Preparation of Anisotropic and Oriented Particles ... - ACS Publications

Nov 17, 2015 - Department of Physical Chemistry, University of Geneva, 30 Quai ... Matter Physics, Laboratory of Crystallography, University of Geneva...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/Langmuir

Preparation of Anisotropic and Oriented Particles on a Flexible Substrate Mahshid Chekini,† Ugo Cataldi,† Plinio Maroni,‡ Laure Guénée,§ Radovan Č erný,§ and Thomas Bürgi*,† †

Department of Physical Chemistry, University of Geneva, 30 Quai Ernest-Ansermet, 1211 Geneva, Switzerland Department of Inorganic and Analytic Chemistry, University of Geneva, 30 Quai Ernest-Ansermet, 1211 Geneva, Switzerland § Department of Quantum Matter Physics, Laboratory of Crystallography, University of Geneva, 24 Quai Ernest-Ansermet, 1211 Geneva, Switzerland ‡

S Supporting Information *

ABSTRACT: Elongated plasmonic nanoparticles show superior optical properties when compared to spherical ones. Facile, versatile and cost-effective bottom-up approaches for fabrication of anisotropic nanoparticles in solution have been developed. However, fabrication of 2-D plasmonic templates from elongated nanoparticles with spatial arrangement at the surface is still a challenge. We used controlled seed-mediated growth in the presence of porous and functionalized surface of flexible polydimethylsiloxane (PDMS) templates to provide directional growth and formation of elongated gold nanoparticles (AuNPs). Atomic force microscopy (AFM) and spectroscopy revealed embedding of the particles within the functionalized porous surface of PDMS. Nanoparticles shapes were observed with transmission electron microscope (TEM), UV−Vis spectroscopy, and X-ray powder diffraction (XRPD) measurements, which revealed an overall orientation of particles at the surface. Anisotropic and oriented particles on a flexible substrate are of interest for sensing applications.

1. INTRODUCTION

alterations of the size, shape and even composition (alloys) of surface immobilized particles and provides better control over large tuning of interparticle distances, particularly by introducing several growth steps.34−43 The assembled nanoparticles can form a well-defined surface for further crystal growth. Since there is no lattice mismatch for growing gold on gold crystal, even three-dimensional layer-by-layer epitaxial growth has been reported.44 Directional growth and formation of anisotropic nanostructures in solution has been widely investigated.21,22,45−48 Directional growth is introduced by surfactant and silver ions based on their interaction with more energetic crystalline facets ((110) and (100)) of gold nanoparticles. Passivation of these surfaces along with favorable growth of more stable and intact (111) crystalline facets lead to spikes or multipods formation. These spikes are mainly directed in the (111) lattice plane.49,50

Plasmonic nanomaterials have found increasing attention in recent years. Their plamon resonances and field enhancement properties have encouraged applications in optoelectronics,1−4 sensing,5−9 metamaterials and enhanced spectroscopy.9−12 The plasmonic properties of nanoparticle arrays depend on their size, shape, and interparticle distances,13−18 with shape control as one of the main factors affecting the optical properties of the sample.15,16,19,20 Introducing anisotropy into the shape of plasmonic nanoparticles changes their optical properties.21−25 Elongated nanoparticles exhibit different plasmon resonances.9,21,22,24 For example, nanorods have longitudinal and transverse plasmon bands (along the long and short axes of the particle, respectively).22,24−28 In bottomup approaches, the templating route is commonly used for fabrication of immobilized anisotropic metallic nanostructures such as nanorods or nanowires at surfaces.29−33 Another facile and cost-effective method is the seed-mediated growth19 with the main drawback of uncontrolled growth orientation and alignment at the surface.20 This method enables further © XXXX American Chemical Society

Received: September 21, 2015 Revised: November 16, 2015

A

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 1. AFM topography image of a plasma-treated and aminosilane-functionalized PDMS surface, analyzed by larger tip cantilever (Olympus OMCL-AC240TS cantilever with a 9 ± 2 nm tip) in panel a, and a slightly stretched (stretching less than 5%) surface analyzed by sharper tip cantilever (SuperSharpSilicon SSS-NCH cantilever with 2 nm tip radius) in panel c. Corresponding phase images are presented in b and d, respectively.

The integration of nanomaterials with flexible matrices, thus combining optical responses and mechanical flexibility, has recently attracted attention.10,38,51−55 Active control of the interparticle distance under stretching and mechanical stress can result in optical tunability and provide a tunable platform for enhanced spectroscopy.38,51,52 Stretching the amorphous and isotropic flexible templates brings the particles closer together in perpendicular direction to the applied stretch. Decreasing the gap sizes between neighboring particles induces coupling in closely spaced plasmonic nanoparticles.38 Coupling between plasmonic particles shifts their plasmon band and results in reversible color change of the templates. However, the particle size and therefore their interparticle distances have to be adjusted, which can be achieved by seed-mediated growth. The realized reversible and multipolar coupling effect and its related alterations upon stretching are polarization dependent.38 Here we report the formation of elongated gold nanoparticles by seed-mediated method after several growth steps at the surface of PDMS templates. We applied AFM and electron microscopy techniques for particle size, shape, and surface topography characterizations. Atomic force microscopy (AFM) studies revealed the porosity of PDMS surface and its potential hosting role for nanoparticles embedding after aminosilanefunctionalization. As a result, the particles grew anisotropically as revealed by transmission electron microscopy (TEM). X-ray

powder diffraction analysis (XRPD) was applied as a promising tool to investigate preferential growth orientation.30−33,42,43,56 The XRPD patterns revealed some changes by disturbing the surface as a result of induced alterations in particles arrangement. Although the TEM measurements revealed the formation of elongated nanoparticles, the measured extinction spectra of the samples only showed one peak related to transverse plasmon resonance.

2. EXPERIMENTAL SECTION 2. 1. Gold Nanoparticle Solution. Solution of 17−19 nm (in diameter) colloidal gold nanoparticles (AuNPs) was synthesized based on the standard Turkevich method.57 Briefly, the gold salt solution (HAuCl4) (600 mL of 2.5 × 10−4 M) was brought to 100 °C under constant stirring in an oil bath. The gold ions were reduced by introducing 15 mL of a 0.03 M sodium citrate solution to the boiling gold salt solution. The red-colored solution was kept at 100 °C for another 30 min before removing from the oil bath. 2. 2. Polydimethylsiloxane (PDMS) Template Fabrication. PDMS is used as a cost-effective, nontoxic, and easy to use elastomer with chemical stability, and high flexibility besides transparency in the visible region.58−61 The sample fabrication method is explained elsewhere in detail.38 In the first step, the solid PDMS films were fabricated. Briefly the mixture of polydimethylsiloxane elastomer and curing agent (10:1 weight ratio, SYLGARD 184 from Dow Corning) were placed in a glass Petri dish to rest at room temperature for 2 h followed by heating at 80 °C for 1 B

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 2. AFM topography image of AuNP attached to PDMS surface (pristine gold nanoparticles deposited on the functionalized PDMS sample as seed with no growth) in a, and its phase image in b, analyzed by a sharper tip cantilever (SuperSharpSilicon SSS-NCH cantilever with 2 nm tip radius), and (c) topographic (d phase) image analyzed by a larger tip cantilever (Olympus OMCL-AC240TS cantilever with a 9 ± 2 nm tip). h. After cooling to room temperature the films were peeled off, and pieces of desired size (26 mm × 16.6 mm × 1.2 mm) were cut out. 2. 3. Surface Functionalization. The PDMS samples were exposed to an air plasma (806 mTorr with a constant flux of air of 31.94 mL/min, 90 s, 7.2 W), which oxidizes the surface. Surface functionalization was carried out by 30 min immersion of the samples in 1% aqueous solution of 3-aminopropyltriethoxysilane (APTES) followed by rinsing with an excessive amount of Milli-Q water (18.2 MΩ cm) and drying under a stream of compressed air. 2. 4. Gold Seed Deposition. The deposition of AuNPs on the surface is charge-driven and relies on forces between the negatively charged citrate-caped AuNPs and positively charged aminosilanefunctionalized surface. The aminosilane-functionalized samples were placed in the colloidal solution of AuNPs for 150 min, followed by rinsing with Milli-Q water and drying under a stream of compressed air. 2. 5. Growth step. Growth steps were carried out by 2 min dipping of the samples in growth solution containing 0.3 mM HAuCl4 and 0.4 mM NH2OH, followed by washing and drying. The samples were stretched and released prior to each growth step. The control sample was prepared on a glass slide cleaned by Piranha solution (3 to 1 mixture of concentrated sulfuric acid and 30% hydrogen peroxide; piranha solution is dangerous and should be handled carefully). After plasma treatment, the glass slide was functionalized with APTES following the same procedure used for PDMS samples. 2. 6. Atomic Force Microscopy (AFM). Surface topography of the samples was measured by AFM (Asylum Research-Oxford, in MFP-3D in AM-AFM mode). We used two types of cantilevers Olympus OMCL-AC240TS and Nanosensors SuperSharpSilicon SSSNCH. The used parameters for Olympus OMCL-AC240TS-R3 are as

follows: free oscillation amplitude (FOA) 35 nm, set point 75% of FOA, resonance frequency 70 kHz and spring constant 2 N/m, scan rate 1 Hz, scan size from 0.4 × 0.4 μm2 to 2 × 2 μm2, with 512 lines and 512 points per line. For SuperSharpSilicon SSS-NCH we used the following parameters: FOA 40 nm, set point 90% of FOA, resonance frequency 300 kHz and spring constant 30 N/m, scan rate 0.5 Hz, scan size 0.4 × 0.4 μm2, with 256 lines and 256 points per line, and for stretched samples scan rate 0.1 Hz, scan size 2 × 2 μm2, with 512 lines and 512 points per line. The AFM images were improved using FFT filtering (Gwyddion 2.32 free SPM software) and changing the contrast. 2. 7. X-ray Powder Diffraction Analysis (XRPD). Samples were placed on a flat sample holder, and X-ray diffraction patterns were measured using an Empyrean (PANalytical) diffractometer in reflection mode with a monochromated CuKα1 radiation (Johansson type Ge monochromator) and a PIXcel3D area detector. Continuous scan from 30° to 90° 2 theta with step size 0.0131° were performed. 2. 8. Transmission electron microscopy (TEM). TEM analyses were carried out by FEI Tecnai G2 Sphera operating at 200 kV. 2. 9. Ultraviolet−visible spectroscopy (UV−Vis). UV−Vis spectroscopy were carried out on a Jasco V-670 UV−Vis spectrophotometer.

3. RESULTS AND DISCUSSION AuNPs deposition on PDMS surface is carried out using a selfassembly method. Since the charge-driven electrostatic attraction promotes AuNPs deposition, the PDMS surface treatment and functionalization is needed (as described in the Experimental Section). These processes not only change the chemical state of the surface layer but also affect the surface C

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

stretching, the particles can be displaced and partly enter into the porous surface. The AuNP growth was carried out in the presence of hydroxylamine, which has been used for synthesis and growth of nanoparticles,68−71 and gold salt solution. Hydroxylamine can reduce Au (III) ions at room temperature.71,72 In acidic conditions, hydroxylamine is oxidized by HAuCl4 only in the presence of a gold surface, i.e., the reduction of Au (III) is catalyzed by the AuNPs.42,69,70 The less stiff Olympus OMCL-AC240TS cantilever is used to study the soft texture of the surface. AFM images of the grown sample (after 12 steps of growing) are presented in Figure 4. These images show the presence of a coarse covering around each AuNP, and some particles are covered due to the embedding. Figure S2 (Supporting Information) shows the 3D representation of particles in Figure 4e, which evidence the presence of the concave part at the top of the structures. It should be noted that these concave structures were observed with several cantilever tips. The AFM images acquired by the sharper tip cantilever (SuperSharpSilicon SSS-NCH), presented in Figure S3 (Supporting Information), show the distribution and coverage of the AuNPs on the surface. Unfortunately these images cannot give a clear idea about the shape and orientation of the particles, but it is estimated that the surface cross-section of the majority of the particles are less than 50 nm. The X-ray powder diffraction patterns of the sample with seed particles and the sample subjected to 12 growth steps are presented in Figure 5. The XRPD pattern shows that the AuNPs are composed of crystalline gold73−75 with a facecentered cubic ( fcc) cell. The difference in the peak intensity of the seed and the grown sample indicates the preferred orientation of grains in the sample, which may be caused by the anisotropic shape of the crystallites. For citrate-capped gold seeds, which have a multiply twinned decahedral structure (five-twin boundaries with all crystal faces formed by {111} planes),76 careful choice of the experimental conditions allows anisotropic growth and formation of rod shape structures.19 Moreover, the presence and orientation of twin planes can direct the shape of the growing particles77 and lead to formation of elongated particles, rods, or wires with 5fold symmetry and five-face facet ends.77,78 These characteristics along with the embedding effect of the surface can introduce a directional growth and anisotropic shape in nanoparticles. XRPD investigation can provide more evidence of preferential orientation. The intensity ratio of the 200 to 111 diffraction peaks in comparison with the conventional bulk sample can be used as an indicator of the predominant orientation. The relative peak area measurement shows the difference in these nanoparticle templates compared to the bulk reference samples, as has been reported in preferential orientation or elongated nanoparticles.41,43,49,50 Significant changes were observed by disturbing the surface (by touching and pressing the surface layer). The integrated intensities of 200, 220, and 311 peaks relative to the 111 peak are presented in Figure 6 for different samples. When this ratio is less than the corresponding bulk value (about 0.5), it suggests that {111} planes are more abundant and, on average, the ⟨111⟩ direction is the preferential growth orientation.73−75,79 This can be explained by considering that the low-energy planes of fcc geometry are more stable.49,61 These surfaces are less occupied by other species like amine and silane groups,

properties. AFM analyses showed the presence of porous surface structure on a plasma-treated and aminosilane-functionalized PDMS surface. Figure 1 shows the AFM images of the analyzed surface by two different cantilevers. The surface height and phase images, acquired with an Olympus OMCLAC240TS cantilever with a 9 ± 2 nm tip radius, are shown in Figure 1a,b. Figure 1c,d shows the height and phase images of slightly stretched surface (stretching less than 5%) analyzed by a sharper tip cantilever (SuperSharpSilicon SSS-NCH cantilever with 2 nm tip radius). The plasma treatment leads to chemical changes at the surface by oxidizing the siloxane groups and formation of Si−O complexes. Depending on the applied conditions, the thickness of the oxide layer can be varied. The oxidized surface shows altered properties compared to the bare PDMS such as porosity and roughness (Figure S1 in Supporting Information).62 Silanization of the surface with APTES in water leads to polymerization and formation of a silane network on the oxide layer of the PDMS surface.63,64 The aminosilane-functionalized and polymerized species at the surface layer contain positively charged amine groups with a high affinity for gold.65−67 The positive charge at the surface provides the driving force for electrostatic attraction and surface deposition of the negatively charged citrate-capped AuNPs. In addition, amine groups have an affinity for gold. AFM images of samples after AuNPs deposition are presented in Figure 2. The AFM images acquired by sharper tip (which is relatively stiff) show the particles on the surface. By contrast, the acquired images by larger tip (with lower force constant and stiffness) reveal the embedding of particles and the presence of a soft covering around them. Materials with higher affinity and stronger bonding ability can replace the citrate on the surface of noble metallic colloids,42 thus we expect the particles being partially embedded into the polymerized surface layer. The embedding of the particles is promoted by stretching and releasing the PDMS templates after AuNP deposition and prior to each growth step. This effect, which is evidenced by the extinction spectra, is only visible after deposition of the seed and prior to the first growth step. Figure 3 shows the measured extinction spectrum before and after sample stretching and releasing. Embedding of particles changes the surrounding medium from air to higher refractive index materials (1.45−1.5 for aminosilane64 and PDMS oxide). This leads to a red-shift in the plasmon resonances of the AuNPs and a significant increase of extinction intensity. This finding strongly suggests that, upon

Figure 3. Extinction spectra of AuNPs deposited on the functionalized PDMS sample; pristine in black and after stretching-releasing in red. D

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 4. AFM images of grown sample (after gold nanoparticles deposition on PDMS sample and after 12 steps of growing) analyzed by larger tip cantilever (Olympus OMCL-AC240TS cantilever with a 9 ± 2 nm tip). Panels a, c, and e are topographic images, and their corresponding phase images are shown in b, d, and f.

thus they can grow more.49 Broadening of the 111 reflection in seed particles compared to other reflections is surprising. Usually an inverse trend has been observed as a result of stacking faults on the {111} planes.80 Corresponding broadening and the calculated sizes (10 nm in the ⟨111⟩ direction and 250 nm in the other directions) do not agree with TEM images. TEM images show the presence of twinned decahedral shape structures (Figure S4 in the Supporting Information), which can influence the directional growth of particles and introduce elongation.77,78 Disturbing the AuNPs orientation at the surface by touching, pressing the surface, and applying

some sheer forces in a qualitative way leads to a change of color of the samples as a result of aggregation and reorientation of particles, and causes an increase in the intensity of the 200 peak against the 111 peak. The TEM micrographs of particles removed from the surface of the PDMS sample after 12 steps of growing are presented in Figure 7, which shows elongated nanoparticles (and even more elongated single or double-pods particles were observed for a sample characterized in previous work;38 TEM micrograph presented in Figure S5 (Supporting Information)). TEM images show that the particles are not fused together. These E

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

Figure 5. X-ray diffraction patterns of seed particles at surface (without growing), grown sample (with 12 steps of growth), partly disturbed, and completely disturbed grown sample, respectively, from bottom to top. (The partly and completely disturbed samples are distinguished by the change of color. The color changed to bluish for partly disturbed sample, while we qualitatively applied some forces by touching and pressing. The color changes to dark blue for the completely disturbed sample when we applied stronger forces and sheer forces at the surface. These changes are consistent throughout the PDMS surface.). Figure 7. TEM micrographs of extracted particles from surface of grown sample (after 12 steps of growing) (scale-bar 50 nm).

Figure 6. Integrated intensity of all X-ray diffraction peaks relative to (111) peak for each sample. From left to right column, respectively: seed particles (no growth), the most grown, partly disturbed, and completely disturbed samples. Figure 8. Extinction spectra of grown sample (after 12 steps of growing).

particles have a larger diameter at one end, and some show a hollow hemispherical (concave) shape at the other end. This shape indicates the faster growth on one side. This can be due to embedding of the other side of the particles and/or a limited diffusion flux of gold ions70,81 toward one side of the growing particles. Since elongated nanoparticles (such as nanorods) should show both longitudinal and transverse plasmon resonances for random orientation, the presence of different peaks in the extinction spectra is expected. However, as shown in Figure 8, we can observe only one peak, which can be assigned to the transverse plasmon resonance. For Au rods with an aspect ratio of 2, a second plasmon resonance around 650 nm would be expected, which is clearly missing. This observation can be a result of particle orientation at the surface. The role of the oxidized and polymerized PDMS surface layer for the anisotropic growth of particles can be further examined by control experiments on glass. The glass slide was functionalized with APTES, and seed nanoparticles were deposited on top with the same procedure used for PDMS samples. The growth of the particles at the glass surface was

faster. The TEM micrograph of extracted particles from the glass surface after three growth steps is presented in Figure 9. The shape of particles grown at the glass surface is distinctly different from the PDMS samples. The particles on glass show quasi-spherical shape in contrast to the highly anisotropic shapes observed for PDMS (with double-pod S5, Supporting Information). In another control experiment, the particles were added to the growth solution. They uniformly grow under these conditions as presented in Figure 9 d. These results confirm the role of the functionalized PDMS surface for the anisotropic growth of the AuNPs. The growth of AuNPs is rather complex. Using hydroxylamine as a mild reducing agent plays an effective role in anisotropic growth of particles. In strongly basic conditions (pH 11−12.5) hydroxylamine can reduce Au (III)68 to metallic Au, which leads to the formation of nonspherical nanoparticles (nanostars and particles with long thorns along with high density of defects).42,68,82 In acidic conditions, it goes through F

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir

4. CONCLUSION The growth of elongated AuNPs at the porous surface of PDMS is reported. The plasma-treated and aminosilanefunctionalized PDMS surface has a porous structure that has been functioning as host for nanoparticles embedding. The AFM studies revealed the embedding of particles in coarse covering layer. The embedded AuNPs anisotropically grow in the presence of hydroxylamine and gold salt solution as evidenced by TEM measurements. Although the hydroxylamine has an effect on elongated growth of particles, the presence of charged aminosilane-functionalized and porous surface of PDMS has the most significant role for directional growth and elongation of AuNPs at the surface. XRPD results indicated the preferential crystalline orientation of grown particles at the surface and the average orientation perturbation through disturbing the surface. Extinction spectra indicate that the growth is preferentially perpendicular to the surface.

Figure 9. TEM micrographs of extracted AuNPs from glass as a control experiment after third (a,b) and fifth growth (c) (scale-bar 200 nm). TEM micrographs of grown AuNPs in solution in d (scale bar 50 nm).



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.langmuir.5b03524. AFM image of oxidized PDMS surface and its pore size distribution, 3D-representation of AFM image shown in Figure 4e and a line profile for this image, AFM images of a grown sample by sharper tip cantilever, TEM micrographs of colloidal AuNp, and TEM micrograph of extracted grown particles from ref 38 (PDF)

oxidation reaction only in the presence of a gold surface by a surface-catalyzed reduction.42,69,70 In this work the growth process was carried out in acidic pH, the starting pH was about 4, which decreased to pH 3 after 2 min in each growth step in the presence of sample. Changes in the samples’ color as a result of increasing the AuNP size was evidenced for each growth step (as also reported in the former work38). It has been reported that 1 s growth can lead to surface roughness of seed nanoparticles at pH 4.2.82 The branch formation has been explained by the formation of small single crystalline AuNPs and their random attachment to the particles during the first 3 s.82 The growth and ripening of particles in solution is faster and less limited compared to immobilized and embedded particles. For pH 4.2, it takes 8 s for ripening of the particles in solution, along with reshaping causing a blue-shift of their plasmon resonances.82 Shape evolution from flower-like to quasi-spherical is pH and Cl− concentration dependent.82 Rodshape particle appearance in the presence of hydroxylamine is caused by the reduced growth rate in two axial directions rather than accelerating growth rate in one direction.70 Since the growth reaction is fast, transport of reactants plays a significant role in the particle growth. The reactant transport is different for particles growing in solution, and the embedded particles inside the porous aminosilane-functionalized surface layer of PDMS. The positive charge of the surface can further reduce the diffusion of Au (III) ions to the growing particles by electrostatic repulsion. Due to the surface charge, large gradients of the local Au (III) ion concentration are expected to contribute to the anisotropic growth. The growth rate of AuNPs at the vicinity of the charged surface is rather slow since the amount of Au (III) ions is strongly depleted. Moreover, the growth behaviors of different crystalline facets are different. The most energetic crystalline facets are most likely to interact with surface NH2 polymeric groups and being passivated through stable bonding. As AFM images show, the surface polymerized aminosilane layer surrounds the particles. This effect can further limit the particle growth and reduce their interface with the growth solution and thus direct their growth outward from the surfaces.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the University of Geneva is acknowledged. We would like to thank the bioimaging center of the University of Geneva for access to electron microscopy.



ABBREVIATIONS PDMS, polydimethylsiloxane; AuNP, gold nanoparticle; AFM, atomic force microscopy; TEM, transmission electron microscope; XRPD, X-ray powder diffraction



REFERENCES

(1) Diré, S.; Babonneau, F.; Sanchez, C.; Livage, J. Sol-gel synthesis of siloxane-oxide hybrid coatings [Si(CH3)2OMO x: M = Si, Ti, Zr, Al] with luminescent properties. J. Mater. Chem. 1992, 2 (2), 239. (2) Chen, H. J.; Shao, L.; Ming, T.; Woo, K. C.; Man, Y. C.; Wang, J. F.; Lin, H. Q. Observation of the Fano Resonance in Gold Nanorods Supported on High-Dielectric-Constant Substrates. ACS Nano 2011, 5 (8), 6754−6763. (3) Yang, Y.; Matsubara, S.; Nogami, M.; Shi, J. L.; Huang, W. M. One-dimensional self-assembly of gold nanoparticles for tunable surface plasmon resonance properties. Nanotechnology 2006, 17 (11), 2821−2827. (4) Maier, S. A. Plasmonics: Metal nanostructures for subwavelength photonic devices. IEEE J. Sel. Top. Quantum Electron. 2006, 12 (6), 1214−1220. (5) Topçu Sulak, M.; Gokdogan, O.; Gulce, A.; Gulce, H. Amperometric glucose biosensor based on gold-deposited poly-

G

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir vinylferrocene film on Pt electrode. Biosens. Bioelectron. 2006, 21 (9), 1719−1726. (6) SadAbadi, H.; Badilescu, S.; Packirisamy, M.; Wuthrich, R. Rapid Microwave-Induced Synthesis of Gold-Polydimethylsiloxane Nanocomposites for Biosensing of Proteins. J. Nanosci. Nanotechnol. 2013, 13 (10), 6880−6887. (7) Homola, J. Surface plasmon resonance sensors for detection of chemical and biological species. Chem. Rev. 2008, 108 (2), 462−493. (8) Stewart, M. E.; Anderton, C. R.; Thompson, L. B.; Maria, J.; Gray, S. K.; Rogers, J. A.; Nuzzo, R. G. Nanostructured plasmonic sensors. Chem. Rev. 2008, 108 (2), 494−521. (9) Dreaden, E. C.; Alkilany, A. M.; Huang, X. H.; Murphy, C. J.; ElSayed, M. A. The golden age: gold nanoparticles for biomedicine. Chem. Soc. Rev. 2012, 41 (7), 2740−2779. (10) Kang, H.; Heo, C. J.; Jeon, H. C.; Lee, S. Y.; Yang, S. M. Durable Plasmonic Cap Arrays on Flexible Substrate with Real-Time Optical Tunability for High-Fidelity SERS Devices. ACS Appl. Mater. Interfaces 2013, 5 (11), 4569−4574. (11) Schuller, J. A.; Barnard, E. S.; Cai, W. S.; Jun, Y. C.; White, J. S.; Brongersma, M. L. Plasmonics for extreme light concentration and manipulation. Nat. Mater. 2010, 9 (3), 193−204. (12) Mulvihill, M. J.; Ling, X. Y.; Henzie, J.; Yang, P. D. Anisotropic Etching of Silver Nanoparticles for Plasmonic Structures Capable of Single-Particle SERS. J. Am. Chem. Soc. 2010, 132 (1), 268−274. (13) Lee, K. S.; El-Sayed, M. A. Gold and silver nanoparticles in sensing and imaging: Sensitivity of plasmon response to size, shape, and metal composition. J. Phys. Chem. B 2006, 110 (39), 19220− 19225. (14) Khlebtsov, N. G.; Dykman, L. A. Optical properties and biomedical applications of plasmonic nanoparticles. J. Quant. Spectrosc. Radiat. Transfer 2010, 111 (1), 1−35. (15) Myroshnychenko, V.; Rodriguez-Fernandez, J.; Pastoriza-Santos, I.; Funston, A. M.; Novo, C.; Mulvaney, P.; Liz-Marzan, L. M.; García de Abajo, F. J. Modelling the optical response of gold nanoparticles. Chem. Soc. Rev. 2008, 37 (9), 1792−1805. (16) Grzelczak, M.; Perez-Juste, J.; Mulvaney, P.; Liz-Marzan, L. M. Shape control in gold nanoparticle synthesis. Chem. Soc. Rev. 2008, 37 (9), 1783−1791. (17) Ye, X. C.; Chen, J.; Diroll, B. T.; Murray, C. B. Tunable Plasmonic Coupling in Self-Assembled Binary Nanocrystal Super lattices Studied by Correlated Optical Microspectrophotometry and Electron Microscopy. Nano Lett. 2013, 13 (3), 1291−1297. (18) Schreiber, R.; Luong, N.; Fan, Z. Y.; Kuzyk, A.; Nickels, P. C.; Zhang, T.; Smith, D. M.; Yurke, B.; Kuang, W.; Govorov, A. O.; Liedl, T. Chiral plasmonic DNA nanostructures with switchable circular dichroism. Nat. Commun. 2013, 4, 2948. (19) Jana, N. R.; Gearheart, L.; Murphy, C. J. Wet chemical synthesis of high aspect ratio cylindrical gold nanorods. J. Phys. Chem. B 2001, 105 (19), 4065−4067. (20) Mieszawska, A. J.; Slawinski, G. W.; Zamborini, F. P. Directing the growth of highly aligned gold nanorods through a surface chemical amidation reaction. J. Am. Chem. Soc. 2006, 128 (17), 5622−5623. (21) Jana, N. R. Gram-scale synthesis of soluble, near-monodisperse gold nanorods and other anisotropic nanoparticles. Small 2005, 1 (8− 9), 875−882. (22) Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J. X.; Gou, L.; Hunyadi, S. E.; Li, T. Anisotropic metal nanoparticles: Synthesis, assembly, and optical applications. J. Phys. Chem. B 2005, 109 (29), 13857−13870. (23) Jana, N. R.; Gearheart, L.; Obare, S. O.; Murphy, C. J. Anisotropic chemical reactivity of gold spheroids and nanorods. Langmuir 2002, 18 (3), 922−927. (24) Huang, X. H.; Neretina, S.; El-Sayed, M. A. Gold Nanorods: From Synthesis and Properties to Biological and Biomedical Applications. Adv. Mater. 2009, 21 (48), 4880−4910. (25) Perez-Juste, J.; Pastoriza-Santos, I.; Liz-Marzan, L. M.; Mulvaney, P. Gold nanorods: Synthesis, characterization and applications. Coord. Chem. Rev. 2005, 249 (17−18), 1870−1901.

(26) Kelly, K. L.; Coronado, E.; Zhao, L. L.; Schatz, G. C. The optical properties of metal nanoparticles: The influence of size, shape, and dielectric environment. J. Phys. Chem. B 2003, 107 (3), 668−677. (27) Sosa, I. O.; Noguez, C.; Barrera, R. G. Optical properties of metal nanoparticles with arbitrary shapes. J. Phys. Chem. B 2003, 107 (26), 6269−6275. (28) Niu, W. X.; Zhang, L.; Xu, G. B. Seed-mediated growth of noble metal nanocrystals: crystal growth and shape control. Nanoscale 2013, 5 (8), 3172−3181. (29) Jones, M. R.; Osberg, K. D.; Macfarlane, R. J.; Langille, M. R.; Mirkin, C. A. Templated Techniques for the Synthesis and Assembly of Plasmonic Nanostructures. Chem. Rev. 2011, 111 (6), 3736−3827. (30) Huczko, A. Template-based synthesis of nanomaterials. Appl. Phys. A: Mater. Sci. Process. 2000, 70 (4), 365−376. (31) Sander, M. S.; Tan, L. S. Nanoparticle arrays on surfaces fabricated using anodic alumina films as templates. Adv. Funct. Mater. 2003, 13 (5), 393−397. (32) Hoyer, P. Multistep replication processes. Curr. Opin. Colloid Interface Sci. 1998, 3 (2), 160−165. (33) Shin, K.; Leach, K. A.; Goldbach, J. T.; Kim, D. H.; Jho, J. Y.; Tuominen, M.; Hawker, C. J.; Russell, T. P. A simple route to metal nanodots and nanoporous metal films. Nano Lett. 2002, 2 (9), 933− 936. (34) Angelome, P. C.; Pastoriza-Santos, I.; Perez-Juste, J.; RodriguezGonzalez, B.; Zelcer, A.; Soler-Illia, G. J. A. A.; Liz-Marzan, L. M. Growth and branching of gold nanoparticles through mesoporous silica thin films. Nanoscale 2012, 4 (3), 931−939. (35) Muller, M. B.; Kuttner, C.; Konig, T. A. F.; Tsukruk, V. V.; Forster, S.; Karg, M.; Fery, A. Plasmonic Library Based on SubstrateSupported Gradiential Plasmonic Arrays. ACS Nano 2014, 8 (9), 9410−9421. (36) Jin, Y. D.; Kang, X. F.; Song, Y. H.; Zhang, B. L.; Cheng, G. J.; Dong, S. J. Controlled nucleation and growth of surface-confined gold nanoparticles on a (3-aminopropyl)trimethoxysilane-modified class slide: A strategy for SPR substrates. Anal. Chem. 2001, 73 (13), 2843− 2849. (37) Taub, N.; Krichevski, O.; Markovich, G. Growth of gold nanorods on surfaces. J. Phys. Chem. B 2003, 107 (42), 11579−11582. (38) Cataldi, U.; Caputo, R.; Kurylyak, Y.; Klein, G.; Chekini, M.; Umeton, C.; Burgi, T. Growing gold nanoparticles on a flexible substrate to enable simple mechanical control of their plasmonic coupling. J. Mater. Chem. C 2014, 2 (37), 7927−7933. (39) Chekini, M.; Oulevey, P.; Bürgi, T. Large-scale plasmonic-SERS templates by successive growth steps. Curr. Appl. Phys. 2015, 15 (3), 253−260. (40) Musick, M. D.; Pena, D. J.; Botsko, S. L.; McEvoy, T. M.; Richardson, J. N.; Natan, M. J. Electrochemical properties of colloidal Au-based surfaces: Multilayer assemblies and seeded colloid films. Langmuir 1999, 15 (3), 844−850. (41) Huang, X.; Li, H.; Li, S. Z.; Wu, S. X.; Boey, F.; Ma, J.; Zhang, H. Synthesis of Gold Square-like Plates from Ultrathin Gold Square Sheets: The Evolution of Structure Phase and Shape. Angew. Chem., Int. Ed. 2011, 50 (51), 12245−12248. (42) Yuan, H.; Ma, W. H.; Chen, C. C.; Zhao, J. C.; Liu, J. W.; Zhu, H. Y.; Gao, X. P. Shape and SPR evolution of thorny gold nanoparticles promoted by silver ions. Chem. Mater. 2007, 19 (7), 1592−1600. (43) Genc, R.; Clergeaud, G.; Ortiz, M.; O'Sullivan, C. Shape directed biomineralization of gold nanoparticles using self-assembled lipid structures. Biomater. Sci. 2014, 2 (8), 1128−1134. (44) Fan, F. R.; Liu, D. Y.; Wu, Y. F.; Duan, S.; Xie, Z. X.; Jiang, Z. Y.; Tian, Z. Q. Epitaxial growth of heterogeneous metal nanocrystals: From gold nano-octahedra to palladium and silver nanocubes. J. Am. Chem. Soc. 2008, 130 (22), 6949−6951. (45) Chen, H. M.; Liu, R. S.; Tsai, D. P. A Versatile Route to the Controlled Synthesis of Gold Nanostructures. Cryst. Growth Des. 2009, 9 (5), 2079−2087. (46) Jana, N. R.; Gearheart, L.; Murphy, C. J. Seed-mediated growth approach for shape-controlled synthesis of spheroidal and rod-like gold H

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX

Article

Langmuir nanoparticles using a surfactant template. Adv. Mater. 2001, 13 (18), 1389−1393. (47) Cao, L. Y.; Zhu, T.; Liu, Z. F. Formation mechanism of nonspherical gold nanoparticles during seeding growth: Roles of anion adsorption and reduction rate. J. Colloid Interface Sci. 2006, 293 (1), 69−76. (48) Sau, T. K.; Murphy, C. J. Room temperature, high-yield synthesis of multiple shapes of gold nanoparticles in aqueous solution. J. Am. Chem. Soc. 2004, 126 (28), 8648−8649. (49) Bakshi, M. S.; Sachar, S.; Kaur, G.; Bhandari, P.; Kaur, G.; Biesinger, M. C.; Possmayer, F.; Petersen, N. O. Dependence of crystal growth of gold nanoparticles on the capping behavior of surfactant at ambient conditions. Cryst. Growth Des. 2008, 8 (5), 1713−1719. (50) Topete, A.; Alatorre-Meda, M.; Villar-Alvarez, E. M.; Cambon, A.; Barbosa, S.; Taboada, P.; Mosquera, V. Simple Control of Surface Topography of Gold Nanoshells by a Surfactant-less Seeded-Growth Method. ACS Appl. Mater. Interfaces 2014, 6 (14), 11142−11157. (51) De Luca, A.; Depalo, N.; Fanizza, E.; Striccoli, M.; Curri, M. L.; Infusino, M.; Rashed, A. R.; La Deda, M.; Strangi, G. Plasmon mediated super-absorber flexible nanocomposites for metamaterials. Nanoscale 2013, 5 (13), 6097−6105. (52) Yu, D. L.; Huang, H. T.; Lu, L. F.; Che, J. F.; Chen, X. Y.; Zhu, X. F.; Song, Y.; Li, D. D. Templated deposition of multiscale periodic metallic nanodot arrays with sub-10nm gaps on rigid and flexible substrates. Nanotechnology 2014, 25 (46), 465303. (53) Maurer, T.; Marae-Djouda, J.; Cataldi, U.; Gontier, A.; Montay, G.; Madi, Y.; Panicaud, B.; Macias, D.; Adam, P.-M.; Lévêque, G.; Bürgi, T.; Caputo, R. The beginnings of plasmomechanics: towards plasmonic strain sensors. Frontiers of Materials Science 2015, 9 (2), 170−177. (54) Liu, J.; Lu, C. Y.; Zhou, H.; Xu, J. J.; Chen, H. Y. Flexible Gold Electrode Array for Multiplexed Immunoelectrochemical Measurement of Three Protein Biomarkers for Prostate Cancer. ACS Appl. Mater. Interfaces 2014, 6 (22), 20137−20143. (55) Chung, A. J.; Huh, Y. S.; Erickson, D. Large area flexible SERS active substrates using engineered nanostructures. Nanoscale 2011, 3 (7), 2903−2908. (56) Seo, D.; Park, J. C.; Song, H. Polyhedral gold nanocrystals with O-h symmetry: From octahedra to cubes. J. Am. Chem. Soc. 2006, 128 (46), 14863−14870. (57) Schulz, F.; Homolka, T.; Bastus, N. G.; Puntes, V.; Weller, H.; Vossmeyer, T. Little Adjustments Significantly Improve the Turkevich Synthesis of Gold Nanoparticles. Langmuir 2014, 30 (35), 10779− 10784. (58) Ng, J. M. K.; Gitlin, I.; Stroock, A. D.; Whitesides, G. M. Components for integrated poly(dimethylsiloxane) microfluidic systems. Electrophoresis 2002, 23 (20), 3461−3473. (59) McDonald, J. C.; Whitesides, G. M. Poly(dimethylsiloxane) as a material for fabricating microfluidic devices. Acc. Chem. Res. 2002, 35 (7), 491−499. (60) Lee, J. N.; Park, C.; Whitesides, G. M. Solvent compatibility of poly(dimethylsiloxane)-based microfluidic devices. Anal. Chem. 2003, 75 (23), 6544−6554. (61) Goyal, A.; Kumar, A.; Patra, P. K.; Mahendra, S.; Tabatabaei, S.; Alvarez, P. J. J.; John, G.; Ajayan, P. M. In situ Synthesis of Metal Nanoparticle Embedded Free Standing Multifunctional PDMS Films. Macromol. Rapid Commun. 2009, 30 (13), 1116−1122. (62) de Menezes Atayde, C.; Doi, I. Highly stable hydrophilic surfaces of PDMS thin layer obtained by UV radiation and oxygen plasma treatments. Phys. Status Solidi C 2010, 7 (2), 189−192. (63) Nakagawa, T.; Tanaka, T.; Niwa, D.; Osaka, T.; Takeyama, H.; Matsunaga, T. Fabrication of amino silane-coated microchip for DNA extraction from whole blood. J. Biotechnol. 2005, 116 (2), 105−111. (64) Vandenberg, E. T.; Bertilsson, L.; Liedberg, B.; Uvdal, K.; Erlandsson, R.; Elwing, H.; Lundstrom, I. Structure of 3-Aminopropyl Triethoxy Silane on Silicon-Oxide. J. Colloid Interface Sci. 1991, 147 (1), 103−118. (65) Wangoo, N.; Bhasin, K. K.; Mehta, S. K.; Suri, C. R. Synthesis and capping of water-dispersed gold nanoparticles by an amino acid:

Bioconjugation and binding studies. J. Colloid Interface Sci. 2008, 323 (2), 247−254. (66) Tiwari, P. M.; Vig, K.; Dennis, V. A.; Singh, S. R. Functionalized Gold Nanoparticles and Their Biomedical Applications. Nanomaterials 2011, 1 (1), 31−63. (67) Ertorer, E.; Avery, J. C.; Pavelka, L. C.; Mittler, S. Surfaceimmobilized Gold Nanoparticles by Organometallic CVD on Amineterminated Glass Surfaces. Chem. Vap. Deposition 2013, 19 (10−12), 338−346. (68) Minati, L.; Benetti, F.; Chiappini, A.; Speranza, G. One-step synthesis of star-shaped gold nanoparticles. Colloids Surf., A 2014, 441, 623−628. (69) Lyon, J. L.; Fleming, D. A.; Stone, M. B.; Schiffer, P.; Williams, M. E. Synthesis of Fe oxide core/Au shell nanoparticles by iterative hydroxylamine seeding. Nano Lett. 2004, 4 (4), 719−723. (70) Brown, K. R.; Walter, D. G.; Natan, M. J. Seeding of colloidal Au nanoparticle solutions. 2. Improved control of particle size and shape. Chem. Mater. 2000, 12 (2), 306−313. (71) Brown, K. R.; Natan, M. J. Hydroxylamine seeding of colloidal Au nanoparticles in solution and on surfaces. Langmuir 1998, 14 (4), 726−728. (72) Brown, K. R.; Lyon, L. A.; Fox, A. P.; Reiss, B. D.; Natan, M. J. Hydroxylamine seeding of colloidal au nanoparticles. 3. Controlled formation of conductive Au films. Chem. Mater. 2000, 12 (2), 314− 323. (73) Jeong, G. H.; Lee, Y. W.; Kim, M.; Han, S. W. High-yield synthesis of multi-branched gold nanoparticles and their surfaceenhanced Raman scattering properties. J. Colloid Interface Sci. 2009, 329 (1), 97−102. (74) Philip, D. Green synthesis of gold and silver nanoparticles using Hibiscus rosa sinensis. Phys. E 2010, 42 (5), 1417−1424. (75) Philip, D. Biosynthesis of Au, Ag and Au-Ag nanoparticles using edible mushroom extract. Spectrochim. Acta, Part A 2009, 73 (2), 374− 381. (76) Liu, M. Z.; Guyot-Sionnest, P. Mechanism of silver(I)-assisted growth of gold nanorods and bipyramids. J. Phys. Chem. B 2005, 109 (47), 22192−22200. (77) Lofton, C.; Sigmund, W. Mechanisms controlling crystal habits of gold and silver colloids. Adv. Funct. Mater. 2005, 15 (7), 1197− 1208. (78) Sun, Y. G.; Ren, Y.; Liu, Y. Z.; Wen, J. G.; Okasinski, J. S.; Miller, D. J. Ambient-stable tetragonal phase in silver nanostructures. Nat. Commun. 2012, 3, 971. (79) Kannan, P.; Abraham John, S. Synthesis of mercaptothiadiazolefunctionalized gold nanoparticles and their self-assembly on Au substrates. Nanotechnology 2008, 19 (8), 085602. (80) Gubicza, J.; Labar, J. L.; Quynh, L. M.; Nam, N. H.; Luong, N. H. Evolution of size and shape of gold nanoparticles during long-time aging. Mater. Chem. Phys. 2013, 138 (2−3), 449−453. (81) Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, 1980; p 718. (82) Zhao, L. L.; Ji, X. H.; Sun, X. J.; Li, J.; Yang, W. S.; Peng, X. G. Formation and Stability of Gold Nanoflowers by the Seeding Approach: The Effect of Intraparticle Ripening. J. Phys. Chem. C 2009, 113 (38), 16645−16651.

I

DOI: 10.1021/acs.langmuir.5b03524 Langmuir XXXX, XXX, XXX−XXX