Tailoring the Core-Satellite Nanoassembly Architectures by Tuning

2 hours ago - We found that the pH of the assembly medium plays a vital role in the hierarchy of the nanoclusters. The architecture along with the siz...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/Langmuir

Cite This: Langmuir XXXX, XXX, XXX−XXX

Tailoring the Core−Satellite Nanoassembly Architectures by Tuning Internanoparticle Electrostatic Interactions Agampodi S. De Silva Indrasekara,†,‡,# Stephen J. Norton,‡ Nicholas K. Geitner,∥,⊥ Bridget M. Crawford,†,‡ Mark R. Wiesner,∥,⊥ and Tuan Vo-Dinh*,†,‡,§ Department of Biomedical Engineering, ‡Fitzpatrick Institute for Photonics, §Department of Chemistry, ∥Department of Civil and Environmental Engineering, and ⊥Center for the Environmental Implications of NanoTechnology, Duke University, Durham, North Carolina 27708, United States

Langmuir Downloaded from pubs.acs.org by KAOHSIUNG MEDICAL UNIV on 11/20/18. For personal use only.



S Supporting Information *

ABSTRACT: The use of plasmonic nanoplatforms has received increasing interest in a wide variety of fields ranging from theranostics to environmental sensing to plant biology. In particular, the development of plasmonic nanoparticles into ordered nanoclusters has been of special interest due to the new chemical functionalities and optical responses that they can introduce. However, achieving predetermined nanocluster architectures from bottom-up approaches in the colloidal solution state still remains a great challenge. Herein, we report a one-pot assembly approach that provides flexibility in precise control of core−satellite nanocluster architectures in the colloidal solution state. We found that the pH of the assembly medium plays a vital role in the hierarchy of the nanoclusters. The architecture along with the size of the satellite gold nanoparticles determines the optical responses of nanoclusters. Using electron microscopy and optical spectroscopy, we introduce a set of design rules for the synthesis of distinct architectures of silica-core gold satellites nanoclusters in the colloidal solution state. Our findings provide insight into advancing the colloidal solution state nanoclusters formation with predictable architectures and optical properties.



INTRODUCTION Noble metal nanoparticles (NPs) have attracted much attention in the field of optical sensing, imaging, and therapy largely due to their unique optical properties, in particular, localized surface plasmon resonance (LSPR).1−6 LSPR of noble metal nanoparticles can be tuned by varying the material composition, size, shape, and local environment.7−11 In addition, assembly of individual NPs has provided another means of optical tunability due to the coupling of surface plasmon of other NPs in close proximity.12−16 The degree of optical tunability of nanoparticle assemblies depends mainly on the interparticle distance, size, and number of constituent NPs as well as the chemical composition of the NPs.14,15,17−19 Many approaches ranging from bottom-up methods, such as electrostatic interactions,20−24 charge-induced dipole interactions,23,25,26 DNA hybridization,19,27−30 and molecularlymediated chemical bonding,12,31−37 to top-down methods33,38,39 have been previously used to develop noble-metal nanoparticle assemblies. Discrete assembled nanostructures, such as dimers and trimers, have been well explored both experimentally and theoretically to optimize the fabrication strategies and also to understand and optimize their optical properties.12,17,40−44 The extension of these discrete assembled nanostructures toward hierarchical structures with complex three dimensionalities is a trending subfield largely due to their ability to introduce chemical and optical versatilities. However, developing approaches to assemble NPs into well-defined, © XXXX American Chemical Society

predetermined architectures in a controlled fashion is still very challenging. Ensuring the reproducibility and colloidal stability for versatile applications of these for hierarchical nanostructures makes them more challenging. Assembling NPs into hierarchical structures widely known as “core−satellites” has been achieved both in colloidal solution state12,15,16,19,37 and colloids dispersed on substrates.14,18,45,46 However, the construction of well-defined, discrete core− satellite assemblies has mainly been demostrated by colloids dispersed on solid substrates.47,48 On the other hand, the fabrication complexity of this method has been an issue for the synthesis of hierarchical structures in the mid-to-large scale for their potential applications, such as imaging, sensing, and phototherapies. Here, we present a set of synthetic rules that advances the development of discrete hierarchical nanostructures of varying architectures in the colloidal solution state. We studied the assembly of optically active gold nanoparticles (AuNPs)“satellites”, onto optically transparent silica nanoparticles (SiNP)-“core”. Electrostatic interaction between AuNP surface and primary amine terminal groups projecting from SiNPs was used to create core−satellite nanostructures. The optical response of these core−satellite nanostructures, in other Received: August 16, 2018 Revised: October 31, 2018 Published: November 8, 2018 A

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 1. Architecture of nanoclusters is pH-dependent. Two-dimensional (2D) projections of TEM micrographs of (a) Raspberry-like assemblies where higher AuNP coverages on SiNP surface are dominant at acidic pH, (b) random assemblies of AuNPs that sparingly cover SiNP are dominant at neutral pH, and (c) bracelet-like assemblies where AuNPs predominantly arranged at the periphery of SiNPs are dominantly observed at basic pH. Inset scale bar: 20 nm. Preparation of Silica−Gold Core−Satellite Nanoclusters. Amine-terminated SiNPs (100 μL of 10 mg/mL) were pelleted by centrifugation at 10 000g for 5 min, resuspended in 1 mL of Mili-Q water, and then used as the working solution (1 mg/mL) for the assembly experiments. Working solution (100 μL) of SiNPs (1 mg/ mL) was used for each assembly experiment described herein. For reactions carried out under acidic conditions (pH 3), SiNPs were treated with 10 μL of 1 M HCl, and the reactions under basic conditions (pH 9) were obtained by adding 10 μL of 100 mM NaOH to SiNP. SiNP in water is used as it is for the reactions under neutral conditions (pH 7). The amount of AuNPs used for assembly experiments was varied between 400 μL and 1.5 mL. Briefly, 100 μL of the working stock solution of SiNP (1 mg/mL) was mixed with a given amount of AuNPs under the desired pH condition and incubated at room temperature under gentle shaking for 3 h. After that each reaction mixture was mixed with 40 μL of 2 mg/mL SH-PEG-COOH and incubated at room temperature under gentle shaking for another 3 h. Then, PEGylated silica−gold core− satellite assemblies were purified by centrifugation at 10 000g for 5 min, resuspended in 1 mL of Mili-Q water, and used for further characterization. For in situ growth of AuNP satellites on nanoclusters, 100 μL of the non-PEGylated nanoclusters was used as “seeds”. The seeds were mixed with 50 μL of 5 mM gold solution, and the growth of the satellite AuNPs was initiated by adding 50 μL of 100 mM ascorbic acid as the reducing agent. After 10 min of reaction, the resultant nanoclusters were PEGylated and purified by centrifugation, as described above. COMSOL Simulations of the Optical Properties of Silica− Gold Core−Satellite Nanostructures. The nanocluster configurations for the COMSOL calculations are based on model systems resembling approximately similar nanocluster configurations obtained from TEM, as shown in Figure 1. The COMSOL calculations are based on the finite-element calculation, which accounts for variations of the electric field within and between the particles. Here, we assume that (i) the E-field polarization is exactly in the plane of the ring, which may not be the case for the measurement and (ii) the data used for the dielectric constant are based on data from bulk material. In each case, there are NPs on the hidden side of the core that “mirror” the visible NPs. This was done for computational convenience; symmetry could then be employed to reduce the calculation time. For example, for the cluster shown in Figure 1a, 30 particles at random essentially tangent to the core particle (with 30 more particles on the

terms, the collective LSPR shift upon the assembly of individual AuNPs onto the SiNP core, was tuned by controlling the satellite coverage on the core as a function of solution pH and the size of the satellites. We studied the experimental condition-dependent architecture and the resultant optical response of these core−satellite nanostructures both experimentally and theoretically. Collectively, both the architecture and optical response of these core−satellite nanostructures provide insight into assembly-mediated optical sensing platform development.



EXPERIMENTAL SECTION

Materials. All of the chemicals used for this study were purchased and used without further purification. Gold chloride solution (5 mM) and sodium citrate were purchased from Sigma-Aldrich. Hydrochloric acid (HCl, 1 M) was purchased from VWR. Amine-terminated silica nanoparticles (SiNPs) (60 nm diameter, concentration = 10 mg/mL) were purchased from Nanocomposix, and citrate-capped gold colloid solutions (5 nm diameter, concentration = 5 × 1013 particles/mL; 10 nm diameter, concentration = 5 × 1013 particles/mL) were purchased from BBI Solutions. Thiol-poly(ethylene glycol) (SH-PEG-COOH MW 3400) was purchased from Nanocs Inc. Citrate-capped 15 nm diameter AuNPs (concentration = 1.5 × 1011 particles/mL) were synthesized according to a modified version of the Turkevich method.12,49 We selected the sizes of the core SiNP and the satellite AuNPs such that the overall size of the nanoclusters will be less than 100 nm in size for their foreseen biological applications. The COMSOL simulations also provided input to finalize the sizes of NPs for further experimentation. Characterization. All of the AuNPs, SiNPs, and silica−gold core−satellite nanoclusters were characterized for their morphology by transmission electron microscopy (TEM), FEI Tecnai G2 Twin transmission electron microscope, and the optical spectra were recorded using a multimode microplate reader, FLUOstar Omega (BMG Labtech). ζ-Potential measurements of AuNPs and SiNPs were performed using a Malvern Zetasizer Nano-S instrument. Three sequential measurements were collected at 25 °C, and the data were fit using Smoluchowski theory. Scattering measurements were taken using a hyperspectral transmission dark-field microscope from Cytoviva HSI. B

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir hidden side) are produced. The COMSOL calculations of the absorption cross-section were obtained by computing eq 1 given below.50 The calculations are based on a number of simplifying assumptions, e.g., they assume that the nanocluster is isolated (i.e., no interaction between different clusters is assumed), and the absorption of the host medium is not taken into account. σa =

n k 0εp″ E02εf

∫V |E(r)|2 dV p

The pH-dependent nanocluster architectures can be understood by the variation of surface charge on SiNP and AuNPs with pH, which in turn affects the type(s) and the strength of interparticle interactions leading to assemblies. We hypothesize that the final nanocluster architecture is a result of two competing forces between SiNPs and AuNPs: (1) molecular interactions between terminal primary amines on SiNPs and the citrate ligands and/or the surface atoms on AuNPs and (2) the steric hindrance between the neighboring AuNPs attached to SiNP. The ζ-potential measurements of both citrate-capped AuNPs and amine-terminated SiNPs at varying pHs (Table S1) considered in this manuscript are helpful to understand the variation of pH-dependent internanoparticle interactions hence nanocluster architectures. Therefore, the participating functional groups on NPs as well as their net surface charge under varying pH provide useful information to understand the pHdependent nanoclusters architectures. On the basis of the functionalization procedure provided by the manufacturer, it is estimated that there are, on average, 2.5 amine groups/nm2 available at the SiNP surface. Therefore, approximately 28 000 primary amines on a SiNP surface (60 nm in diameter) are estimated to be available to interact with AuNPs. However, it should be noted that only a portion of these amine groups would be accessible for interaction and conjugation to AuNPs due to the orientation, packing density and size of AuNPs, and the isoelectric point of the amines on SiNP, which is close to pH 7.5. On the basis of that and also as evident by the ζpotential measurements (Table S1), under the acidic and neutral pH conditions used in this study, SiNPs carry an overall positive surface charge, whereas basic pH conditions impart an overall negatively charged surface on SINP, i.e, the primary amines on SiNP could be considered to be completely protonated, whereas they are completely deprotonated at basic pH. On the other hand, as reflected by the ζ-potential measurements (Table S1), AuNPs exhibit a net negative surface charge at all of the pH conditions. Hence, citrate ligands on AuNPs can be expected to be mostly deprotonated, but to varying degrees at varying pH, whereas complete deprotonation at the basic pH. At acidic pH, electrostatic interactions between the positively charge SiNP (protonated primary amines on SiNP) and negatively charged AuNPs could be dominant and largely contributing to the nanocluster formation. This could explain the higher density of AuNPs present under acidic pH in comparison to other pHs that leads to raspberry-like nanoclusters. In addition, hydrogen bonding between the protonated primary amines on SiNP and citrate ligands on AuNPs could also have some contributions to the higher satellite density of nanoclusters. At basic pHs, where both SiNP and AuNPs exhibit net negative surface charge (Table S1), electrostatic repulsion between those nanoparticles could be considered to be dominant over the amine−gold weak covalent bonds leading to assemblies. The interplay between strong repulsive forces, weak covalent bonding along with the steric/electrostatic repulsion between relatively large AuNPs that already reside on SiNP (10 and 15 nm diameter) could therefore explain the unique bracelet-like core−satellite nanoclusters architecture, predominantly present at basic pH (Figure S2). Therefore, based on our observations, we speculate that stronger electrostatic interactions mainly drive the assembly at acidic pH, whereas the balance between electrostatic repulsion and weak covalent bonding determines the degree and orientation of assembly at basic pH. Neutral pH

(1)

The integration in eq 1 is over the volume of the interior of each particle in the cluster. COMSOL was used to compute the interior electric fields. In this equation, n is the refractive index of the fluid medium (water), k0 is the free-space wavenumber, E0 is the magnitude of the optical electric field incident on the particle, εf is the dielectric constant of the fluid, εp″ is the imaginary part of the particle’s dielectric constant, and E(r) is the electric field in the interior of the particle. In general, both the real and imaginary parts of the dielectric constant will depend on wavelength, and this dependence will influence the shape of the LSPR spectrum. In the COMSOL calculation, tabulated data for the dielectric function for gold were obtained from Johnson and Christy’s calculations.51



RESULTS AND DISCUSSION Core−satellite nanoclusters composed of SiNP (core) and AuNPs (satellites) are fabricated using an electrostatic and/or weakly covalent interaction-mediated, one-step assembly approach. SiNPs with terminal primary amine groups readily available for cross-linking and citrate-capped AuNPs with high affinity toward primary amine groups were employed as the building blocks. The sizes of the building blocks (SiNPs of 60 nm in diameter and AuNPs of 5, 10, and 15 nm in diameter; Figure S1) were chosen to achieve two main characteristics for the nanoclusters; (a) the overall size to not exceed 100 nm diameter and (b) their optical response to appear in the farvisible or near-infrared region for their potential applications in biological systems. In a typical assembly reaction, colloidal suspensions of SiNP and AuNPs are allowed to interact through molecular interactions, followed by poly(ethylene glycol) coating to ensure the colloidal stability of nanoclusters (Scheme 1). Scheme 1. Fabrication of Core−Satellite Nanoclusters Using Amine-Terminated Silica Nanoparticles (SiNP) and Gold Nanoparticles (AuNPs)

The architecture of SiNP−AuNP core−satellite nanoclusters is modulated by the pH of the assembly medium. We identified three different populations of nanocluster architectures that form preferably under basic, acidic, and neutral pH conditions (Figure 1). When the reaction medium was basic (pH 9), satellite AuNPs are orderly arranged on the core SiNPs resembling a “bracelet”. When the pH of the reaction medium was adjusted to pH 3, a dominant population of nanoclusters where a high density of satellites randomly arranged all over the core SiNP resembling a “raspberry” was observed. Under neutral conditions (pH 7), a population of transitional stage nanoclusters, which we refer to as “random” configuration, was observed. C

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir condition displays a mixed existence of these attractive and repulsive forces leading to “randomly” arranged nanoclusters. However, we believe that a more mechanistic understanding of the molecular forces and energies leading to the unique bracelet-like assembly architecture observed at basic pH could be realized with further fundamental studies and computation, which is beyond the scope of this study. The trends observed for the pH-dependent nanocluster architectures are consistent when 10 and 15 nm diameter AuNPs are used as satellites, but not for 5 nm diameter AuNPs. The architecture of nanoclusters obtained from 5 nm diameter AuNPs under all of the pHs and also at higher AuNP concentrations result random architectures. This observation suggests that the steric factors owing to the size of the satellites might also play a significant role in the nanocluster architectures. When the size ratio between the core and satellites is close to and below 5, the steric hindrance may start to play a significant role along with the interparticle interactions. The raspberry-like and bracelet-like assemblies are only prominent when the satellites are 10 and 15 nm AuNPs (Figure 1), but not for 5 nm AuNPs (Figure S3). To better understand the morphology of these nanoclusters, they were characterized using the tilt-angel TEM, in which the plane of observation was changed along the z-axis by 40 and 60° (Figure S4). The tilt-angle TEM micrographs demonstrate the three-dimensional architecture, particularly, of bracelet-like nanoclusters confirming their true architecture and ruling out that it is only a 2D view of nanoclusters. The architecture of core−satellite nanoclusters affects their optical response. Ensemble extinction spectra of nanoclusters were used to compare the changes in the localized surface plasmon resonance peak position (LSPR) of core−satellite nanoclusters in comparison to the constituent satellite AuNPs. Any changes in the LSPR could be accounted for and also used to understand the plasmonic coupling between the AuNPs because the SiNP is nonplasmonic and optically transparent. We observed no significant changes in the LSPR peak maxima for nanoclusters composed of 5 nm diameter AuNPs under all pHs and AuNP concentrations. The LSPR of these nanoclusters are centered at 518 nm similar to constituent 5 nm diameter AuNPs. This can be attributed to the large separation between AuNPs in the random nanocluster architecture, which supports no plasmon coupling between satellite AuNPs. In addition, no significant change in the average LSPR position was observed for the randomly arranged nanoclusters formed under neutral conditions with all of the sizes of AuNP satellites used in this study. Nanoclusters formed under basic pH and composed of 10 nm diameter AuNPs, on average 10 nm red shift of the LSPR peak position in the ensemble extinction spectrum was observed with respect to the constituent AuNPs (LSPRmax = 518 nm) at all of the concentrations considered in this study (Figure 2). This red shift of the LSPR peak position could be attributed to the coupling between AuNPs. The possibility of plasmon coupling between satellites could be validated by the nanocluster architecture that exhibit closely placed or near touching AuNP on the core particles (Figure 1). Red shift of the LSPR peak position is also observed for the nanoclusters formed under acidic pH, and it also varies with the concentration of AuNPs (Figure 2b). When the 10 nm AuNP concentration is below 1.5 μM, no significant changes in the LSPR peak maximum of the clusters were observed. But when the concentrations of AuNPs were 1.5 and 2.0 μM, on

Figure 2. Optical response of nanoclusters is influenced by the pH and the concentration of the AuNP in the assembly reaction medium. (a) Ensemble extinction spectra and (b) LSPRmax peak position of nanoclusters as a function of the satellite AuNP (d = 10 nm) concentration at acidic (pH 3) neutral (pH 7), and basic (pH 9) conditions.

average 40 nm red shift of the LSPR peak position was observed. This could be attributed to the larger number of AuNPs assembled on SiNP, facilitating plasmon coupling to a greater extent. However, no changes in the LSPR peak position were observed for the random nanoclusters formed at pH 7 and composed of 10 nm diameter satellite AuNPs. Nanoclusters composed of 15 nm diameter AuNPs exhibit the largest variation in the ensemble optical response. On average, 4 nm red shift of the LSPR peak position was observed for the bracelet-like nanoclusters (Figure S5). Moreover, the largest red shift of the LSPR peak position among all of the types of nanoclusters, which is, on average, 40 nm was observed for the raspberry-like nanoclusters prepared under acidic pH. This large red shift could be attributed to strong plasmon coupling between larger satellite AuNPs. Another noticeable change is the plasmon peak broadening as the satellite AuNP density on the core increases. This can be attributed to the retardation effect, as the size of the nanoclusters becomes significant compared to the wavelength of light. COMSOL calculation and experimental single-particle darkfield spectral measurements demonstrate the architectureoptical response relationship. For a more accurate comparison of theoretical and experimental spectra of nanoclusters, singleparticle dark-field spectral measurements for nanocluster composed of 10 nm diameter satellite AuNPs were acquired. Absorption cross-section and LSPR maxima of selected nanoclusters as a function of their architecture and AuNP size were measured. Except for the bracelet-like architectures, the calculated and experimental LSPR peak positions for nanoclusters composed of 10 nm AuNP satellites are in good agreement (Figure 3). Studies with dimers have shown that an E-field polarization aligned along a particle gap will red shift the LSPR compared to a transverse alignment. This alignment along the gaps occurs more frequently when the polarization is in the plane of the bracelet and likely contributes to the red shifting shown in Figure 3b. The less shifted spectra of Figure 3b could be attributed to the heterogeneity of the nanoclusters present in the samples that does not represent the exact architecture used for simulation and also the polarization-averaged readout in the ensemble extinction spectra. In addition, the calculated spectra of the randomly arranged nanoclusters with either 10 or 15 nm AuNPs exhibit the LSPR peak position at 530 nm, whereas the scattering cross-section of the latter is much higher, which can be attributed to the larger size of satellite AuNPs (Figure S6). D

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 3. Core−satellite assembly architecture affects their optical response. COMSOL-calculated optical spectra of (a) raspberry-like assemblies AuNPs (LSPRmax = 520 nm) and randomly arranged clusters with 10 nm (LSPRmax = 530 nm), (b) optical response of bracelet-like nanoclusters with 10 nm diameter AuNPs is dependent with respect to the incident light. Absorption cross-section versus light moving horizontally when the Efield polarized vertically in the plane of AuNPs (solid line, LSPRmax = 520 and 640 nm) and the E-field polarized normal to the plane of AuNPs (dotted line, LSPRmax = 520 nm). The number and the arrangement of AuNPs of the model nanoclusters are matched to the representative clusters shown in Figure 1. (c) Representative single-particle experimental dark-field scattering spectra of a bracelet-like (LSPRmax = 630 nm), a raspberrylike (LSPRmax = 590 nm), and a randomly arranged nanocluster (LSPRmax = 577 nm) composed 10 nm diameter AuNPs.

nanoclusters could be adapted and facilitate more fundamental understanding of the optical properties of functional hierarchical nanoclusters. These nanoclusters can also be further developed and used as optical nanosensors for their application in biological systems.

Representative experimental single-particle optical spectrum of a bracelet-like nanocluster composed of 10 nm diameter satellite AuNPs provides one plasmon peak centered around 630 nm. The LSPR peak position from single-particle darkfield spectroscopy agrees better with the COMSOL calculation over the ensemble optical spectral measurements. In addition, the single-particle spectrum of a raspberry-like nanocluster composed of 10 nm diameter AuNP satellites provides an LSPR maximum at 590 nm, which is more red shifted than both calculated and ensemble spectral measurements. This observation can be attributed to the high density of AuNPs satellites that may be supporting plasmon coupling to a greater extent than the model nanocluster architecture used for calculations. In situ growth of satellites provides promise for the further modulations of nanoclusters architectures. Our ultimate goal is to develop far-visible or near-IR absorbing nanoclusters with an overall size suitable for biological applications. Therefore, we explored experimental modifications that could improve the optical response of nanoclusters. Our preliminary data from in situ growth of satellite AuNPs on nanoclusters show great promise for achieving the desired morphological characteristics. AuNPs on nanoclusters were used as seeds in the presence of excess gold chlorides and ascorbic acid, a strong reducing agent, which promoted the in situ growth of AuNPs. This leads to the transformation of nanoclusters architectures, where AuNPs are larger in size and in physical contact with each other (Figure S7) yielding more red-shifted ensemble LSPR peak position than the starting nanoassembly. In depth investigation of this approach will be considered in a future manuscript.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.langmuir.8b02792. Experimental section; characterization of starting nanomaterials and nanoclusters (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Agampodi S. De Silva Indrasekara: 0000-0002-9731-7885 Nicholas K. Geitner: 0000-0003-4313-372X Mark R. Wiesner: 0000-0001-7152-7852 Tuan Vo-Dinh: 0000-0003-3701-3326 Present Address

# Department of Chemistry, University North Carolina at Charlotte, 9201 University City Blvd, Charlotte, North Carolina 28223, United States (A.S.D.S.I.).

Funding

This work is supported by the U.S. Department of Energy Office of Science under Award Number DE-SC0014077.



CONCLUSIONS In conclusion, by taking the simplified fabrication strategy and design rules presented in this study into account, more controllable, predetermined nanoparticle assembly architectures could be fabricated in the colloidal solution state. The pH and size of AuNPs can be used as synthetic parameters to tailor the nanocluster architecture and optical response, respectively. More importantly, changes in the LSPR could be accounted for by the plasmonic coupling between the satellite AuNPs assembled on the nonplasmonic SiNP core. Therefore, our approach and design of varying architectures of core−satellite

Notes

The authors declare no competing financial interest.



REFERENCES

(1) Alkilany, A. M.; Thompson, L. B.; Boulos, S. P.; Sisco, P. N.; Murphy, C. J. Gold nanorods: Their potential for photothermal therapeutics and drug delivery, tempered by the complexity of their biological interactions. Adv. Drug Delivery Rev. 2012, 64, 190−199. (2) Huang, X.; El-Sayed, M. A. Plasmonic photo-thermal therapy (PPTT). Alexandria J. Med. 2011, 47, 1−9. (3) Indrasekara, A. S. D. S.; Meyers, S.; Shubeita, S.; Feldman, L. C.; Gustafsson, T.; Fabris, L. Gold nanostar substrates for SERS-based

E

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir chemical sensing in the femtomolar regime. Nanoscale 2014, 6, 8891− 8899. (4) Perets, E. A.; Indrasekara, A. S. D. S.; Kurmis, A.; Atlasevich, N.; Fabris, L.; Arslanoglu, J. Carboxy-terminated immuno-SERS tags overcome non-specific aggregation for the robust detection and localization of organic media in artworks. Analyst 2015, 140, 5971− 5980. (5) Vo-Dinh, T.; Dhawan, A.; Norton, S. J.; Khoury, C. G.; Wang, H.; Misra, V.; Gerhold, M. D. Plasmonic Nanoparticles and Nanowires: Design, Fabrication and Application in Sensing. J. Phys. Chem. C 2010, 114, 7480−7488. (6) Wang, H.-N.; Vo-Dinh, T. Plasmonic Coupling Interference (PCI) Nanoprobes for Nucleic Acid Detection. Small 2011, 7, 3067− 3074. (7) De Silva Indrasekara, A. S.; Johnson, S. F.; Odion, R. A.; VoDinh, T. Manipulation of the Geometry and Modulation of the Optical Response of Surfactant-Free Gold Nanostars: A Systematic Bottom-Up Synthesis. ACS Omega 2018, 3, 2202−2210. (8) Kelly, K.; Coronado, E.; Zhao, L.; Schatz, G. The optical properties of metal nanoparticles: The influence of size, shape, and dielectric environment. J. Phys. Chem. B 2003, 107, 668−677. (9) Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J.; Gou, L.; Hunyadi, S. E.; Li, T. Anisotropic Metal Nanoparticles: Synthesis, Assembly, and Optical Applications. J. Phys. Chem. B 2005, 109, 13857−13870. (10) Hao, E.; Schatz, G.; Hupp, J. Synthesis and Optical Properties of Anisotropic Metal Nanoparticles. J. Fluoresc. 2004, 14, 331−341. (11) Khoury, C. G.; Vo-Dinh, T. Gold Nanostars For SurfaceEnhanced Raman Scattering: Synthesis, Characterization and Optimization. J. Phys. Chem. C 2008, 112, 18849−18859. (12) Indrasekara, A. S. D. S.; Paladini, B. J.; Naczynski, D. J.; Starovoytov, V.; Moghe, P. V.; Fabris, L. Dimeric Gold Nanoparticle Assemblies as Tags for SERS-Based Cancer Detection. Adv. Healthcare Mater. 2013, 2, 1370−1376. (13) Lee, A.; Andrade, G. F.; Ahmed, A.; Souza, M. L.; Coombs, N.; Tumarkin, E.; Liu, K.; Gordon, R.; Brolo, A. G.; Kumacheva, E. Probing dynamic generation of hot-spots in self-assembled chains of gold nanorods by surface-enhanced Raman scattering. J. Am. Chem. Soc. 2011, 133, 7563−7570. (14) Yoon, J. H.; Lim, J.; Yoon, S. Controlled Assembly and Plasmonic Properties of Asymmetric Core-Satellite Nanoassemblies. ACS Nano 2012, 6, 7199−7208. (15) Gandra, N.; Abbas, A.; Tian, L.; Singamaneni, S. Plasmonic Planet-Satellite Analogues: Hierarchical Self-Assembly of Gold Nanostructures. Nano Lett. 2012, 12, 2645−2651. (16) Xu, L.; Kuang, H.; Xu, C.; Ma, W.; Wang, L.; Kotov, N. A. Regiospecific Plasmonic Assemblies for in Situ Raman Spectroscopy in Live Cells. J. Am. Chem. Soc. 2012, 134, 1699−1709. (17) Mark, P. R.; Fabris, L. Understanding nanoparticle assembly: A simulation approach to SERS-active dimers. J. Colloid Interface Sci. 2012, 369, 134−143. (18) Yoon, J. H.; Yoon, S. Probing Interfacial Interactions Using Core-Satellite Plasmon Rulers. Langmuir 2013, 29, 14772−14778. (19) Chou, L. Y. T.; Song, F.; Chan, W. C. W. Engineering the Structure and Properties of DNA-Nanoparticle Superstructures Using Polyvalent Counterions. J. Am. Chem. Soc. 2016, 138, 4565−4572. (20) Romo-Herrera, J. M.; Alvarez-Puebla, R.; Liz-Marzan, L. Controlled assembly of plasmonic colloidal nanoparticle clusters. Nanoscale 2011, 3, 1304−1315. (21) Wang, X.; Li, G.; Chen, T.; Yang, M.; Zhang, Z.; Wu, T.; Chen, H. Polymer-Encapsulated Gold-Nanoparticle Dimers: Facile Preparation and Catalytical Application in Guided Growth of Dimeric ZnONanowires. Nano Lett. 2008, 8, 2643−2647. (22) Gschneidtner, T. A.; Fernandez, Y. A. D.; Syrenova, S.; Westerlund, F.; Langhammer, C.; Moth-Poulsen, K. A Versatile SelfAssembly Strategy for the Synthesis of Shape-Selected Colloidal Noble Metal Nanoparticle Heterodimers. Langmuir 2014, 30, 3041− 3050.

(23) Cheng, L.; Song, J.; Yin, J.; Duan, H. Self-Assembled Plasmonic Dimers of Amphiphilic Gold Nanocrystals. J. Phys. Chem. Lett. 2011, 2, 2258−2262. (24) Zohar, N.; Haran, G. Modular Plasmonic Antennas Built of Ultrathin Silica-Shell Silver-Core Nanoparticles. Langmuir 2014, 30, 7919−7927. (25) Chen, G.; Wang, Y.; Tan, L. H.; Yang, M.; Tan, L. S.; Chen, Y.; Chen, H. High-Purity Separation of Gold Nanoparticle Dimers and Trimers. J. Am. Chem. Soc. 2009, 131, 4218−4219. (26) Ringler, M.; Schwemer, A.; Wunderlich, M.; Nichtl, A.; Karzinger, K.; Klar, T. A.; Feldmann, J. Shaping Emission Spectra of Fluorescent Molecules with Single Plasmonic Nanoresonators. Phys. Rev. Lett. 2008, 100, No. 203002. (27) Zheng, Y.; Thai, T.; Reineck, P.; Qiu, L.; Guo, Y.; Bach, U. DNA-Directed Self-Assembly of Core-Satellite Plasmonic Nanostructures: A Highly Sensitive and Reproducible Near-IR SERS Sensor. Adv. Funct. Mater. 2013, 23, 1519−1526. (28) Loweth, C. J.; Caldwell, W. B.; Peng, X.; Alivisatos, A. P.; Schultz, P. G. DNA-Based Assembly of Gold Nanocrystals. Angew. Chem., Int. Ed. 1999, 38, 1808−1812. (29) Zanchet, D.; Micheel, C. M.; Parak, W. J.; Gerion, D.; Williams, S. C.; Alivisatos, A. P. Electrophoretic and Structural Studies of DNADirected Au Nanoparticle Groupings. J. Phys. Chem. B 2002, 106, 11758−11763. (30) Chen, J. I. L.; Chen, Y.; Ginger, D. S. Plasmonic Nanoparticle Dimers for Optical Sensing of DNA in Complex Media. J. Am. Chem. Soc. 2010, 132, 9600−9601. (31) Indrasekara, A. S. D. S.; Thomas, R.; Fabris, L. Plasmonic properties of regiospecific core-satellite assemblies of gold nanostars and nanospheres. Phys. Chem. Chem. Phys. 2015, 17, 21133−21142. (32) Novak, J. P.; Nickerson, C.; Franzen, S.; Feldheim, D. L. Purification of Molecularly Bridged Metal Nanoparticle Arrays by Centrifugation and Size Exclusion Chromatography. Anal. Chem. 2001, 73, 5758−5761. (33) Kinkhabwala, A.; Yu, Z.; Fan, S.; Avlasevich, Y.; Mullen, K.; Moerner, W. E. Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna. Nat. Photonics 2009, 3, 654−657. (34) Guarrotxena, N.; Bazan, G. C. Antibody-functionalized SERS tags with improved sensitivity. Chem. Commun. 2011, 47, 8784−8786. (35) Braun, G. B.; Lee, S. J.; Laurence, T.; Fera, N.; Fabris, L.; Bazan, G. C.; Moskovits, M.; Reich, N. O. Generalized Approach to SERS-Active Nanomaterials via Controlled Nanoparticle Linking, Polymer Encapsulation, and Small-Molecule Infusion. J. Phys. Chem. C 2009, 113, 13622−13629. (36) Fabris, L.; Schierhorn, M.; Moskovits, M.; Bazan, G. C. Aptatag-Based Multiplexed Assay for Protein Detection by SurfaceEnhanced Raman Spectroscopy. Small 2010, 6, 1550−1557. (37) Schütz, M.; Schlucker, S. Molecularly linked 3D plasmonic nanoparticle core/satellite assemblies: SERS nanotags with singleparticle Raman sensitivity. Phys. Chem. Chem. Phys. 2015, 17, 24356− 24360. (38) Zhu, W.; Banaee, M. G.; Dongxing, W.; Yizhuo, C.; Crozier, K. B. Lithographically Fabricated Optical Antennas with Gaps Well Below 10 nm. Small 2011, 7, 1761−1766. (39) Chirumamilla, M.; Andrea, T.; Anisha, G.; Gobind, D.; Zaccaria, R. P.; Roman, K.; Eliana, R.; Marco, L.; Carlo, L.; De, A. F.; Di, F. E. 3D Nanostar Dimers with a Sub 10 nm Gap for Single/ Few Molecule Surface Enhanced Raman Scattering. Adv. Mater. 2014, 26, 2353−2358. (40) Nordlander, P.; Oubre, C.; Prodan, E.; Li, K.; Stockman, M. I. Plasmon Hybridization in Nanoparticle Dimers. Nano Lett. 2004, 4, 899−903. (41) Chuntonov, L.; Haran, G. Trimeric Plasmonic Molecules: The Role of Symmetry. Nano Lett. 2011, 11, 2440−2445. (42) Norton, S. J.; Vo-Dinh, T. Optical response of linear chains of metal nanospheres and nanospheroids. J. Opt. Soc. Am. A 2008, 25, 2767−2775. F

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir (43) Khoury, C. G.; Norton, S. J.; Vo-Dinh, T. Plasmonics of 3-D Nanoshell Dimers Using Multipole Expansion and Finite Element Method. ACS Nano 2009, 3, 2776−2788. (44) Dhawan, A.; Norton, S. J.; Gerhold, M. D.; Vo-Dinh, T. Comparison of FDTD numerical computations and analytical multipole expansion method for plasmonics-active nanosphere dimers. Opt. Express 2009, 17, 9688−9703. (45) Hofmann, A.; Schmiel, P.; Stein, B.; Graf, C. Controlled Formation of Gold Nanoparticle Dimers Using Multivalent Thiol Ligands. Langmuir 2011, 27, 15165−15175. (46) Sardar, R.; Heap, T. B.; Shumaker-Parry, J. Versatile Solid Phase Synthesis of Gold Nanoparticle Dimers Using an Asymmetric Functionalization Approach. J. Am. Chem. Soc. 2007, 129, 5356− 5357. (47) Tanaka, Y.; Matsumoto, A.; Tsuruoka, T.; Nawafune, H.; Akamatsu, K. Gold Nanoparticle Satellite Nanostructures on a Glass Substrate for Dual-responsive Optical Sensing of Organic Liquids. Chem. Lett. 2014, 43, 1084−1085. (48) Cha, H.; Yoon, J. H.; Yoon, S. Probing Quantum Plasmon Coupling Using Gold Nanoparticle Dimers with Tunable Interparticle Distances Down to the Subnanometer Range. ACS Nano 2014, 8, 8554−8563. (49) Turkevich, J.; Stevenson, P. C.; Hillier, J. A study of the nucleation and growth processes in the synthesis of colloidal gold. Discuss. Faraday Soc. 1951, 11, 55−75. (50) Ishimaru, A. Electromagnetic Wave Propagation, Radiation and Scattering; Englewood Cliffs, 1991. (51) Johnson, P. B.; Christy, R. W. Optical Constants of the Noble Metals. Phys. Rev. B 1972, 6, 4370−4379.

G

DOI: 10.1021/acs.langmuir.8b02792 Langmuir XXXX, XXX, XXX−XXX