c), suggesting that precise control over the binding sites and stoichiometry in Au NCs−BSA conjugates can be achieved by modulating the relative molar ratios of Au25 NCs and BSA. We extracted images of representative conjugates from a larger TEM image to highlight the binding in Figure 2. More comprehensive views of the Au NCs distribution in BSA can be found in Figures S7 and S8. To gain a better understanding of this nano-bio conjugate formation, we used a computational approach based on MD simulations to D https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Figure 4. Computational modeling of the specific binding sites for Au25(p-MBS)18on BSA using 2:1 Au NCs−BSA conjugates. (a) Snapshots from the 200 ns MD trajectory of [Au25(p-MBS)18]2−BSA showing the proximity of Au25(p-MBS)18 nanoclusters (Au25 NC1 and Au25 NC2) with BSite1 (blue surface) and BSite2 (green surface) at different points of the simulation. (b) Distances between the center of mass (COM) of Au25 NC1 and COM of BSite1 during the simulation time. (c) Distances between COM of Au25 NC2 and COM of BSite2 during the simulation time. (d) Number of H-bonds formed between Au25 NC1 and any region of BSA (cyan line) or specific residues that compose BSite1 (blue line) as a function of simulation time. (e) Number of H-bonds formed between Au25 NC2 and any region of BSA (cyan line) or specific residues that compose BSite2 (green line) as a function of simulation time. (f and g) Zoom-in snapshots at different points of the 200 ns MD trajectory of [Au25(p-MBS)18]2−BSA showing the H-bonds (black dashed lines) formed between the ligand layer of (f) Au25 NC1 and residues from BSite1, and (g) Au25 NC2 and residues from BSite2. Color code for Au25(p-MBS)18: gold atoms are shown in large mustard spheres, sulfur atoms at the metal− ligand interface in small yellow spheres, p-MBS ligands in sticks with carbon atoms in gray, sulfur atoms in yellow, oxygen atoms in red, and hydrogen atoms in white. Color code for BSA: Full structure of BSA is shown in pastel cyan ribbons and binding site residues in sticks, with carbon atoms in pastel cyan, oxygen atoms in red, nitrogen atoms in blue, and hydrogen atoms in white. MBS)18]1−BSA (see Figures S9−10, and supplemental computational procedures for details). By using these initial configurations, two binding sites for Au25(p-MBS)18 (BSite1 and BSite2) were successfully identified (Figures 3a and 3b) and interestingly, featured different affinities (Figures 3c-3f). investigate the specific interactions between Au25(p-MBS)18 and BSA. Computational Modeling of the Specific Binding Sites for Au NCs on BSA. Atomistic MD simulations of 200 ns were conducted for three initial models of [Au25(pE https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Figure 5. Electron transfer from BSA to Au NCs. (a and b) Ultrafast TA spectra of (a) Au25(p-MBS)18 and (b) [Au25(p-MBS)18]1−BSA conjugate. (c and d), Global fitting of the TA maps of (c) Au25(p-MBS)18 and (d) [Au25(p-MBS)18]1−BSA conjugate. (e) XPS spectra of Au25(p-MBS)18 and [Au25(p-MBS)18]1−BSA conjugate. (f) Schematic diagram of the excited state dynamics of [Au25(p-MBS)18]1−BSA conjugate. BSite1 includes residues from BSA domains I and II, including Asp1, Thr2, Lys4, His9, Lys12, and Lys261 and it roughly follows the curvature of the nanocluster’s ligand surface (Figure 3c). BSite2 is located at a small loop in domain I containing the residues Lys116, Lys173, Gly174, and Ala175 (Figure 3d). In both cases, we observed that hydrogen bonds (H-bonds) and electrostatic interactions are the main facilitators of Au25(p-MBS)18−BSA binding. Lysine residues, which are positively charged basic amino acids, greatly benefit the stable interactions with the negatively charged ligand layer (sulfonate groups) of Au25(p-MBS)18. However, the number of H-bonds between Au25(p-MBS)18 and BSA is higher in BSite1 (Figure 3e) than in BSite2 (Figure 3f), and this attraction occurs faster in BSite1 (H-bond formation at 0.4 and 26 ns for BSite1 and BSite2, respectively). In addition, energy decomposition also shows that the contribution of the Coulombic terms (electrostatic interactions) to the total interaction energy is higher in BSite1 (Figures S11−12). We further built a model for [Au25(p-MBS)18]2−BSA based on the identified binding sites BSite1 and BSite2, and analyzed its stability through 200 ns MD simulations (Figure 4). Interestingly, when both binding sites are occupied simultaneously, they exhibit similar affinity for Au25(p-MBS)18, as the initial location remained almost intact during the simulation (Figure 4a). The short distances (<2 nm) between the center of mass (COM) of each binding site and Au25 NCs indicate that both Au25(p-MBS)18 are tightly bound to the protein and stayed inside BSite1 (Figure 4b) and BSite2 (Figure 4c). When F https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS comparing the formation of H-bonds between the Au25 NCs and BSA, a nearly identical number of interactions is observed inside the two binding sites (Figures 4d and 4e). However, no contribution of electrostatic interactions was observed for BSite2 as it was seen for BSite1 (Figure S13), suggesting that BSite1 has a higher affinity for Au25(p-MBS)18. Overall, the same residues previously described were observed to be the main contributors to the Au25(p-MBS)18−BSA interactions (Figures 4f and 4g). Nevertheless, the flexibility of the loops and the subtle movements of Au25 NCs inside the binding sites promote the formation of new interactions with charged residues neighboring the defined binding sites (for example, with Lys239 in BSite1) and, more interestingly, special arrangements of their side chains inside the nanocluster’s ligand layer. For instance, in BSite1, the side chain of Lys4 can be found between three p-MBS groups, very close to the metal−ligand interface (Figure 4f). The mutual attraction between the Lys4 positively charged amino group and the partially negative sulfur atoms at the gold−sulfur interface induces a more open conformation of the p-MBS groups, which eventually could rigidify the metal−ligand interface and thus cause the photoluminescence enhancement effect of Au25(p-MBS)18 when conjugated with BSA, as reported before for similar thiolate-protected Au NCs.45 Thus, despite the high stability of the [Au25(p-MBS)18]2− BSA observed in the MD simulations, the potential different affinity of the binding sites can account for the detection of free Au25(p-MBS)18 in PAGE analysis and [Au25(p-MBS)18]1− BSA presented in MALDI-TOF spectra at a BSA:Au25(pMBS)18 ratio of 1:2 (Figures 1g and 2a). Specifically, BSite2 exhibits a lower affinity, which cannot hold the bound Au25(pMBS)18 throughout the assay conditions used. Electron Transfer from BSA to Au NCs. In our previous discussion, the strong effect of BSA on photoluminescence enhancement of Au25(p-MBS)18 is clear when the conjugate is observed at room temperature (Figure 1d), and without structural disruption of any individual component (Figures S14, S15), which implies the energy or electron transfer between Au25(p-MBS)18 and BSA. Given that the illumination of 470 nm exceeds the excitation wavelength of BSA (Figure 1c), Förster energy transfer between Au25(p-MBS)18 and BSA can be ruled out. As a sufficient electron-rich environment would accelerate the electron transfer process,44,46 we suspect an electron transfer process from BSA to Au25(p-MBS)18. Moreover, BSA contains multiple electron-rich amino acids, and their reduction potential is typically low enough to facilitate the electron transfer process to other materials.47 To shed light on the underlying electron transfer mechanism, a more detailed spectroscopy study was conducted. We chose the [Au25(p-MBS)18]1−BSA conjugates as the model system due to its higher stability and monodispersity, which can provide understanding of this nano-bio interactions at the molecular level. As shown in Figures 5a and 5b, the TA maps of the Au25(p-MBS)18 and [Au25(p-MBS)18]1−BSA conjugates were taken under the excitation of 470 nm, and both showed similar excited-state absorption (ESA) overlapped with ground-state bleaching (GSB). Singular value decomposition (SVD) and global fitting were carried out to extract the time constants of these two TA maps (Figures S16 and S17). Two components, 1 ps and >1 ns, were identified in the Au25(pMBS)18 TA map (Figure 5c). The 1 ps process is attributed to the internal conversion (IC) and ISC as no additional relaxation is observed and the longer lifetime (>1 ns) Article corresponds to the photoluminescence emission decay. Upon the formation of the [Au25(p-MBS)18]1−BSA conjugate, a new electron dynamic process was identified with a decay of 479 fs (Figure 5d). This superfast decay shows enhanced adsorption around 630 and 680 nm. As BSA showed no absorption at 470 nm (Figure 1c), this new component originates from Au25(pMBS)18. Increased absorption implies increased electron density; according to the electronic structure of Au25 NCs, the 680 nm absorption belongs to the Au (sp-sp) transition. Both the slightly increased steady absorptions of [Au25(pMBS)18]1−BSA at around 470 nm in UV−vis spectra (Figure S14) and the decreased content of Au(I) in the shell of Au25(pMBS)18 in X-ray photoelectron spectroscopy (XPS) (Figure 5e) confirmed the electron transfer from BSA to Au25(pMBS)18. Thus, this 479 fs decay was assigned to electron transfer from BSA to Au25(p-MBS)18. Another two components of [Au25(p-MBS)18]1−BSA conjugate, 992 fs and >1 ns, are IC and ISC relaxation and photoluminescence emission decay, respectively. Slightly decreased IC and ISC relaxation of the Au25(p-MBS)18−BSA conjugate compared with that of Au25(p-MBS)18 may result from the rigidified metal−ligand interface after the binding with BSA, which facilitates the ISC process. This also agrees with the enhanced triplet excited state emission after the [Au25(p-MBS)18]1−BSA conjugate formation (Table 1). To figure out how this electron transfer occurs and considering that ligands play an important role in the interactions between Au NCs and BSA, we further constructed three Au25 NCs protected by para-mercaptobenzoic acid (pMBA), 6-mercaptohexanoic acid (MHA), and 3-mercapto-1propanesulfonic acid (MPS). These ligands were chosen to represent different structural motifs, including aromatic rings (p-MBA), alkyl chains of varying lengths (MHA and MPS), and terminal functional groups with different acidities (carboxylic groups and sulfonic groups). We hypothesized that the electronic structures and molecular characteristics of these ligands would influence the electron transfer process between BSA and Au25 NCs. All three Au25 NCs, namely, Au25(p-MBA)18, Au25(MHA)18 and Au25(MPS)18, exhibit NIR photoluminescence under 470 nm excitation (Figure S18). Following incubation with BSA, these Au25 NCs all showed enhanced NIR-II fluorescence. However, only Au25(p-MBA)18 showed comparable BSA photoluminescence enhancement effects and PAGE results (Figure S19) with Au25(p-MBS)18. The weak fluorescence enhancement of alkyl-thiolate-protected Au25 NCs unambiguously indicates that the delocalized electrons on benzene ring is crucial for promoting the electron process between BSA and Au25 NCs. Delocalized orbitals are more likely to be the electron acceptor in the XH-π (X = C or N),48 facilitating the electron transfer from BSA to Au25 NCs. The S XPS spectra of Au25(p-MBS)18 confirm this hypothesis, where increased electron density is observed on S after binding with BSA (Figure S20). Additionally, we observed that Au25(MPS)18, which is protected by a shorter alkyl chain, exhibited a stronger emission enhancement compared to Au25(MHA)18. This can be explained by the shorter alkyl chain facilitating faster electron transfer due to the reduced distance between the interacting surfaces of BSA and Au25 NCs. However, the influence of the terminal functional groups is less significant. Thus, we can conclude that the electron transfer from BSA to Au25 NCs is significantly influenced by the ligand structure. In the case of Au25(p-MBS)18, the electron transfer is from BSA to p-MBS on the surface of Au25(p-MBS)18 and then G https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Article Figure 6. Synergistic effect of the Au NCs−BSA conjugate. (a) Temperature-dependent photoluminescence spectra of the [Au25(p-MBS)18]1−BSA conjugate. (b) Temperature-dependent CD spectra of the [Au25(p-MBS)18]1−BSA conjugate. (c) Urea concentration-dependent photoluminescence spectra of the [Au25(p-MBS)18]1−BSA conjugate. All the experiments were conducted in 0.01 M PBS solution at pH 7.4 and room temperature unless otherwise indicated. affects the photoluminescence of Au25(p-MBS)18 through LMCT. This highlights the importance of ligand selection in tuning the electron transfer and optical properties of Au NCs in protein-conjugated systems. As the electron transfer process highly depends on the proximity between the electron donor and acceptor,49,50 and the LMCT process relies on the ligands and motif conformation, we hypothesized that the electronic structure of Au25(p-MBS)18 can be further tuned by manipulating the structure of BSA. Thus, we subsequently explored the synergistic effect in [Au25(p-MBS)18]1−BSA conjugates. Synergistic Effect of Au NCs−BSA Conjugate. The configuration of BSA is irreversibly/reversibly alterable by diverse stimuli, including temperature and H-bond disruptor. First, we used high temperatures to denature the protein, i.e., loosen the structural configuration of BSA. Incubated under high temperatures (>60 °C) for 30 min, the structure of BSA suffers from a severe distortion (Figure S21), which has been identified mainly as the loosening of α helix and turn structure.51 Just as we suspected, regulating the structure of BSA in this way can further regulate the electronic structure and thus the photoluminescence of the Au25(p-MBS)18. NIR-II photoluminescence of the [Au25(p-MBS)18]1−BSA conjugate intensified with increasing temperature (Figure 6a), which is not observed for free Au25(p-MBS)18 (Figure S22). However, temperature-varied UV−vis absorption and CD spectra showed that the [Au25(p-MBS)18]1−BSA conjugate can also maintain its stability at 80 °C for at least 30 min without compromising the structures of either Au25(p-MBS)18 (Figure S23) or BSA (Figure 6b). Considering the already proved distinct stability of Au25(p-MBS)18 under high temperatures,37 these results suggest that the enhanced nano-bio conjugate stability is greatly attributed to the structural templating effect of Au25(p-MBS)18, which also induces thermal stability on BSA by increasing its denaturation temperature (Figure 6b), consistent with a previously reported effect of sulfonatecontaining ligands on the stability and conformation of BSA.52 Although the structure of Au25(p-MBS)18 and BSA in conjugates remains unchanged under high temperatures (<80 °C), the fact that the protein structure tends to become loose due to the destruction of hydrogen bonds at high temperatures cannot be ignored. Under high temperatures, the loosening tendency of BSA stretches the shell of Au25(p-MBS)18 through its interactions with p-MBS, which can increase the Au(I)− Au(I) distance in the protecting shell of NCs and thus lead to the enhanced emission with blue-shifted emission maximum. It is documented in the literature that increasing Au(I)−Au(I) distance in the frame of aurophilic interactions is capable of H https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS ■ inducing blue-shift of cluster emission, in good agreement with our observation.53,54 This is further supported by the photoluminescence lifetime of [Au25(p-MBS)18]1−BSA conjugate at 80 °C, where the ratio of triplet excited states emission drastically enhanced (Table 1). Under extreme conditions, such as 100 °C for 30 min, the conjugate is disrupted, where Au25(p-MBS)18 decomposition (Figure S23) and BSA denaturation (Figure 6a) are observed. Consequently, the photoluminescence starts to decay (Figure 6b). Thus, the optical properties of Au NCs can reflect the structure of the protein in the formed conjugates. By monitoring the UV−vis spectra of Au25(p-MBS)18 (Figure S24), it is evident that the [Au25(p-MBS)18]1−BSA conjugate shows distinct stability, remaining structurally stable in PBS solution for at least 6 days. In addition to temperature, the electronic structure of Au25(p-MBS)18 can be tuned by adding urea, which can disrupt the H-bonds within BSA and thus alter the structure of BSA.55 By cyclic adding and removing (by dialysis) of aliquots amount of urea into the solution of [Au25(p-MBS)18]1−BSA conjugate, the photoluminescence intensity of Au25(p-MBS)18 exhibits a cyclic ascending and descending pattern (Figure 6c). This reversible photoluminescence enhancement effect corroborates that the stress exerted by BSA on the shell of Au25(p-MBS)18 induces minor and reversible structure distortion. In this vein, the BSA can act as tweezers in the [Au25(p-MBS)18]1−BSA conjugate, finely and controllably tailoring the photoluminescence of Au25 NCs. This observation further demonstrates that the optical properties of Au NCs can reflect the state of the protein in nano-bio conjugates, implying the potential of Au NCs to serve as protein structure probes in bioapplications. These findings suggest a synergistic effect in the nanocluster− protein conjugates, where the nanocluster provides thermal stability to the protein, and simultaneously, the structure of protein manipulate the electronic structure of nanoclusters. Article ASSOCIATED CONTENT sı Supporting Information * The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c09879. Detailed experimental procedures, descriptions of characterization methods, supplemental computational methodologies, time-resolved PL spectroscopy, PL spectra, fitted Hill equation plot, MALDI-TOF mass spectra, negative-stained TEM micrographs, UV−vis spectra, CD spectra, XPS spectra, MD models of BSite1 and BSite2, global fitting of TA spectra (PDF) ■ AUTHOR INFORMATION Corresponding Authors Qiaofeng Yao − Key Laboratory of Organic Integrated Circuits, Ministry of Education & Tianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University, Tianjin 300072, China; Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China; orcid.org/ 0000-0002-5129-9343; Email: qfyao@tju.edu.cn Wenping Hu − Key Laboratory of Organic Integrated Circuits, Ministry of Education & Tianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University, Tianjin 300072, China; Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China; orcid.org/ 0000-0001-5686-2740; Email: huwp@tju.edu.cn Hannu Häkkinen − Departments of Physics and Chemistry, Nanoscience Center, University of Jyväskylä, FI-40014 Jyväskylä, Finland; orcid.org/0000-0002-8558-5436; Email: hannu.j.hakkinen@jyu.fi Jianping Xie − Department of Chemical and Biomolecular Engineering, National University of Singapore, Singapore 117585, Singapore; Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China; orcid.org/0000-0002-3254-5799; Email: chexiej@nus.edu.sg ■ CONCLUSION In summary, our study has revealed the stoichiometry and structural basis for the nanoclusters−protein interactions at the atomic and quantitative levels, presenting the specific and selective noncovalent interactions between Au25(p-MBS)18 and BSA, in which the MD simulations highlighted the crucial role of charged residues within specific domains (such as lysine and histidine) in forming the stable interactions. By finely tuning the molar ratio of Au25(p-MBS)18 to BSA, we achieved molecularly precise [Au25(p-MBS)18]x−BSA conjugates (x = 1 or 2) without structural disruption of any individual component. Spectroscopic analyses further revealed that BSA enhances the NIR-II emission of Au25(p-MBS)18 through electron transfer between BSA residues and ligands on the surface of the Au NCs. The [Au25(p-MBS)18]1−BSA conjugate exhibits a synergistic effect, in which Au25(p-MBS)18 enhances the stability of BSA and the structural dynamics of BSA can be employed to delicately manipulate the electronic and physical structure of Au25(p-MBS)18, while keeping the nanocluster structure largely unaltered. This study not only elucidates the fundamental interactions between NCs and proteins but also illuminates the potential for strategic manipulation of the interactions between nanomaterials and proteins through ligand engineering, which opens promising avenues for the development of advanced nanomedicine. Authors Bihan Zhang − Department of Chemical and Biomolecular Engineering, National University of Singapore, Singapore 117585, Singapore; Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China María Francisca Matus − Departments of Physics and Chemistry, Nanoscience Center, University of Jyväskylä, FI40014 Jyväskylä, Finland; orcid.org/0000-0002-4816531X Xiaorong Song − MOE Key Laboratory for Analytical Science of Food Safety and Biology & State Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou 350116, China; orcid.org/0000-0001-5484-0978 Zhennan Wu − State Key Laboratory of Integrated Optoelectronics, College of Electronic Science and Engineering, Jilin University, Changchun 130012, China; orcid.org/ 0000-0001-5887-7129 Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.4c09879 I https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS Author Contributions (14) Chen, S.; Li, L.; Zhao, C.; Zheng, J. Surface Hydration: Principles and Applications toward Low-Fouling/Nonfouling Biomaterials. Polymer 2010, 51 (23), 5283−5293. (15) Mahmoudi, M.; Bertrand, N.; Zope, H.; Farokhzad, O. C. Emerging Understanding of the Protein Corona at the Nano-Bio Interfaces. Nano Today 2016, 11 (6), 817−832. (16) Wheeler, K. E.; Chetwynd, A. J.; Fahy, K. M.; Hong, B. S.; Tochihuitl, J. A.; Foster, L. A.; Lynch, I. Environmental Dimensions of the Protein Corona. Nat. Nanotechnol. 2021, 16 (6), 617−629. (17) Silva, N. H. C. S.; Vilela, C.; Marrucho, I. M.; Freire, C. S. R.; Pascoal Neto, C.; Silvestre, A. J. D. Protein-Based Materials: From Sources to Innovative Sustainable Materials for Biomedical Applications. J. Mater. Chem. B 2014, 2 (24), 3715−3740. (18) Li, P.; Sun, L.; Xue, S.; Qu, D.; An, L.; Wang, X.; Sun, Z. Recent Advances of Carbon Dots as New Antimicrobial Agents. SmartMat 2022, 3 (2), 226−248. (19) Chen, S.; Zheng, J.; Li, L.; Jiang, S. Strong Resistance of Phosphorylcholine Self-Assembled Monolayers to Protein Adsorption: Insights into Nonfouling Properties of Zwitterionic Materials. J. Am. Chem. Soc. 2005, 127 (41), 14473−14478. (20) Brewer, S. H.; Glomm, W. R.; Johnson, M. C.; Knag, M. K.; Franzen, S. Probing BSA Binding to Citrate-Coated Gold Nanoparticles and Surfaces. Langmuir 2005, 21 (20), 9303−9307. (21) Röcker, C.; Pötzl, M.; Zhang, F.; Parak, W. J.; Nienhaus, G. U. A Quantitative Fluorescence Study of Protein Monolayer Formation on Colloidal Nanoparticles. Nat. Nanotechnol. 2009, 4 (9), 577−580. (22) Jin, R.; Zeng, C.; Zhou, M.; Chen, Y. Atomically Precise Colloidal Metal Nanoclusters and Nanoparticles: Fundamentals and Opportunities. Chem. Rev. 2016, 116 (18), 10346−10413. (23) Chakraborty, I.; Pradeep, T. Atomically Precise Clusters of Noble Metals: Emerging Link between Atoms and Nanoparticles. Chem. Rev. 2017, 117 (12), 8208−8271. (24) Yu, Y.; Luo, Z.; Chevrier, D. M.; Leong, D. T.; Zhang, P.; Jiang, D.-e.; Xie, J. Identification of a Highly Luminescent Au22(SG)18 Nanocluster. J. Am. Chem. Soc. 2014, 136 (4), 1246−1249. (25) Wu, Z.; Yao, Q.; Chai, O. J. H.; Ding, N.; Xu, W.; Zang, S.; Xie, J. Unraveling the Impact of Gold(I)-Thiolate Motifs on the Aggregation-Induced Emission of Gold Nanoclusters. Angew. Chem., Int. Ed. 2020, 59 (25), 9934−9939. (26) Zhu, M.; Aikens, C. M.; Hollander, F. J.; Schatz, G. C.; Jin, R. Correlating the Crystal Structure of a Thiol-Protected Au25 Cluster and Optical Properties. J. Am. Chem. Soc. 2008, 130 (18), 5883−5. (27) Lopez-Acevedo, O.; Kacprzak, K. A.; Akola, J.; Häkkinen, H. Quantum Size Effects in Ambient CO Oxidation Catalysed by LigandProtected Gold Clusters. Nat. Chem. 2010, 2 (4), 329−334. (28) Walter, M.; Akola, J.; Lopez-Acevedo, O.; Jadzinsky, P. D.; Calero, G.; Ackerson, C. J.; Whetten, R. L.; Grönbeck, H.; Häkkinen, H. A Unified View of Ligand-Protected Gold Clusters as Superatom Complexes. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (27), 9157−9162. (29) Zheng, K.; Xie, J. Cluster Materials as Traceable Antibacterial Agents. Acc. Mater. Res. 2021, 2 (11), 1104−1116. (30) Liu, H.; Hong, G.; Luo, Z.; Chen, J.; Chang, J.; Gong, M.; He, H.; Yang, J.; Yuan, X.; Li, L.; Mu, X.; Wang, J.; Mi, W.; Luo, J.; Xie, J.; Zhang, X.-D. Atomic-Precision Gold Clusters for NIR-II Imaging. Adv. Mater. 2019, 31 (46), 1901015. (31) Loynachan, C. N.; Soleimany, A. P.; Dudani, J. S.; Lin, Y.; Najer, A.; Bekdemir, A.; Chen, Q.; Bhatia, S. N.; Stevens, M. M. Renal Clearable Catalytic Gold Nanoclusters for in Vivo Disease Monitoring. Nat. Nanotechnol. 2019, 14 (9), 883−890. (32) Liu, C.-L.; Wu, H.-T.; Hsiao, Y.-H.; Lai, C.-W.; Shih, C.-W.; Peng, Y.-K.; Tang, K.-C.; Chang, H.-W.; Chien, Y.-C.; Hsiao, J.-K.; Cheng, J.-T.; Chou, P.-T. Insulin-Directed Synthesis of Fluorescent Gold Nanoclusters: Preservation of Insulin Bioactivity and Versatility in Cell Imaging. Angew. Chem., Int. Ed. 2011, 50 (31), 7056−7060. (33) Chen, X.; Ren, X.; Gao, X. Peptide or Protein-Protected Metal Nanoclusters for Therapeutic Application. Chin. J. Chem. 2022, 40 (2), 267−274. ◆ B.Z. and M.F.M. contributed equally to this work. Notes The authors declare no competing financial interest. ■ ACKNOWLEDGMENTS The experimental work is financially supported by the Ministry of Education, Singapore (A-8000054-01-00), National Natural Science Foundation of China (22071174, 22371204), and the Fundamental Research Funds for the Central Universities. The computational work at the University of Jyväskylä is supported by the Academy of Finland. The computations were performed in the LUMI supercomputer, owned by the EuroHPC Joint Undertaking and hosted by CSC (Finland), through the Finnish Grand Challenge Project BIOINT. We thank Sami Malola for help in building the initial Au NC models for molecular dynamics simulations. ■ Article REFERENCES (1) Ding, D.; Li, K.; Liu, B.; Tang, B. Z. Bioprobes Based on AIE Fluorogens. Acc. Chem. Res. 2013, 46 (11), 2441−2453. (2) Xavier, P. L.; Chaudhari, K.; Verma, P. K.; Pal, S. K.; Pradeep, T. Luminescent Quantum Clusters of Gold in Transferrin Family Protein, Lactoferrin Exhibiting FRET. Nanoscale 2010, 2 (12), 2769−2776. (3) Zhang, L.; Lin, Y.; Li, S.; Guan, X.; Jiang, X. In Situ Reprogramming of Tumor-Associated Macrophages with Internally and Externally Engineered Exosomes. Angew. Chem., Int. Ed. 2023, 62 (11), e202217089. (4) Song, L.; Li, P.-P.; Yang, W.; Lin, X.-H.; Liang, H.; Chen, X.-F.; Liu, G.; Li, J.; Yang, H.-H. Low-Dose X-Ray Activation of W(VI)Doped Persistent Luminescence Nanoparticles for Deep-Tissue Photodynamic Therapy. Adv. Funct. Mater. 2018, 28 (18), 1707496. (5) Du, B.; Yu, M.; Zheng, J. Transport and Interactions of Nanoparticles in the Kidneys. Nat. Rev. Mater. 2018, 3 (10), 358− 374. (6) Cai, Y.; Ni, D.; Cheng, W.; Ji, C.; Wang, Y.; Müllen, K.; Su, Z.; Liu, Y.; Chen, C.; Yin, M. Enzyme-Triggered Disassembly of Perylene Monoimide-Based Nanoclusters for Activatable and Deep Photodynamic Therapy. Angew. Chem., Int. Ed. 2020, 59 (33), 14014− 14018. (7) Rosi, N. L.; Mirkin, C. A. Nanostructures in Biodiagnostics. Chem. Rev. 2005, 105 (4), 1547−1562. (8) Chen, J.; Hao, L.; Hu, J.; Zhu, K.; Li, Y.; Xiong, S.; Huang, X.; Xiong, Y.; Tang, B. Z. A Universal Boronate-Affinity CrosslinkingAmplified Dynamic Light Scattering Immunoassay for Point-of-Care Glycoprotein Detection. Angew. Chem., Int. Ed. 2022, 61 (7), e202112031. (9) Song, X.; Zhu, W.; Ge, X.; Li, R.; Li, S.; Chen, X.; Song, J.; Xie, J.; Chen, X.; Yang, H. A New Class of NIR-II Gold Nanocluster-Based Protein Biolabels for in Vivo Tumor-Targeted Imaging. Angew. Chem., Int. Ed. 2021, 60 (3), 1306−1312. (10) Zhang, L.; Wang, L.; Xie, Y.; Wang, P.; Deng, S.; Qin, A.; Zhang, J.; Yu, X.; Zheng, W.; Jiang, X. Triple-Targeting Delivery of CRISPR/Cas9 to Reduce the Risk of Cardiovascular Diseases. Angew. Chem., Int. Ed. 2019, 58 (36), 12404−12408. (11) Matus, M. F.; Häkkinen, H. Understanding Ligand-Protected Noble Metal Nanoclusters at Work. Nat. Rev. Mater. 2023, 8 (6), 372−389. (12) Matus, M. F.; Häkkinen, H. Atomically Precise Gold Nanoclusters: Towards an Optimal Biocompatible System from a Theoretical-Experimental Strategy. Small 2021, 17 (27), 2005499. (13) Zheng, Y.; You, S.; Ji, C.; Yin, M.; Yang, W.; Shen, J. Development of an Amino Acid-Functionalized Fluorescent Nanocarrier to Deliver a Toxin to Kill Insect Pests. Adv. Mater. 2016, 28 (7), 1375−1380. J https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX Journal of the American Chemical Society pubs.acs.org/JACS (34) Shang, L.; Brandholt, S.; Stockmar, F.; Trouillet, V.; Bruns, M.; Nienhaus, G. U. Effect of Protein Adsorption on the Fluorescence of Ultrasmall Gold Nanoclusters. Small 2012, 8 (5), 661−665. (35) Bertorelle, F.; Wegner, K. D.; Perić Bakulić, M.; Fakhouri, H.; Comby-Zerbino, C.; Sagar, A.; Bernadó, P.; Resch-Genger, U.; Bonačić-Koutecký, V.; Le Guével, X.; Antoine, R. Tailoring the NIR-II Photoluminescence of Single Thiolated Au25 Nanoclusters by Selective Binding to Proteins**. Chem.�Eur. J. 2022, 28 (39), e202200570. (36) Qu, M.; Xue, F.; Wei, J.-Y.; Qiao, M.-M.; Ren, W.-Q.; Li, S.-L.; Zhang, X.-M. Kernels-Different Aunanoclusters Enhanced Catalytic Performance Via Modification of Ligand and Electronic Effects. Chin. J. Chem. 2022, 40 (21), 2575−2581. (37) Zhang, B.; Wu, Z.; Cao, Y.; Yao, Q.; Xie, J. Ultrastable Hydrophilic Gold Nanoclusters Protected by Sulfonic Thiolate Ligands. J. Phys. Chem. C 2021, 125, 489−497. (38) Majorek, K. A.; Porebski, P. J.; Dayal, A.; Zimmerman, M. D.; Jablonska, K.; Stewart, A. J.; Chruszcz, M.; Minor, W. Structural and Immunologic Characterization of Bovine, Horse, and Rabbit Serum Albumins. Mol. Immunol. 2012, 52 (3), 174−182. (39) Adkins, J. N.; Varnum, S. M.; Auberry, K. J.; Moore, R. J.; Angell, N. H.; Smith, R. D.; Springer, D. L.; Pounds, J. G. Toward a Human Blood Serum Proteome: Analysis by Multidimensional Separation Coupled with Mass Spectrometry *. Mol. Cell. Proteomics 2002, 1 (12), 947−955. (40) Zhou, M.; Jin, R. Optical Properties and Excited-State Dynamics of Atomically Precise Gold Nanoclusters. Annu. Rev. Phys. Chem. 2021, 72 (1), 121−142. (41) Wu, Z.; Jin, R. On the Ligand’s Role in the Fluorescence of Gold Nanoclusters. Nano Lett. 2010, 10 (7), 2568−2573. (42) Turro, N. J.; Ramamurthy, V.; Scaiano, J. C. Principles of Molecular Photochemistry: An Introduction. University science books: 2009. (43) Hirai, H.; Takano, S.; Nakashima, T.; Iwasa, T.; Taketsugu, T.; Tsukuda, T. Doping-Mediated Energy-Level Engineering of M@Au12 Superatoms (M = Pd, Pt, Rh, Ir) for Efficient Photoluminescence and Photocatalysis. Angew. Chem., Int. Ed. 2022, 61 (36), e202207290. (44) Teng, K.-X.; Niu, L.-Y.; Yang, Q.-Z. A Host-Guest Strategy for Converting the Photodynamic Agents from a Singlet Oxygen Generator to a Superoxide Radical Generator. Chem. Sci. 2022, 13 (20), 5951−5956. (45) Pyo, K.; Thanthirige, V. D.; Kwak, K.; Pandurangan, P.; Ramakrishna, G.; Lee, D. Ultrabright Luminescence from Gold Nanoclusters: Rigidifying the Au(I)-Thiolate Shell. J. Am. Chem. Soc. 2015, 137 (25), 8244−8250. (46) Chen, D.; Yu, Q.; Huang, X.; Dai, H.; Luo, T.; Shao, J.; Chen, P.; Chen, J.; Huang, W.; Dong, X. A Highly-Efficient Type I Photosensitizer with Robust Vascular-Disruption Activity for Hypoxic-and-Metastatic Tumor Specific Photodynamic Therapy. Small 2020, 16 (23), 2001059. (47) Reynaud, J. A.; Malfoy, B.; Bere, A. The Electrochemical Oxidation of Three Proteins: Rnaase a, Bovine Serum Albumin and Concanavalin a at Solid Electrodes. J. Electroanal. Chem. Interface Electrochem. 1980, 116, 595−606. (48) Plevin, M. J.; Bryce, D. L.; Boisbouvier, J. Direct Detection of CH/Π Interactions in Proteins. Nat. Chem. 2010, 2 (6), 466−471. (49) Romero, N. A.; Nicewicz, D. A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116 (17), 10075−10166. (50) Marcus, R. A. On the Theory of Oxidation Reduction Reactions Involving Electron Transfer. I. J. Chem. Phys. 1956, 24 (5), 966−978. (51) Murayama, K.; Tomida, M. Heat-Induced Secondary Structure and Conformation Change of Bovine Serum Albumin Investigated by Fourier Transform Infrared Spectroscopy. Biochemistry 2004, 43 (36), 11526−11532. (52) Celej, M. S.; Montich, G. G.; Fidelio, G. D. Protein Stability Induced by Ligand Binding Correlates with Changes in Protein Flexibility. Protein Sci. 2003, 12 (7), 1496−1506. Article (53) Yang, J.-G.; Li, K.; Wang, J.; Sun, S.; Chi, W.; Wang, C.; Chang, X.; Zou, C.; To, W.-P.; Li, M.-D.; Liu, X.; Lu, W.; Zhang, H.-X.; Che, C.-M.; Chen, Y. Controlling Metallophilic Interactions in Chiral Gold(I) Double Salts Towards Excitation Wavelength-Tunable Circularly Polarized Luminescence. Angew. Chem., Int. Ed. 2020, 59 (17), 6915−6922. (54) Huang, Y.; Han, X.; Wang, L.; Pei, R. Assembly-Driven Aggregation-Induced Emission of Gold Nanoclusters with Excitation Wavelength-Dependent Emission and Mechanochromic Property. Adv. Opt. Mater. 2024, 12 (20), 2400078. (55) Liang, M.; Fan, K.; Zhou, M.; Duan, D.; Zheng, J.; Yang, D.; Feng, J.; Yan, X. H-Ferritin-Nanocaged Doxorubicin Nanoparticles Specifically Target and Kill Tumors with a Single-Dose Injection. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (41), 14900−14905. K https://doi.org/10.1021/jacs.4c09879 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX