Tailoring the Structure of 58-Electron Gold Nanoclusters: Au103S2

rules that dictate the structural versus electronic stability and to what extent the .... The central S atom of the dimeric staple is located on t...
0 downloads 0 Views 6MB Size
Article pubs.acs.org/JACS

Tailoring the Structure of 58-Electron Gold Nanoclusters: Au103S2(SNap)41 and Its Implications Tatsuya Higaki,† Chong Liu,‡ Meng Zhou,† Tian-Yi Luo,‡ Nathaniel L. Rosi,‡ and Rongchao Jin*,† †

Department of Chemistry, Carnegie Mellon University, Pittsburgh, Pennsylvania 15213, United States Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States



S Supporting Information *

ABSTRACT: We report the synthesis and crystal structure determination of a gold nanocluster with 103 gold atoms protected by 2 sulfidos and 41 thiolates (i.e., 2-naphthalenethiolates, S-Nap), denoted as Au103S2(S-Nap)41. The crystallographic analysis reveals that the thiolate ligands on the nanocluster form local tetramers by intracluster interactions of C−H···π and π···π stacking. The herringbone pattern formation via intercluster interactions is also observed, which leads to a linearly connected zigzag pattern in the single crystal. The kernel of the nanocluster is a Marks decahedron of Au79, which is the same as the kernel of the previously reported Au102(pMBA)44 (pMBA = −SPh-p-COOH); this is a surprise given the much bulkier naphthalene-based ligand than pMBA, indicating the robustness of the decahedral structure as well as the 58-electron configuration. Despite the same kernel, the surface structure of Au103 is quite different from that of Au102, indicating the major role of ligands in constructing the surface structure. Other implications from Au103 and Au102 include (i) both nanoclusters show similar HOMO−LUMO gap energy (i.e., Eg ≈ 0.45 eV), indicating the kernel is decisive for Eg while the surface is less critical; and (ii) significant differences are observed in the excited-state lifetimes by transient absorption spectroscopy analysis, revealing the kernel-to-surface relaxation pathway of electron dynamics. Overall, this work demonstrates the ligand-effected modification of the gold−thiolate interface independent of the kernel structure, which in turn allows one to map out the respective roles of kernel and surface in determining the electronic and optical properties of the 58e nanoclusters.



subsequent work has identified the trimeric,35,44 tetrameric, and pentameric staples,45−48 and even an octameric ring motif.49 Recent research efforts have focused on the relationship between the kernel and the surface (or interface), as well as the effects on the material properties.50−53 Several reports have clarified the close correlation between the surface structure and properties of nanoclusters such as the catalytic activity,54,55 chirality,56−58 photoluminescence,59−62 and the stability.63,64 The intriguing role of surface ligands in determining the structural patterning of nanoclusters has also been discovered in several cases.38−41,65 For example, the structural (quasi)isomerism of Au28(SR)20 with different −R groups was reported to be induced by the ligands.65 Recent research has also revealed that the kernel structure of the Au30 nanocluster can be controlled to be hcp versus fcc through the choice of thiolate ligands (i.e., 1-adamantanethiol vs tert-butylthiol) and reaction conditions.39 However, the overall relationship between the Aun(SR)m nanocluster structure and the ligand still remains elusive.3

INTRODUCTION

Atomically precise metal nanoclusters have attracted wide research interest in both experiment1−12 and theory.13−20 Some unprecedented structures and novel properties have been discovered, which enable new applications in catalysis, sensing, and optoelectronics.1,21,22 Among the research activities, structural analysis constitutes the very basis for understanding the material properties of nanoclusters.22−29 Since the early success in structure determination of Au 102 (SR) 44 , 30 Au25(SR)18,31,32 and Au38(SR)24,33 significant progress has been achieved in recent years,1 especially the recent success in crystallization of the 2.2 nm Au246(SR)80 nanocluster.34 The earlier work of structural analysis focused on the identification of the types of kernel and surface motifs.1,35 For example, after the discovery of decahedral and icosahedral types, later work has reported the hexagonal close-packed (hcp), body-centered cubic (bcc), and face-centered cubic (fcc) structure types in gold nanoclusters,36−39 whereas bulk gold is of fcc only. Various growth modes of nanoclusters have also been discovered, such as shell-by-shell34,40,41 and layer-bylayer.42,43 In terms of the surface structure, monomeric and dimeric staple motifs were discovered in early work,30−33 and © 2017 American Chemical Society

Received: May 7, 2017 Published: June 29, 2017 9994

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society

Figure 1. Crystallographic structure of the Au103S2(S-Nap)41 nanocluster: (A) total structure of Au103S2(S-Nap)41 (left), kernel structure of Marks decahedral Au79 (middle), and surface staple motifs (right); (B) highlighted view of intracluster ligand−ligand interactions in the surface staple motifs; and (C) anatomy of the tetrameric herringbone structure via C−H···π interactions (top) and the parallel tetramer via staggered π···π interactions (bottom). Color labels: magenta = Au in the kernel, blue = Au in the staple motifs, yellow = S, white = H, and all of the other colors are used for C in different positions.

rules that dictate the structural versus electronic stability and to what extent the ligand plays its role; and (ii) the respective roles of the kernel and surface in governing the properties of nanoclusters. On the other hand, theoretical simulations on large Aun(SR)m nanoclusters (n being a few hundred) are very challenging in achieving high-level calculations. Thus, it is of paramount importance that experimental strategies should be devised to reveal the interplay between the kernel and surface of nanoclusters in deciding the material properties. Herein, we report a new nanocluster formulated as Au103S2(S-Nap)41, which possesses a Au79 kernel being identical to the kernel of the previously reported Au102(pMBA)44.30 Yet these two nanoclusters possess largely different surface structures. This fulfills our goal of tailoring a nanocluster’s surface structure without affecting the kernel, and accordingly gaining insights into the respective roles of the kernel and surface in deciding the electronic and optical properties. Interestingly, the surface ligands on the new Au103S2(SNap)41 cluster show aesthetic patterns via T-shaped or parallel π···π interactions of aromatic naphthalene groups within the nanocluster and between neighboring nanoclusters.

Currently, there have been many successes in structural analysis of relatively small nanoclusters,1,5,22,66 but the structure elucidation of large gold nanoclusters (e.g., more than 100 gold atoms) by X-ray crystallography is still very challenging.34 The pursuit of large nanoclusters is important as such nanoclusters can provide crucial information about the complex, aesthetic surface patterns and the gold−thiolate interfacial structures, as well as their relationship to plasmonic nanoparticles. Therefore, research on the large nanoclusters needs significant efforts to enrich the structure library of nanoclusters for a comprehensive understanding of structural patterning and structure−property correlation.1 Generally, the nanocluster structure can be divided into two parts, that is, the kernel and the surface, which are intimately coupled.1 Therefore, it remains difficult to elucidate how the kernel and surface components respectively determine the material properties and which part plays a dominant role. More critically, there have been no effective strategies yet on how to tailor the kernel or surface structure without affecting the other. Research in tailoring the nanocluster structure is critical in clarifying many major issues, including (i) the fundamental 9995

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society



RESULTS AND DISCUSSION Synthesis and Mass Spectrometry Analysis. The Au103S2(S-Nap)41 nanocluster was synthesized with 2-naphthalenethiol via ligand exchange-induced size/structure transformation (LEIST)3 of Au99(SPh)42. First, Au99(SPh)42 was prepared according to a literature procedure. 67 The Au99(SPh)42 nanocluster was then dissolved in toluene (1 mL), and 4-tert-butyltoluene (1 mL) and 2-naphthalenethiol (240 mg) were added. The reaction was allowed to proceed at 80 °C for 48 h. The product was precipitated with methanol, followed by redissolving and reprecipitation several times to remove excess thiol. Pure Au103S2(S-Nap)41 nanocluster was extracted with dichloromethane. Details of the experiments are described in the Supporting Information. The yield of Au103S2(S-Nap)41 was ∼40% (in gold atom basis). The product was first characterized by electrospray ionization (ESI) mass spectrometry (Figure S1). Cesium acetate was added to form cluster/cesium adducts for ESI analysis. An intense peak was observed at 13573.0, which corresponds to [Au 103S 2(S-Nap)41 + 2Cs]2+ (calculated m/z, 13572.9; deviation, 0.1). Other peaks are also assigned (see notes in Figure S1). The ESI−MS analysis indicates that Au103S2(SNap)41 is charge-neutral, which is also refected in X-ray crystallographic analysis (vide infra). Crystallization and Structure Analysis of the Au103S2(S-Nap)41. Single-crystal growth of the product was performed by vapor diffusion of methanol into a 1,2,4trichlorobenzene solution of the nanoclusters (∼3 mg/mL), followed by X-ray crystallographic analysis. The overall structure of Au103S2(S-Nap)41 (i.e., including the kernel, staples motifs, and carbon tails) is found to be chiral because it conforms to the C2 point group (Figure 1A). On the surface of the nanocluster, the carbon tails of the ligands show several types of aesthetic patterns via C−H···π and π···π interactions (Figure 1B, with details shown in Figure 1C). For example, two naphthalene groups form a dimer (Figure 1C, I and II) via the edge-to-face C−H···π interaction, termed as a T-shaped interaction, where one naphthalene ring perpendicularly stands on the other ring. The two pairs of T-shaped dimers form a cyclic tetramer (Figure 1C, I and II) via lateral C−H···π interaction between the pairs. Regarding the T-shaped pair, the long axes of the naphthalene rings form an angle of 51° for the naphthalene groups to fill the void within the tetramer. The average distance for the C−H···π interaction is 2.76 ± 0.05 Å (Figure 1B, II). This distance is consistent with the reported value for typical C−H···π interaction (2.73 ± 0.13 Å)68 but somewhat shorter than the C−H···π distance in the Au246 nanocluster with −SPh-p-CH3 ligand (2.88 ± 0.42 Å).34 Another tetrameric unit is more deformed because it is connected via C−H···π interaction with the other tetramer (Figure 1B, I and II). This type of arrangement is reminiscent of a “herringbone pattern” and is ubiquitously observed in crystal structures of polycyclic aromatic hydrocarbons69 (including naphthalene) when a T-shaped interaction is favored by the presence of H atoms and rim-C atoms over the stacking interaction. The herringbone-like arrangement was also found in the self-assembled monolayer (SAM) of 2-naphthalenethiol as a stable and dominant pattern.70 The previously reported Au133(TBBT)52 (TBBT = −SPh-p-tBu) also showed “swirl” patterns comprising four ligands on the cluster surface.40 At the bottom of the nanocluster, four naphthalene groups are packed up via π···π stacking (Figure 1B, III and C, III). Each

naphthalene ring is staggered by 97°, and the average spacing is 3.40 ± 0.10 Å (Figure 1C), forming a linear tetramer, which is then sandwiched by the other four ligands from the top and bottom via C−H···π interaction. The Au103S2(S-Nap)41 structure displays complex surface patterns through intracluster ligand interactions. Furthermore, intercluster ligand interactions give rise to another type of tetramer, which is formed by a T-shaped dimer on one cluster (Figure 2A, blue) with another dimer on the neighboring

Figure 2. (A) Intercluster ligand−ligand interactions via C−H···π in the crystal packing structure of the Au103S2(S-Nap)41 nanoclusters. (B) Combination of intercluster and intracluster ligand−ligand interactions to form “herringbone pattern”. Color labels: light blue = Au in the staple motifs, yellow = S, white = H, and all of the other colors are for C in different positions.

cluster (Figure 2A, magenta). These connected dimers indeed comprise part of the intracluster tetramers, respectively (Figure 2B), and they form three sequential cyclic tetramers. The average distance for the C−H···π interaction in the intercluster tetramer is 2.58 ± 0.08 Å, which is shorter than the distances of intracluster tetramers. The intracluster ligand interactions would be preserved in the solution state, while the intercluster ligand interactions would not after the crystal is dissolved. The emergence of herringbone pattern in the Au103S2(SNap)41 nanocluster could provide a critical basis for structural study of SAM with aromatic thiolates because the crystal structure of large nanoclusters is demonstrated to provide insights into the structure enigma of SAM.43 The connectivity via herringbone-like C−H···π interaction continues all of the way along the z-axis of the unit cell, leading to the formation of the needle-like single crystal along the [001] direction. The asformed, linearly wired superstructure of the Au103S2(S-Nap)41 nanoclusters is further laterally connected via intercluster Tshaped C−H···π interactions (Figure S2). As compared to the C5 symmetrical C−H···π interaction in the previously reported Au246(SR)80 nanocluster,34 the Au103S2(S-Nap)41 nanocluster shows a C2 symmetrical intracluster interaction, which is a 9996

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society

possesses μ3 connectivity. The sulfido is completely covered by carbon tails of neighboring ligands, so the absence of carbon tail at this position is dictated by the sterically crowded environment (Figure S4). Because there is no external sulfido source added in our synthesis, the sulfido group must come from the thiol (i.e., via S−C bond breaking). Of note, the S−C bond energy is ∼2.6 eV, and the bond breakage was also previously observed in thermal reactions.37,38 Comparison between Au 1 0 3 S 2 (S-Nap) 4 1 and Au102(pMBA)44. The Au103S2(S-Nap)41 nanocluster possesses nominal 58 free valence electrons of Au 6s1 origin, that is, 103(Au atoms) − 41(ligands) − 2 × 2(sulfidos) = 58e, reminiscent of the isoelectronic Au102(pMBA)44 nanocluster.30 These two nanoclusters share the same Au79 kernel structure, but differ in the gold−thiolate interfacial structures. It is also worth noting that Gao et al. theoretically predicted Au104(SCH3)46 with a Marks decahedral Au79 kernel, which is also a 58-electron nanocluster.71 The stability of 58-electron gold nanoclusters is considered to arise from electronic shell closing (i.e., 1S21P61D102S21F142P61G18) based on theoretical analysis.72 It was also theoretically revealed that the HOMO− LUMO gap of Au102 nanocluster is generated by protecting the charge-neutral Au79 (having no HOMO−LUMO gap) with the outer shell of Au23(pMBA)44; the latter results in depletion of 21-electron from the Au79 kernel, and hence the remaining 58 electrons give rise to closed electron shells and a gap of ∼0.5 eV.71,72 The stability of the nanoclusters was also discussed in terms of the geometrical factor because the Au79 kernel in Au102(pMBA)44 has a Marks decahedron structure with its total surface energy reduced and thus is stabilized.73 The fact that the Au103S2(S-Nap)41 and Au102(pMBA)44 nanoclusters possess the same Au79 kernel (Figure S3, Table S1) despite the significantly different carbon tail structures of ligands (pMBA vs S-Nap) strongly indicates the high geometrical stability of the Au79 kernel. A comparison of the surface structures of Au103 and Au102 is as follows. The top and bottom of both nanoclusters are protected, respectively, by five monomeric staples in C5 symmetry, and the overall structures have a C2 axis in the lateral direction (Figure 4A). However, the protection pattern on the waist of Au102(pMBA)44 is totally different from that of Au103S2(S-Nap)41 (Figure 4B). The Au102(pMBA)44 possesses nine monomeric staples and two dimeric staples on the waist of the Au79 kernel, in contrast with the six monomeric staples, one dimeric staple, and two trimeric staples for the case of Au103S2(S-Nap)41 (Figure 4C,D). In Au102(pMBA)44, the Au atom of one monomeric staple is located on the C2 axis, and the rest of the monomeric staples and two dimeric staples are symmetrically distributed on each side of the axis. Interestingly, the same third Au40 shell is involved to connect the kernel and the staple motifs in both cases. In other words, the number and the position of Au atoms used for staple motif foot-holding are exactly the same in both clusters despite the difference in the number and type of surface motifs in the two nanoclusters. Overall, the structural difference between the two nanoclusters lies in the Au−S interfacial layer. In addition, the surface ligands of Au103S2(S-Nap)41 display aesthetic patterns via T-shaped C−H···π or parallel π···π interactions of aromatic naphthalene groups within the nanocluster as well as between neighboring nanoclusters. In contrast, the structure of the Au102(pMBA)44 nanocluster showed only simple hydrogen bonding and π···π stacking.30

combination of C−H···π and π···π interactions, hence a linearly connected 1D superstructure. The Au103S2(S-Nap)41 nanocluster possesses a Au79 kernel of Marks decahedron. The dissection of the kernel is shown in Figure 3. The first shell of the nanocluster is a Au7 decahedron

Figure 3. Kernel and surface structure of the Au103S2(S-Nap)41 nanocluster: (A) the first shell Au7 (light blue); (B) the second shell Au32 (gray); (C) the third shell Au40 (violet); (D) the fourth shell Au24 (blue) and S43 (yellow); and (E,F) top and side views of surface staple motifs. μ3-Sulfidos are indicated with arrows. Color labels: yellow = S, and all of the other colors are used for Au in different positions.

with D5h symmetry (Figure 3A). The Au7 decahedron is then covered by the second shell of Au32 to form a Au39 Ino decahedron (Figure 3B). The third shell (Au40) covers the top, waist, and bottom of the Au39 decahedron with the C5 symmetry retained (Figure 3C), forming a Au79 Marks decahedron. The quasi-D5h Au79 kernel is further protected by a surface layer consisting of 24 gold atoms and 43 sulfur atoms. The top and the bottom of the kernel are respectively protected by five monomeric staple motifs (i.e., -S-Au-S-) in C5 symmetry (Figure 3D). The waist of the Au79 kernel is protected by one dimeric staple (-S-Au-S-Au-S-), six monomeric staples, and two trimeric-like staples (-S-Au-S*-Au-S-AuS-, note: S* bears no carbon tail) (Figure 3E, indicated by two arrows). The central S atom of the dimeric staple is located on the C2 axis of the cluster. Three out of the six monomeric staples and one out of the two trimeric staples are distributed on one side of the C2 axis, and the rest are on the other side (Figure 3F). The sulfido (S*) in the trimeric-like staple is additionally bonded to a gold atom in the Au40 shell, so it 9997

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society

insight into the fundamental rule for surface protection and the stability of nanoclusters. UV−Vis−NIR Absorption and HOMO−LUMO Gap. The steady-state UV−vis−NIR spectrum of the Au103S2(S-Nap)41 nanocluster shows peaks at 434, 542, 634, and 744 nm (Figure 5A), with continuous absorption up to 3000 nm (Figure 5B);

Figure 5. UV−vis (A) and NIR (B) absorption spectra of the Au103S2(S-Nap)41 nanocluster in CCl4. The asterisk indicates the solvent absorption peak (i.e., overtone).

note, for the convenience of seeing fine spectral features, the spectrum is shown in two ranges (300−1000 and 1000−3000 nm). The spectrum of Au 103 S 2 (S-Nap) 41 shows more pronounced peaks than Au102(pMBA)44, albeit the peak positions are similar (Figure S5). The similarity in the absorption spectra indicates that the kernel structure plays a more decisive role. The optical HOMO−LUMO gap of Au103S2(S-Nap)41 is determined to be 0.42 eV by extrapolation of absorbance to the baseline (Figure 5B, inset). This optical gap is consistent with the electrochemically determined HOMO−LUMO gap (0.38 eV, after subtraction of 0.22 eV charging energy; of note, the charging energy is present regardless of neutral or charged nanoclusters10,11a,26b) by differential pulse voltammetry (DPV, see Figure 6). The Au102(pMBA)44 was reported to have a

Figure 4. Comparison of surface structures of the Au103S2(S-Nap)41 (A,C) and the Au102(pMBA)44 (B,D) nanoclusters: (A,B) side views of Au−S staple motifs of the two nanoclusters; (C,D) top views of staple motifs at waist positions. Color labels: yellow = S, and all of the other colors are used for Au in staple motifs at different positions.

On the Ligand’s Role in Dictating the Structure. Previous reports revealed the intriguing role of ligand in determining the structures of nanoclusters.38−41,65 For example, Au133(TBBT)52 possesses an isotropic shape with an icosahedron-based kernel protected by helical patterns of staple motifs,40 whereas Au130(pMBT)50 (pMBT = −SPh-p-CH3) shows a barrel-shaped Ino decahedron-based kernel protected by ripple-like patterns of staple motifs.41 The differences in the kernel and protecting patterns were ascribed to the effect of the substituents at the para-position in the ligand (i.e., −C(CH3)3 vs −CH3).40,41 The repulsion between the ligands becomes more prominent in larger-sized nanoclusters because an increase in cluster size leads to a decrease in surface curvature as well as the spacing between ligands. Also, the position of substituent on the benzene ring was reported to be a critical factor for controlling the size and structure of nanoclusters.74 In our present work, however, the 2-naphthalenethiol ligand yields the Au103S2(SNap)41, which is close in size to the Au102(pMBA)44 protected by 4-mercaptobenzoic acid, despite the bulkier S-Nap ligand than pMBA. This case can be distinguished partly because of the contraction of surface ligands via attractive aromatic interaction, instead of simple repulsion between bulky carbon tails. The tertiary butyl group of TBBT contains all aliphatic carbons with steric effect toward all directions regardless of the bond rotation. However, the bulkiness of the 2-naphthalenethiol comes from the extension of aromatic conjugation, which can provide stronger, attractive interactions via quadrupoles on aromatic rings. The contracted surface structure shows up as aesthetic patterns, such as the herringbone pattern, and it allows the Au103S2(S-Nap)41 nanocluster to possess the same kernel structure as Au102(pMBA)44 without being hindered by steric effect. The introduction of attractive interaction among ligands in this work is thus important and can also give us deep

Figure 6. DPV of the Au103S2(S-Nap)41 nanocluster.

similar value of experimental HOMO−LUMO gap, 0.45 eV.75 The Au104 nanocluster was reported to have a theoretical HOMO−LUMO gap similar to that of Au102(SCH3)44.71 Thus, the HOMO−LUMO gap energy is also determined by the kernel of the nanocluster. Ultrafast Electron Dynamics. Femtosecond transient absorption (fs-TA) spectroscopy was performed to probe the electron dynamics of the Au103S2(S-Nap)41 nanocluster, in 9998

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society

Figure 7. (A) Transient absorption data map of Au103S2(S-Nap)41 pumped at 480 nm. (B) Evolution associated spectra (EAS) obtained from the global fitting. (C) Kinetic traces at selected wavelengths and the corresponding fits. (D) Normalized decay kinetics around 540 nm as a function of laser fluence.

energy in the core then relaxes via core phonons or interactions with surface structure.76,80 The comparison between Au103S2(SNap)41 and Au102(pMBA)44 suggests that the surface structure modifies the core relaxation dynamics so strongly that two Marks decahedral nanoclusters show different excited-state lifetimes by a factor of 10. The observation of surface interaction in large nanoclusters is very interesting because of the smaller population of the surface atoms over the kernel atoms as compared to the smaller-sized nanoclusters. The excited-state structural relaxation in small nanoclusters such as Au25(SR)1878 would be less likely in large nanoclusters such as Au103 and Au102; thus we believe it is more reasonable to have a kernel-to-surface relaxation picture of electrons. Overall, the Au103S2(S-Nap)41 nanocluster shows significantly different electron dynamics as compared to that of Au102(pMBA)44 nanocluster. The results imply the important role of the gold−thiolate interfacial structure in affecting electron dynamics of the nanoclusters.

particular, the effect of surface structure. The TA spectrum of the nanocluster as a function of time delay after 480 nm pump pulse is shown in Figure 7A. The TA signal consists of a strong, net ground-state bleaching (GSB) signal around 550 nm as well as two excited-state absorption (ESA) peaks around 700 and 800 nm (Figure 7A), and the majority of TA signal decays to zero within 2 ns. Global fitting was applied to extract the excited-state species of Au103S2(S-Nap)41 (Figure 7B). Fitting the population dynamics of the TA signal required three decaying components (2.0, 16.6, and 420 ps) with good fitting quality (e.g., see the 540 and 670 nm kinetic curves shown in Figure 7C). The 2.0 ps component should be assigned to the internal conversion from the Sn to S1 state. The 16.6 ps component can be ascribed to the vibrational cooling from hot S1 to S1, while the 420 ps component represents the relaxation from S1 to the ground state. According to the evolution associated spectra (EAS), the second and third components (Figure 7B, green and red) show almost the same spectral features at all wavelengths. The electron dynamics of Au103S2(S-Nap)41 is independent of the pump fluence (Figure 7D), demonstrating that the Au103S2(S-Nap)41 nanocluster has a nonmetallic electronic structure. The excitonic behavior of this nanocluster is also consistent with the previous observation of molecular-like dynamics in the Au102(pMBA)44 nanocluster.76,77 However, regarding the relaxation from S1 to the ground state, the excited-state lifetime of the Au103S2(S-Nap)41 (420 ps) is almost 10 times shorter than the previously reported lifetime for the Au102(pMBA)44 (>3 ns).76,77 Upon photoexcitation, transitions from ground state to excited state occur through core-based orbitals predominantly,78,79 and the excited



CONCLUSION In summary, a new nanocluster formulated as Au103S2(S-Nap)41 is obtained, which provides important implications for the structural and electronic stabilities, as well as the structural insights into the electron dynamics and optical absorption properties. Interestingly, the naphthalene ligands on the nanocluster form tetrameric patterns by local intracluster C− H···π and π···π interactions on the surface. The ligands also interact between adjacent nanoclusters to form a 3D macroscopic crystal lattice. The Au79 kernel of the nanocluster 9999

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society

(9) Goswami, N.; Yao, Q.; Chen, T.; Xie, J. Coord. Chem. Rev. 2016, 329, 1. (10) Kwak, K.; Tang, Q.; Kim, M.; Jiang, D.-e.; Lee, D. J. Am. Chem. Soc. 2015, 137, 10833. (11) (a) Rambukwella, M.; Sementa, L.; Barcaro, G.; Fortunelli, A.; Dass, A. J. Phys. Chem. C 2015, 119, 25077. (b) Nimmala, P. R.; Theivendran, S.; Barcaro, G.; Sementa, L.; Kumara, C.; Jupally, V. R.; Apra, E.; Stener, M.; Fortunelli, A.; Dass, A. J. Phys. Chem. Lett. 2015, 6, 2134. (c) Antonello, S.; Dainese, T.; Pan, F.; Rissanen, K.; Maran, F. J. Am. Chem. Soc. 2017, 139, 4168. (12) Stoll, T.; Sgrò, E.; Jarrett, J. W.; Réhault, J.; Oriana, A.; Sala, L.; Branchi, F.; Cerullo, G.; Knappenberger, K. L., Jr. J. Am. Chem. Soc. 2016, 138, 1788. (13) Akola, J.; Walter, M.; Whetten, R. L.; Häkkinen, H.; Grönbeck, H. J. Am. Chem. Soc. 2008, 130, 3756. (14) (a) Pei, Y.; Gao, Y.; Zeng, X. C. J. Am. Chem. Soc. 2008, 130, 7830. (b) Lopez-Acevedo, O.; Tsunoyama, H.; Tsukuda, T.; Häkkinen, H.; Aikens, C. M. J. Am. Chem. Soc. 2010, 132, 8210. (15) (a) Aikens, C. M. J. Phys. Chem. Lett. 2011, 2, 99. (b) Tian, Z.; Liu, W.; Cheng, L. Prog. Chem. 2015, 27, 1743. (16) Pelayo, J. J.; Whetten, R. L.; Garzón, I. L. J. Phys. Chem. C 2015, 119, 28666. (17) Tang, Q.; Ouyang, R.; Tian, Z.; Jiang, D.-e. Nanoscale 2015, 7, 2225. (18) (a) Xu, W. W.; Li, Y.; Gao, Y.; Zeng, X. C. Nanoscale 2016, 8, 7396. (b) Xu, W. W.; Li, Y.; Gao, Y.; Zeng, X. C. Nanoscale 2016, 8, 1299. (c) Liu, C.; Pei, Y.; Sun, H.; Ma, J. J. Am. Chem. Soc. 2015, 137, 15809. (d) Xiong, L.; Peng, B.; Ma, Z.; Wang, P.; Pei, Y. Nanoscale 2017, 9, 2895. (19) (a) Molina, B.; Tlahuice-Flores, A. Phys. Chem. Chem. Phys. 2016, 18, 1397. (b) Varnholt, B.; Guberman-Pfeffer, M. J.; Oulevey, P.; Antonello, S.; Dainese, T.; Gascón, J. A.; Bürgi, T.; Maran, F. J. Phys. Chem. C 2016, 120, 25378. (20) (a) Baseggio, O.; De Vetta, M.; Fronzoni, G.; Stener, M.; Sementa, L.; Fortunelli, A.; Calzolari, A. J. Phys. Chem. C 2016, 120, 12773. (b) Iida, K.; Noda, M.; Nobusada, K. J. Phys. Chem. C 2016, 120, 2753. (c) Takagi, N.; Ishimura, K.; Matsui, M.; Fukuda, R.; Matsui, T.; Nakajima, T.; Ehara, M.; Sakaki, S. J. Am. Chem. Soc. 2015, 137, 8593. (21) (a) Li, G.; Jin, R. Acc. Chem. Res. 2013, 46, 1749. (b) Yamazoe, S.; Yoskamtorn, T.; Takano, S.; Yadnum, S.; Limtrakul, J.; Tsukuda, T. Chem. Rec. 2016, 16, 2338. (c) Zhang, L.; Wang, E. Nano Today 2014, 9, 132. (d) Yuan, X.; Luo, Z.; Yu, Y.; Yao, Q.; Xie, J. Chem. - Asian J. 2013, 8, 858. (22) Xu, W. W.; Zhu, B.; Zeng, X. C.; Gao, Y. Nat. Commun. 2016, 7, 13574. (23) (a) Chen, S.; Xiong, L.; Wang, S.; Ma, Z.; Jin, S.; Sheng, H.; Pei, Y.; Zhu, M. J. Am. Chem. Soc. 2016, 138, 10754. (b) Wang, S.; Jin, S.; Yang, S.; Chen, S.; Song, Y.; Zhang, J.; Zhu, M. Sci. Adv. 2015, 1, e1500441. (24) Higaki, T.; Liu, C.; Chen, Y.; Zhao, S.; Zeng, C.; Jin, R.; Wang, S.; Rosi, N. L.; Jin, R. J. Phys. Chem. Lett. 2017, 8, 866. (25) (a) Conn, B. E.; Atnagulov, A.; Yoon, B.; Barnett, R. N.; Landman, U.; Bigioni, T. P. Sci. Adv. 2016, 2, e1601609. (b) He, X.; Wang, Y.; Jiang, H.; Zhao, L. J. Am. Chem. Soc. 2016, 138, 5634. (26) (a) Liao, L.; Zhuang, S.; Yao, C.; Yan, N.; Chen, J.; Wang, C.; Xia, N.; Liu, X.; Li, M.-B.; Li, L.; Bao, X.; Wu, Z. J. Am. Chem. Soc. 2016, 138, 10425. (b) Dong, H.; Liao, L.; Zhuang, S.; Yao, C.; Chen, J.; Tian, S.; Zhu, M.; Liu, X.; Li, L.; Wu, Z. Nanoscale 2017, 9, 3742. (27) (a) Alhilaly, M. J.; Bootharaju, M. S.; Joshi, C. P.; Besong, T. M.; Emwas, A.-H.; Juarez-Mosqueda, R.; Kaappa, S.; Malola, S.; Adil, K.; Shkurenko, A.; Häkkinen, H.; Eddaoudi, M.; Bakr, O. M. J. Am. Chem. Soc. 2016, 138, 14727. (b) Bootharaju, M. S.; Dey, R.; Gevers, L. E.; Hedhili, M. N.; Basset, J.-M.; Bakr, O. M. J. Am. Chem. Soc. 2016, 138, 13770. (28) (a) Yang, H.; Yan, J.; Wang, Y.; Su, H.; Gell, L.; Zhao, X.; Xu, C.; Teo, B. K.; Häkkinen, H.; Zheng, N. J. Am. Chem. Soc. 2017, 139, 31. (b) Yan, J.; Su, H.; Yang, H.; Hu, C.; Malola, S.; Lin, S.; Teo, B. K.; Häkkinen, H.; Zheng, N. J. Am. Chem. Soc. 2016, 138, 12751.

possesses quasi-D 5h symmetry with Marks decahedron structure, which is the same as in the previously reported Au102(pMBA)44 nanocluster. The two nanoclusters show similar UV−vis−NIR spectral profiles and HOMO−LUMO gap energy, indicating the dominant role of the kernel. However, the ultrafast electron dynamics are quite different in that the long-lived excited-state lifetime of Au103S2(S-Nap)41 (420 ps) is significantly shorter than that of the Au102(pMBA)44 (>3 ns), that is, by a factor of 10. This work demonstrates the tailoring of surface structure by ligand−ligand interactions, as well as the interplay between the kernel and the surface. The observations offer further insights into the universal principles of the surface structure prediction of Aun(SR)m nanoclusters, which are expected to be realized in future work via a combination of theory and experiment.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b04678. Details of the synthesis, crystallization, and X-ray analysis, and supporting Figures S1−S5 and Tables S1 and S2 (PDF) X-ray crystallographic data for Au103S2(S-Nap)41 (CIF)



AUTHOR INFORMATION

Corresponding Author

*[email protected] ORCID

Tatsuya Higaki: 0000-0002-6759-9009 Tian-Yi Luo: 0000-0002-9973-9328 Nathaniel L. Rosi: 0000-0001-8025-8906 Rongchao Jin: 0000-0002-2525-8345 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS R.J. acknowledges the financial support by the Air Force Office of Scientific Research under AFOSR award no. FA9550-15-19999 (FA9550-15-1-0154) and 2016 Defense University Research Instrumentation Program (DURIP). We thank Dr. Zhongrui Zhou for assistance in ESI−MS analysis.



REFERENCES

(1) Jin, R.; Zeng, C.; Zhou, M.; Chen, Y. Chem. Rev. 2016, 116, 10346. (2) Maity, P.; Xie, S.; Yamauchi, M.; Tsukuda, T. Nanoscale 2012, 4, 4027. (3) Zeng, C.; Chen, Y.; Das, A.; Jin, R. J. Phys. Chem. Lett. 2015, 6, 2976. (4) Kurashige, W.; Niihori, Y.; Sharma, S.; Negishi, Y. Coord. Chem. Rev. 2016, 320−321, 238. (5) Fernando, A.; Weerawardene, K. L. D. M.; Karimova, N. V.; Aikens, C. M. Chem. Rev. 2015, 115, 6112. (6) (a) Wang, Q.-M.; Lin, Y.-M.; Liu, K.-G. Acc. Chem. Res. 2015, 48, 1570. (b) Zeng, J.-L.; Guan, Z.-J.; Du, Y.; Nan, Z.-A.; Lin, Y.-M.; Wang, Q.-M. J. Am. Chem. Soc. 2016, 138, 7848. (7) Udayabhaskararao, T.; Pradeep, T. J. Phys. Chem. Lett. 2013, 4, 1553. (8) (a) Yau, S. H.; Varnavski, O.; Goodson, T., III. Acc. Chem. Res. 2013, 46, 1506. (b) Abeyasinghe, N.; Kumar, S.; Sun, K.; Mansfield, J. F.; Jin, R.; Goodson, T., III. J. Am. Chem. Soc. 2016, 138, 16299. 10000

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001

Article

Journal of the American Chemical Society (29) (a) Erickson, J. D.; Mednikov, E. G.; Ivanov, S. A.; Dahl, L. F. J. Am. Chem. Soc. 2016, 138, 1502. (b) Kenzler, S.; Schrenk, C.; Schnepf, A. Angew. Chem., Int. Ed. 2017, 56, 393. (30) Jadzinsky, P. D.; Calero, G.; Ackerson, C. J.; Bushnell, D. A.; Kornberg, R. D. Science 2007, 318, 430. (31) Zhu, M.; Aikens, C. M.; Hollander, F. J.; Schatz, G. C.; Jin, R. J. Am. Chem. Soc. 2008, 130, 5883. (32) Heaven, M. W.; Dass, A.; White, P. S.; Holt, K. M.; Murray, R. W. J. Am. Chem. Soc. 2008, 130, 3754. (33) Qian, H.; Eckenhoff, W. T.; Zhu, Y.; Pintauer, T.; Jin, R. J. Am. Chem. Soc. 2010, 132, 8280. (34) Zeng, C.; Chen, Y.; Kirschbaum, K.; Lambright, K. J.; Jin, R. Science 2016, 354, 1580. (35) (a) Jiang, D.-e.; Chen, W.; Whetten, R. L.; Chen, Z. J. Phys. Chem. C 2009, 113, 16983. (b) Pei, Y.; Gao, Y.; Shao, N.; Zeng, X. C. J. Am. Chem. Soc. 2009, 131, 13619. (36) Zeng, C.; Qian, H.; Li, T.; Li, G.; Rosi, N. L.; Yoon, B.; Barnett, R. N.; Whetten, R. L.; Landman, U.; Jin, R. Angew. Chem., Int. Ed. 2012, 51, 13114. (37) Liu, C.; Li, T.; Li, G.; Nobusada, K.; Zeng, C.; Pang, G.; Rosi, N. L.; Jin, R. Angew. Chem., Int. Ed. 2015, 54, 9826. (38) Higaki, T.; Liu, C.; Zeng, C.; Jin, R.; Chen, Y.; Rosi, N. L.; Jin, R. Angew. Chem., Int. Ed. 2016, 55, 6694. (39) Higaki, T.; Zeng, C.; Chen, Y.; Hussain, E.; Jin, R. CrystEngComm 2016, 18, 6979. (40) Zeng, C.; Chen, Y.; Kirschbaum, K.; Appavoo, K.; Sfeir, M. Y.; Jin, R. Sci. Adv. 2015, 1, e1500045. (41) Chen, Y.; Zeng, C.; Liu, C.; Kirschbaum, K.; Gayathri, C.; Gil, R. R.; Rosi, N. L.; Jin, R. J. Am. Chem. Soc. 2015, 137, 10076. (42) Zeng, C.; Chen, Y.; Iida, K.; Nobusada, K.; Kirschbaum, K.; Lambright, K. J.; Jin, R. J. Am. Chem. Soc. 2016, 138, 3950. (43) Zeng, C.; Liu, C.; Chen, Y.; Rosi, N. L.; Jin, R. J. Am. Chem. Soc. 2016, 138, 8710. (44) Das, A.; Li, T.; Nobusada, K.; Zeng, C.; Rosi, N. L.; Jin, R. J. Am. Chem. Soc. 2013, 135, 18264. (45) Das, A.; Li, T.; Li, G.; Nobusada, K.; Zeng, C.; Rosi, N. L.; Jin, R. Nanoscale 2014, 6, 6458. (46) (a) Pei, Y.; Pal, R.; Liu, C.; Gao, Y.; Zhang, Z.; Zeng, X. C. J. Am. Chem. Soc. 2012, 134, 3015. (b) Jiang, D.-e.; Overbury, S. H.; Dai, S. J. Am. Chem. Soc. 2013, 135, 8786. (c) Tlahuice-Flores, A.; JoseYacamán, M.; Whetten, R. L. Phys. Chem. Chem. Phys. 2013, 15, 19557. (47) (a) Das, A.; Liu, C.; Byun, H. Y.; Nobusada, K.; Zhao, S.; Rosi, N. L.; Jin, R. Angew. Chem., Int. Ed. 2015, 54, 3140. (b) Chen, S.; Wang, S.; Zhong, J.; Song, Y.; Zhang, J.; Sheng, H.; Pei, Y.; Zhu, M. Angew. Chem., Int. Ed. 2015, 54, 3145. (48) Song, Y.; Wang, S.; Zhang, J.; Kang, X.; Chen, S.; Li, P.; Sheng, H.; Zhu, M. J. Am. Chem. Soc. 2014, 136, 2963. (49) Zeng, C.; Liu, C.; Chen, Y.; Rosi, N. L.; Jin, R. J. Am. Chem. Soc. 2014, 136, 11922. (50) Liu, P.; Qin, R.; Fu, G.; Zheng, N. J. Am. Chem. Soc. 2017, 139, 2122. (51) Tang, Q.; Jiang, D.-e. Chem. Mater. 2016, 28, 5976. (52) Kang, X.; Zhou, M.; Wang, S.; Jin, S.; Sun, G.; Zhu, M.; Jin, R. Chem. Sci. 2017, 8, 2581. (53) Konishi, K.; Iwasaki, M.; Sugiuchi, M.; Shichibu, Y. J. Phys. Chem. Lett. 2016, 7, 4267. (54) Wang, Y.; Wan, X.-K.; Ren, L.; Su, H.; Li, G.; Malola, S.; Lin, S.; Tang, Z.; Häkkinen, H.; Teo, B. K.; Wang, Q.-M.; Zheng, N. J. Am. Chem. Soc. 2016, 138, 3278. (55) Zhao, S.; Jin, R.; Abroshan, H.; Zeng, C.; Zhang, H.; House, S. D.; Gottlieb, E.; Kim, H. J.; Yang, J. C.; Jin, R. J. Am. Chem. Soc. 2017, 139, 1077. (56) Knoppe, S.; Bürgi, T. Acc. Chem. Res. 2014, 47, 1318. (57) (a) Zhu, M.; Qian, H.; Meng, X.; Jin, S.; Wu, Z.; Jin, R. Nano Lett. 2011, 11, 3963. (b) Zeng, C.; Chen, Y.; Liu, C.; Nobusada, K.; Rosi, N. L.; Jin, R. Sci. Adv. 2015, 1, e1500425. (58) Takano, S.; Tsukuda, T. J. Phys. Chem. Lett. 2016, 7, 4509. (59) Wu, Z.; Jin, R. Nano Lett. 2010, 10, 2568.

(60) Wang, T.; Wang, D.; Padelford, J. W.; Jiang, J.; Wang, G. J. Am. Chem. Soc. 2016, 138, 6380. (61) Pyo, K.; Thanthirige, V. D.; Kwak, K.; Pandurangan, P.; Ramakrishna, G.; Lee, D. J. Am. Chem. Soc. 2015, 137, 8244. (62) Mathew, A.; Natarajan, G.; Lehtovaara, L.; Häkkinen, H.; Kumar, R. M.; Subramanian, V.; Jaleel, A.; Pradeep, T. ACS Nano 2014, 8, 139. (63) Tlahuice-Flores, A.; Whetten, R. L.; Jose-Yacaman, M. J. Phys. Chem. C 2013, 117, 20867. (64) Yamazoe, S.; Takano, S.; Kurashige, W.; Yokoyama, T.; Nitta, K.; Negishi, Y.; Tsukuda, T. Nat. Commun. 2016, 7, 10414. (65) Chen, Y.; Liu, C.; Tang, Q.; Zeng, C.; Higaki, T.; Das, A.; Jiang, D.-e.; Rosi, N. L.; Jin, R. J. Am. Chem. Soc. 2016, 138, 1482. (66) (a) Pei, Y.; Zeng, X. C. Nanoscale 2012, 4, 4054. (b) Jiang, D.-e. Nanoscale 2013, 5, 7149. (c) Jin, R. Nanoscale 2015, 7, 1549. (d) Lei, Z.; Wan, X.-K.; Yuan, S.-F.; Wang, J.-Q.; Wang, Q.-M. Dalton Trans. 2017, 46, 3427. (67) (a) Li, G.; Zeng, C.; Jin, R. J. Am. Chem. Soc. 2014, 136, 3673. (b) Liu, C.; Lin, J.; Shi, Y.; Li, G. Nanoscale 2015, 7, 5987. (68) Nishio, M. Phys. Chem. Chem. Phys. 2011, 13, 13873. (69) Desiraju, G. R. Chem. Commun. 1997, 1475. (70) (a) Jiang, P.; Nion, A.; Marchenko, A.; Piot, L.; Fichou, D. J. Am. Chem. Soc. 2006, 128, 12390. (b) Ossowski, J.; Wächter, T.; Silies, L.; Kind, M.; Noworolska, A.; Blobner, F.; Gnatek, D.; Rysz, J.; Bolte, M.; Feulner, P.; Terfort, A.; Cyganik, P.; Zharnikov, M. ACS Nano 2015, 9, 4508. (71) Gao, Y.; Shao, N.; Zeng, X. C. ACS Nano 2008, 7, 1497. (72) Walter, M.; Akola, J.; Lopez-Acevedo, O.; Jadzinsky, P. D.; Calero, G.; Ackerson, C. J.; Whetten, R. L.; Grönbeck, H.; Häkkinen, H. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 9157. (73) Marks, L. D. Philos. Mag. A 1984, 49, 81. (74) Chen, Y.; Zeng, C.; Kauffman, D. R.; Jin, R. Nano Lett. 2015, 15, 3603. (75) Hulkko, E.; Lopez-Acevedo, O.; Koivisto, J.; Levi-Kalisman, Y.; Kornberg, R. D.; Pettersson, M.; Häkkinen, H. J. Am. Chem. Soc. 2011, 133, 3752. (76) Mustalahti, S.; Myllyperkiö, P.; Malola, S.; Lahtinen, T.; Salorinne, K.; Koivisto, J.; Häkkinen, H.; Pettersson, M. ACS Nano 2015, 9, 2328. (77) Yi, C.; Zheng, H.; Tvedte, L. M.; Ackerson, C. J.; Knappenberger, K. L. J. Phys. Chem. C 2015, 119, 6307. (78) Weerawardene, K. L D. M.; Aikens, C. M. J. Am. Chem. Soc. 2016, 138, 11202. (79) Li, Y.; Galli, G.; Gygi, F. ACS Nano 2008, 2, 1896. (80) Mustalahti, S.; Myllyperkiö, P.; Lahtinen, T.; Malola, S.; Salorinne, K.; Tero, T.-R.; Koivisto, J.; Häkkinen, H.; Pettersson, M. J. Phys. Chem. C 2015, 119, 20224.

10001

DOI: 10.1021/jacs.7b04678 J. Am. Chem. Soc. 2017, 139, 9994−10001