15
A ngewandte Chemi e Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination Number and Extraordinary Enhancement in Surface- Enhanced Raman Scattering (SERS)** Nicolas Pazos-Perez, Claudia Simone Wagner, Jose M. Romo-Herrera, Luis M. Liz-MarzƁn, F Javier Garcȷa de Abajo,* Alexander Wittemann,* Andreas Fery,* and RamɃn A. Alvarez-Puebla* . Angewandte Communications 12688 # 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 12688 –12693

Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

AngewandteChemie

Nanoparticle ClustersDOI: 10.1002/anie.201207019

Organized Plasmonic Clusters with High CoordinationNumber and Extraordinary Enhancement in Surface-Enhanced Raman Scattering (SERS)**Nicolas Pazos-Perez, Claudia Simone Wagner, Jose M. Romo-Herrera,Luis M. Liz-Marz�n, F Javier Garc�a de Abajo,* Alexander Wittemann,*Andreas Fery,* and Ram�n A. Alvarez-Puebla*

.AngewandteCommunications

12688 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 12688 –12693

Page 2: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Noble metal nanoparticles exhibit optical excitations knownas surface plasmons that produce large enhancement of thelocal light intensity under external illumination, particularlywhen the nanoparticles are arranged in closely spacedconfigurations.[1] The interparticle gap distance[2] plays a crit-ical role in the generation of hotspots with high electro-magnetic fields, and thus such assembled nanoparticles findapplication to ultrasensitive detection, for example throughsurface-enhanced Raman scattering[3] (SERS) and nonlinearoptics, among other feats.[4] Controlled assembly usingcolloidal chemistry is an emerging and promising field forhigh-yield production of metal nanoparticle clusters withsmall interparticle gaps.[5] However, most of the reportedmethods rely on the use of nucleic acids or other organicmolecules as linking elements,[6] which yield long separationdistances and thus weak plasmon coupling. Additionally, onlysimple clusters, such as dimers and trimers, have beenefficiently synthesized. Herein, we report the controlledassembly of gold nanospheres into well-defined nanoparticleclusters with large coordination numbers (up to 7) and highsymmetry. We further demonstrate ultrasensitive direct andindirect SERS sensing, thus corroborating the outstandingoptical performance of these clusters with robust enhance-ment factors that are over three orders of magnitude higherthan those of single particles.

The optical response of plasmonic particles upon lightirradiation is mainly determined by particle size and shape, aswell as the dielectric environment and geometrical arrange-ment.[2b, 7] Recent advances in wet-chemical synthesis enablethe fabrication of nanostructures with engineered complexsize, shape, and composition.[8] However, the colloidal syn-thesis of organized associations of these particles to fullyexploit their plasmonic couplings remains a challenge, despiteincreasing efforts to produce monodisperse stable colloidalclusters. Pioneering work[9, 10] demonstrated the possibility ofobtaining nanoparticle dimers and trimers by using molecularlinkers such as thiolated nucleic acids or other long organicmolecules. This molecular linking approach is still frequently

used, even though the final product is a mixture of clusterswith different numbers of particles. Actually, numerousvariations of this method exist for the production ofhomogeneous samples containing dimers and other geome-tries with low coordination numbers;[6b, 11] with this approacha purification process is always required, typically includingsedimentation,[12] electrophoresis,[13] and/or density gradientcentrifugation.[14] However, the application of these clustersto plasmonics in general, and to SERS in particular, is stilllimited because of three severe constraints. First, the use ofthiolated linkers chemically passivates the plasmonic surface,thereby inhibiting the adsorption of other target species;second, the linker itself produces a SERS signal. These issuesare crucial, considering that the thiolated molecule is placedprecisely at the hotspot. And finally, perhaps the mostimportant limitation is the size of the molecule: linkersbased on nucleic acids require a minimum size of approx-imately 10 nm, leading to interparticle distances that are toolong to produce effective hotspots. To avoid this drawback,a novel approach based on the linkage and purification ofdimers and subsequent in situ overgrowth was recentlyreported.[15] Despite the improvement, this method is stillsubject to vibrational contamination from the DNA linkers,and it has been only demonstrated for dimers.

Other approaches for the generation of dimers in colloidalsolution include the use of asymmetric functionalization orpolymer ligands. In the former, particles are functionalizedwith different capping agents, which can chemically bind toeach other through simple reactions. The most popular choiceis the functionalization of separate nanoparticle batches withamino and carboxylic groups. After mixing, carbodiimidechemistry serves to bind the two different fragments, therebyleading to formation of particle clusters.[16] This approach,however, is affected by the same problems as those observedwhen using DNA or organic linkers. In contrast, polymerligands exploit electrostatic interactions between a polyelec-trolyte and the particles. This method has proven effective toform nanoparticle dimers with poly(vinyl pyrrolidone),

[*] Dr. N. Pazos-Perez,[+] Prof. A. FeryPhysikalische Chemie II, Universit�t BayreuthUniversit�tsstrasse 30, 95440 Bayreuth (Germany)E-mail: [email protected]

Dr. C. S. Wagner,[+] Prof. A. WittemannColloid chemistry, University of KonstanzUniversit�tsstrasse 10, 78464 Konstanz (Germany)E-mail: [email protected]

Dr. J. M. Romo-HerreraCentro de Nanociencias y Nanotecnolog�a, UNAM, km 107Carretera Tijuana-Ensenada, C.P. 22860, Ensenada B.C. (M�xico)

Dr. J. M. Romo-Herrera, Prof. L. M. Liz-Marz�nDepartamento de Qu�mica-F�sica, Universidade de Vigo36310 Vigo (Spain)

Prof. L. M. Liz-Marz�nIkerbasque, Basque Foundation for Science, 48011 Bilbao (Spain)and Centre for Cooperative Research in Biomaterials (CIC bioma-GUNE), Paseo de Miram�n 182, 20009 San Sebasti�n (Spain)

Prof. F. J. Garc�a de AbajoInstituto de Quimica-Fisica Rocasolano-CSICSerrano 119, 28006, Madrid (Spain)E-mail: [email protected]

Prof. R. A. Alvarez-PueblaICREA (Catalan Institution for Research and Advanced Studies)Passeig Llu�s Companys 23, 08010 Barcelona (Spain) andDepartment of Electronic Engineering & Center for ChemicalTechnology of Catalonia, Universitat Rovira i VirgiliAvda. Pa�sos Catalans 26, 43007 Tarragona (Spain)E-mail: [email protected]

[+] These authors contributed equally to this work.

[**] This work was funded by the Spanish Ministerio de Economiay Competitividad (CTQ2011-23167, MAT2010-15374, MAT2010-14885, and Consolider NanoLight.es), the Xunta de Galicia(09TMT011314PR), and the European Research Council (AdvancedGrant PLASMAQUO; Starting Grant METAMECH; and FP7/2008Metachem Project under grant agreement no. 228762); Germanresearch foundation (DFG) within collaborative research center 840,TPs A3 and B5.

Supporting information for this article is available on the WWWunder http://dx.doi.org/10.1002/anie.201207019.

Re-use of this article is permitted in accordance with the Terms andConditions set out at http://angewandte.org/open.

AngewandteChemie

12689Angew. Chem. Int. Ed. 2012, 51, 12688 –12693 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Page 3: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

poly(ethylene glycol), and poly(methyl methacrylate).[17] Theuse of polymers instead of DNA or organic linkers not onlyallows efficiently gluing colloidal particles with smallerinterparticle gaps,[18] but also concentrating the analytes[19]

within close proximity to the plasmonic surface while avoid-ing vibrational signal contamination, because the polymersdisplay low Raman cross-sections.[20]

We combine here these two properties by resorting toblock copolymers based on poly(ethylene oxide)-b-poly-(propylene oxide)-b-poly(ethylene oxide) (Pluronic F-68,PF68) to produce highly symmetric gold nanoparticle clusterswith coordination numbers (CNs) up to 7. The resultingclusters are efficiently separated by density gradient centri-fugation into homogeneous aliquots of dimers (CN = 2),trimers (CN = 3), tetramers (CN = 4, mostly forming tetrahe-dra), and a complex mixture of trigonal bipyramids; octahe-dra; and pentagonal bipyramids (CN = 5, 6, and 7). Theoptical properties of the resulting materials are extensivelycharacterized and theoretically modeled to explain theirunique optical performance for electromagnetic field en-hancement.

The assembly of the colloidal nanoparticles into clusterscomprises initially the fabrication of monodisperse sphericalgold nanoparticles (NPs) of average diameter of approxi-mately 50 nm by a seed-mediated approach.[21] After prepa-ration, the NPs capped with cetyl trimethylammoniumbromide (CTAB) show a zeta potential of + 48 mV. TheNPs are then concentrated and transferred into an aqueoussolution of PF68, which leads to a reversal of the zetapotential to �19 mV. The colloidal dispersion of gold NPscoated with PF68 is then emulsified with toluene by using anultrasonic homogenizer. This procedure yields narrowlydispersed emulsion droplets bearing nanoparticles at theirsurface.[22] Packing of the NPs into clusters is induced bygradual toluene evaporation (in a rotary evaporator). Theisotropy of the starting particles enables the construction ofhighly symmetric clusters. Finally, the separation of thedifferent populations of clusters is achieved by densitygradient centrifugation. Separation is then induced by thespecific sedimentation velocity of each different cluster undercentrifugation, giving rise to the typical stripes shown inFigure 1a. The different fractions of clusters can then becollected using a piercing unit and separated again to ensurepurity. All these processes lead to the complete separation ofclusters with coordination numbers of 2, 3, and 4 (Figure 1b).Owing to the close density values for higher CNs (5–7),gradient centrifugation does not have enough resolution andthese clusters are obtained as a mixture within the samestripe.

For the characterization of the interparticle distances,which is a critical parameter for the efficient optical couplingof plasmon modes, an extensive collection of dimers andtrimers was investigated by HR-TEM. Figure 1c showsrepresentative gaps between two particles, with valuesaround 2 nm. Also, an average side length of 14–15 nm wasmeasured for the triangular cavity formed by the assembly ofthree particles within trimers; this length is in good agreementwith the geometrical prediction for 50 nm particles with aninterparticle separation of 2 nm. Interestingly, these interpar-

ticle distances can be tuned either by decreasing the amountof PF68 (which leads to a decrease in the interparticle gap) orby decreasing the size of the nanoparticles (which decreasesthe cavity size). However, if the final aim of the cluster isoptical sensing, the size of both gaps and particles should beengineered to maximize the optical enhancement whileallowing the analyte to diffuse inside the cluster. With suchconsiderations in mind, clusters with 2 nm gaps combine highoptical coupling, generated by the strong electromagneticfields from 50 nm particles,[23] and sufficient space fordiffusion of small- and medium-size analytes within the gaps.

The size of the particles is also appropriate to carry outoptical and SERS characterization at the single-cluster level.To this end, dilute solutions of each sample were spin-coatedonto a marked glass window. The slide surfaces were imagedwith SEM to localize isolated clusters with different coordi-nation numbers. Then, the areas containing the clusters werestudied by dark-field optical microspectroscopy[24] and SERSafter retention of benzenethiol from the gas phase (Figure 2).This experiment allowed us to record the scattering spectrumand the optical enhancement from the same clusters.

Figure 3 shows measured and calculated[25] (gold spheresof 50 nm in diameter were considered, with the metaldescribed through its measured dielectric function[26] andincluding a 0.6 nm coating of refractive index 1.3 to effectivelyrepresent the linking copolymer layer) Vis-NIR spectra ofclusters with CNs ranging from 1 to 7. The agreement for lowCNs is reasonably good, thus corroborating the opticalbehavior expected from plasmon hybridization and interac-tion between different gaps in the corresponding particlearrangements (see below). However, the comparison for CNsabove 4 is less satisfactory, presumably owing to structuraldeformations of the clusters in the immersion oil (required tomatch the dielectric constants between the objective and thesample), thereby leading to more elongated arrangementsthat result in a strong redshift of the spectral weight. The

Figure 1. Preparation of gold nanoclusters with high coordinationnumbers. a) Density gradient separation to obtain a specific popula-tion from the initial mixture of clusters. b) SEM micrographs of thedifferent cluster populations obtained after careful fractionation of thecorresponding stripes in image (a) by using a piercing unit. c) Repre-sentative HR-TEM images of particle dimers and trimers illustratingthe small interparticle gaps and cavities generated by this fabricationmethod.

.AngewandteCommunications

12690 www.angewandte.org � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 12688 –12693

Page 4: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

spectra evolve towards more complex structures as the CNincreases. Isolated nanospheres are dominated by a singledipolar-plasmon band at 543 nm, while the optical response ofdimers is reminiscent of elongated anisotropic particles,characterized by two bands attributed to transversal andlongitudinal plasmonic contributions. The latter is pushedtowards the infrared region as a result of strong interparticleinteraction through the gap, whereas the former is due to non-interacting transverse dipole plasmons that remain unaffectedat the constituent individual particles. Trimers and higher-CNclusters exhibit more complex spectra, in which the short-wavelength transverse-dipole plasmon and the long-wave-length gap-mediated plasmon are modulated by the presence

of additional spheres and the interaction among differentgaps. While the dimer maintains a strong polarizationdependence of its optical spectrum, a fully symmetricaltrimer (D3h) has, by symmetry, a polarization-isotropicextinction spectrum. This results in the two experimentallyobserved bands, with a resolvable distortion revealed bya shoulder in the experiment; we attribute this distortion tosymmetry breaking from slight differences in the shape of thegold particles. Theoretical predictions for the tetrahedralclusters (Td) show two main contributions, the electric-dipoleresonance at 580 nm and the magnetic-dipole resonancearound 800 nm.[27] Analogously to the trimer, both resonancesare distorted by symmetry breaking in the experimentalspectra. Higher coordination numbers (5 (D3h symmetry), 6(Oh), and 7 (D5h)) present complex calculated spectra thatreflect an intricate interplay between electric and magneticmultipoles in the different spherical particles, conformed bythe symmetry of each cluster configuration, although theexperimental spectra are strongly affected by the immersionoil, as noted above.

While much work has been invested into the optimizationof dimer geometries to produce large electric field enhance-ments, as well as the resulting giant enhancement of SERSsignals,[15,17c,28] more complex geometries have not beenthoroughly analyzed. As we show below, our results demon-strate that such higher coordination clusters can be used asefficient SERS platforms. To that end, we allowed adsorptionof benzenethiol, a well-known SERS probe, from the gasphase[3] onto the isolated clusters, previously localized on themarked slide. High-resolution confocal SERS was then usedto study the clusters at the same positions where the dark-field microspectroscopy images had been acquired. Resultsfor the measured optical enhancement of each structure areshown in Figure 4, normalized to the signal observed fora number of individual particles equal to the CN (enhance-ment = 1). All clusters exhibit the strong SERS signalscharacteristic of benzenethiol (C-H bending at 1022 cm�1

and ring breathings at 1073 and 999 cm�1). Notably, individualsingle particles yielded no signal, and therefore their SERSspectrum was obtained from samples containing 27 well-separated particles within the sampled area. As expected, theSERS signals of a dimer and a trimer show an intensityincrease around 100-and 300-fold, respectively. These en-hancement factors (EFs) with respect to non-interactingindividual particles, are well-understood and ascribed to thegeneration of optical hotspots within the gaps formedbetween the particles.[2b, 29] For tetrahedral clusters, theSERS enhancement dramatically increases up to the levelof three orders of magnitude. A similar abrupt step isobserved for the trigonal bipyramid and the octahedra,while the pentagonal bipyramid reaches EFs close to 5000-fold. It should be stressed that the long-wavelength plasmonfeatures observed in Figure 3 are displaced to the blue whenthe clusters are placed in air (see Figure S1 in the SupportingInformation), so that the current illumination conditions (at633 and 785 nm) are off-resonant. Furthermore, the near-fielddistributions show large field enhancement at the gapsbetween the particles (Figure 4b). This finding is consistentwith a tight-binding description of the optical response of

Figure 2. Schematic procedure for the optical and SERS character-ization of single clusters. a) SEM and optical images of a marked glasssurface spin-coated with one of the cluster samples. b) The cluster islocalized and imaged with SEM. c, d) The region localized with SEM isimaged with Raman and dark-field spectroscopies.

Figure 3. Optical response of nanoparticle clusters with coordinationnumbers from 1 to 7. a) Dark-field single-particle optical spectroscopyof the clusters. Next to it, SEM pictures of the corresponding clustersare shown. b) Electromagnetic simulations of the same clusters.

AngewandteChemie

12691Angew. Chem. Int. Ed. 2012, 51, 12688 –12693 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Page 5: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

these structures (see the Supporting Information), thusshowing that the enhancement is associated with the gaps,and thus, other regions in the clusters are rather inactive forSERS. Therefore, as a first approximation, we interpret ourresults as receiving a contribution to the EF proportional tothe number of gaps G in each cluster:

EF ¼ CNþA G: ð1Þ

Notice that G increases strongly with the CN (e.g., G = 1,3, 6, 9, 12, and 15 for CN = 2–7, respectively). The constantprefactor A equals the EF of the dimer relative to theindividual particle. We show in Figure 4c a fit of our data tothis simple expression, which yields A = 288.1� 8.83 under633 nm excitation and A = 98.6� 9.46 under 785 nm excita-tion. The smaller enhancement in the latter is expected,because the light wavelength is more displaced towards thered with respect to the airborne cluster plasmon bands (seethe Supporting Information). Incidentally, the SERS EFsexpected for these energies from the surface-integratedintensity enhancement calculated from MESME[30] at theexcitation and emitted wavelengths are 2092 and 297,respectively. The smaller values observed in the experimentcan be attributed to a combination of nonlocal effects[31] andnonuniform coverage of the gold surface by the analyte, sothat the latter is not always present at the gap center, wherethe field enhancement is maximum.

In summary, we have shown that by using PF68 coatingand emulsion clustering it is possible to produce plasmonicnanoparticle molecules with high symmetry and coordinationindex, and that they can be separated by applying densitygradient centrifugation. PF68 produces narrow interparticlegaps with subsequent strong optical interactions while allow-ing the analytes to diffuse inside the gaps, where giganticelectric fields are generated, as we have shown by directlymeasuring the SERS enhancement in the clusters. Ourgeometrical nanostructures not only open a new path forthe investigation of optical interactions between nanoparti-cles, but they also have great potential for applications tosensing and nonlinear nanophotonics.

Received: August 30, 2012Published online: November 4, 2012

.Keywords: enhancement factors · hotspots · nanoparticles ·plasmons · surface-enhanced Raman scattering

[1] a) H. Xu, E. J. Bjerneld, M. Kall, L. Borjesson, Phys. Rev. Lett.1999, 83, 4357 – 4360; b) K. Li, M. I. Stockman, D. J. Bergman,Phys. Rev. Lett. 2003, 91, 227402.

[2] a) F. J. Garc�a de Abajo, Rev. Mod. Phys. 2010, 82, 209 – 275;b) N. J. Halas, S. Lal, W.-S. Chang, S. Link, P. Nordlander, Chem.Rev. 2011, 111, 3913 – 3961.

[3] R. A. Alvarez-Puebla, A. Agarwal, P. Manna, B. P. Khanal, P.Aldeanueva-Potel, E. Carb�-Argibay, N. Pazos-P�rez, L. Vig-derman, E. R. Zubarev, N. A. Kotov, L. M. Liz-Marz�n, Proc.Natl. Acad. Sci. USA 2011, 108, 8157 – 8161.

[4] M. Danckwerts, L. Novotny, Phys. Rev. Lett. 2007, 98, 026104.[5] J. M. Romo-Herrera, R. A. Alvarez-Puebla, L. M. Liz-Marzan,

Nanoscale 2011, 3, 1304 – 1315.[6] a) S. Angioletti-Uberti, B. M. Mognetti, D. Frenkel, Nat. Mater.

2012, 11, 518 – 522; b) R. J. Macfarlane, B. Lee, M. R. Jones, N.Harris, G. C. Schatz, C. A. Mirkin, Science 2011, 334, 204 – 208.

[7] F. J. Garc�a de Abajo, Rev. Mod. Phys. 2007, 79, 1267 – 1290.[8] a) M. Grzelczak, J. Perez-Juste, P. Mulvaney, L. M. Liz-Marzan,

Chem. Soc. Rev. 2008, 37, 1783 – 1791; b) Y. Xia, Y. Xiong, B.Lim, S. E. Skrabalak, Angew. Chem. 2009, 121, 62 – 108; Angew.Chem. Int. Ed. 2009, 48, 60 – 103.

Figure 4. Optical enhancement of nanoparticle clusters with coordina-tion numbers from 1 to 7. a) Comparison between the enhancementfactors obtained for each sample, normalized to the enhancementproduced by a single particle excited with a 633 nm laser line. Inset:SERS spectra of benzenethiol on the pentagonal bipyramid (CN 7).b) Near-field intensity (log thermal scale) for low-CN clusters in air at633 nm along planes passing by the centers of the lower spheres. Themaximum intensity relative to the incident intensity jEext j 2 is 14,10100, 5400, and 4000 for CN =1–4, respectively. c) Experimental datafitting with the following expression EF =CN + AG, where G repre-sents the number of gaps and A is a constant factor reflecting theeffective enhancement of the gap relative to the surface of an isolatedsphere.

.AngewandteCommunications

12692 www.angewandte.org � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 12688 –12693

Page 6: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

[9] C. J. Loweth, W. B. Caldwell, X. Peng, A. P. Alivisatos, P. G.Schultz, Angew. Chem. 1999, 111, 1925 – 1929; Angew. Chem. Int.Ed. 1999, 38, 1808 – 1812.

[10] L. C. Brousseau Iii, J. P. Novak, S. M. Marinakos, D. L. Feld-heim, Adv. Mater. 1999, 11, 447 – 449.

[11] J.-H. Lee, G.-H. Kim, J.-M. Nam, J. Am. Chem. Soc. 2012, 134,5456 – 5459.

[12] G. Von White, M. G. Provost, C. L. Kitchens, Ind. Eng. Chem.Res. 2012, 51, 5181 – 5189.

[13] M. Hanauer, S. Pierrat, I. Zins, A. Lotz, C. Sçnnichsen, NanoLett. 2007, 7, 2881 – 2885.

[14] D. Steinigeweg, M. Sch�tz, M. Salehi, S. Schl�cker, Small 2011, 7,2443 – 2448.

[15] a) D.-K. Lim, K.-S. Jeon, H. M. Kim, J.-M. Nam, Y. D. Suh, Nat.Mater. 2010, 9, 60 – 67; b) D.-K. Lim, K.-S. Jeon, J.-H. Hwang, H.Kim, S. Kwon, Y. D. Suh, J.-M. Nam, Nat. Nanotechnol. 2011, 6,452 – 460.

[16] R. Sardar, T. B. Heap, J. S. Shumaker-Parry, J. Am. Chem. Soc.2007, 129, 5356 – 5357.

[17] a) W. Li, P. H. C. Camargo, L. Au, Q. Zhang, M. Rycenga, Y.Xia, Angew. Chem. 2010, 122, 168 – 172; Angew. Chem. Int. Ed.2010, 49, 164 – 168; b) L. Cheng, J. Song, J. Yin, H. Duan, J. Phys.Chem. Lett. 2011, 2, 2258 – 2262; c) G. B. Braun, S. J. Lee, T.Laurence, N. Fera, L. Fabris, G. C. Bazan, M. Moskovits, N. O.Reich, J. Phys. Chem. C 2009, 113, 13622 – 13629.

[18] a) J. A. Fan, C. Wu, K. Bao, J. Bao, R. Bardhan, N. J. Halas, V. N.Manoharan, P. Nordlander, G. Shvets, F. Capasso, Science 2010,328, 1135 – 1138; b) V. N. Manoharan, M. T. Elsesser, D. J. Pine,Science 2003, 301, 483 – 487; c) G. Meng, N. Arkus, M. P.Brenner, V. N. Manoharan, Science 2010, 327, 560 – 563.

[19] a) T. I. Abdullin, O. V. Bondar, Y. G. Shtyrlin, M. Kahraman, M.Culha, Langmuir 2010, 26, 5153 – 5159; b) R. A. �lvarez-Puebla,R. Contreras-C�ceres, I. Pastoriza-Santos, J. P�rez-Juste, L. M.Liz-Marz�n, Angew. Chem. 2009, 121, 144 – 149; Angew. Chem.Int. Ed. 2009, 48, 138 – 143.

[20] R. A. Alvarez-Puebla, L. M. Liz-Marzan, Chem. Soc. Rev. 2012,41, 43 – 51.

[21] J. Rodr�guez-Fern�ndez, J. P�rez-Juste, F. J. Garc�a de Abajo,L. M. Liz-Marz�n, Langmuir 2006, 22, 7007 – 7010.

[22] C. S. Wagner, Y. Lu, A. Wittemann, Langmuir 2008, 24, 12126 –12128.

[23] P. N. Njoki, I. I. S. Lim, D. Mott, H.-Y. Park, B. Khan, S. Mishra,R. Sujakumar, J. Luo, C.-J. Zhong, J. Phys. Chem. C 2007, 111,14664 – 14669.

[24] C. Novo, A. M. Funston, I. Pastoriza-Santos, L. M. Liz-Marz�n,P. Mulvaney, Angew. Chem. 2007, 119, 3587 – 3590; Angew.Chem. Int. Ed. 2007, 46, 3517 – 3520.

[25] a) F. J. Garc�a de Abajo, A. Howie, Phys. Rev. Lett. 1998, 80,5180 – 5183; b) F. J. Garc�a de Abajo, A. Howie, Phys. Rev. B2002, 65, 115418.

[26] P. B. Johnson, R. W. Christy, Phys. Rev. B 1972, 6, 4370 – 4379.[27] Y. A. Urzhumov, G. Shvets, J. A. Fan, F. Capasso, D. Brandl, P.

Nordlander, Opt. Express 2007, 15, 14129 – 14145.[28] M. Shanthil, R. Thomas, R. S. Swathi, K. George Thomas, J.

Phys. Chem. Lett. 2012, 3, 1459 – 1464.[29] R. Jin, Angew. Chem. 2010, 122, 2888 – 2892; Angew. Chem. Int.

Ed. 2010, 49, 2826 – 2829.[30] F. J. Garc�a de Abajo, Phys. Rev. B 1999, 60, 6086 – 6102.[31] F. J. Garc�a de Abajo, J. Phys. Chem. C 2008, 112, 17983 – 17987.

AngewandteChemie

12693Angew. Chem. Int. Ed. 2012, 51, 12688 –12693 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Page 7: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Supporting Information

� Wiley-VCH 2012

69451 Weinheim, Germany

Organized Plasmonic Clusters with High Coordination Number andExtraordinary Enhancement in Surface-Enhanced Raman Scattering(SERS)**Nicolas Pazos-Perez, Claudia Simone Wagner, Jose M. Romo-Herrera, Luis M. Liz-Marz�n,F Javier Garc�a de Abajo,* Alexander Wittemann,* Andreas Fery,* and Ram�n A. Alvarez-Puebla*

anie_201207019_sm_miscellaneous_information.pdf

Page 8: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Supplementary Information:

Organized plasmonic clusters with high coordination number and extraordinary SERS

enhancement Nicolas Pazos-Pérez,1,T Claudia Simone Wagner,2,T Jose M. Romo-Herrera,3,4 Luis M. Liz-Marzán,4,5,6 F.

Javier García de abajo,7 Alexander Wittemann,2,* Andreas Fery,1,* Ramón A. Alvarez-Puebla8,9*

1Physikalische Chemie II, Universität Bayreuth, Universitätsstr. 30, 95440 Bayreuth, Germany; 2Colloid

chemistry, University of Konstanz, Universitaetsstrasse 10, 78464 Konstanz, Germany; 3 Centro de

Nanociencias y Nanotecnología, UNAM, km 107 Carretera Tijuana-Ensenada, C.P. 22860, Ensenada

B.C., México; 4Departamento de Química-Física, Universidade de Vigo, 36310 Vigo, Spain; 5Basque

Foundation for Science, 48011 Bilbao, Spain. 6Centre for Cooperative Research in Biomaterials (CIC

biomaGUNE), Paseo de Miramón 182, 20009 San Sebastián, Spain; 7Instituto de Quimica-Fisica

Rocasolano-CSIC, Serrano 119, 28006, Madrid, Spain; 8ICREA, Passeig Lluís Companys 23, 08010

Barcelona, Spain. 9Department of Electronic Engineering & Center for Chemical Technology of

Catalonia. Universitat Rovira i Virgili, Avda. Països Catalans, 2643007 Tarragona, Spain

*e-mail: [email protected]; [email protected]; [email protected]

TN.P.-P. and C.S.W. have contributed equally to the present work.

Page 9: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Experimental Methods

Chemicals. Trisodium citrate, tetrachloroauric(III) acid, sodium borohydride, ascorbic acid, and

cetyltrimethylammonium bromide (CTAB), glycerol (all from Merk) toluene and cyclohexane (Riedel-de

Haën), Pluronic®F-68 (PF68, Sigma) and benzenethiol (BT, Aldrich) were of analytical grade and used

as received. Deionized water (Millipore Academic A10) was used throughout the experiments.

Synthesis of gold nanoparticles. The gold nanoparticles (50 nm) were synthesized by a seed-mediated

approach, encompassing first the production of small gold seeds by preparing a solution (20 mL)

containing trisodium citrate (2.5 10-4 M) and HAuCl4 (2.5 10-4 M). The solution was vigorously stirred

while sodium borohydride (0.1 M, 600 µL) was added. After stirring for 1 h, 500 µL of the seeds were

added to a growth solution (250 mL) containing CTAB (0.1 M), HAuCl4 (1.08 M, 1213 µL) and ascorbic

acid (0.1 M, 1837 µL).

Stabilization and concentration of Au-CTAB with pluronic. To ensure particle-size monodispersity

and stabilization in PF68, 250 mL of the Au solution (2.510-4 M) were centrifuged (8000 rpm, 27 C, 10

min) and redispersed three times. After each centrifugation, precipitate was dispersed in PF68 (1 w%) and

supernatant discarded. The nanoparticle dispersion was left undisturbed during ~12 h at 27 C. Then, the

precipitate was discarded and the supernatant centrifuged three more times (8000rpm, 27 C, 10 min),

dispersing the precipitates in PF68 and discarding the supernatants. The final volume was 27 mL (1.2510-

3 M in gold)

Cluster fabrication from oil-in-water emulsions. The assembly of the gold nanospheres into clusters

was accomplished through the agglomeration of the gold spheres while adsorbed onto toluene droplets.

Briefly, suspensions of gold particles in 0.5 wt.% aqueous solution of PF68 were prepared. Subsequently,

0.4 mL toluene was added to 27 mL of these suspensions. Emulsification was accomplished with an

ultrasonic horn that was dipped into the sample, so that the end of the sonotrode was touching the oil–

water interphase. This procedure, in combination with the used geometry of the rosette cell (Bandelin RZ

3), ensured efficient mixing of the two phases immediately after sonication. The ultrasonication steps

Page 10: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

were performed under ice-cooling using a high-shear homogenizer (Bandelin Sonoplus HD 3200, 200 W,

probe KE 76). Three sonication steps, each for 5 min with 2 min rest in between, were performed at 30%

amplitude and a frequency of 20 kHz. Packing of the particles into clusters was induced by evaporation of

the toluene using a rotary evaporator (Heidolph Laborota 4000 efficient).

Cluster separation. Fractionation of the suspensions into clusters of the same coordination number

was accomplished through density gradient centrifugation. Concentration gradients were built from 40

wt.% and 60 wt.% aqueous glycerol solutions using a gradient maker. The clusters of different sizes were

separated by their sedimentation velocity during centrifugation at 5000 rpm for 6 min (Kontron

Instruments Centrikon T-1080, Sorvall Surespin 630 rotor). The cluster fractions were isolated through a

piercing unit (Kontron Instruments). Glycerol was removed by centrifugation and redispersion of the

particles in deionized water.

Characterization. Vis-NIR spectra were recorded using an Agilent 8453 diode array

spectrophotometer. Zeta potential was determined using a Malvern Instruments Zetasizer 2000. Scanning

electron microscopy (SEM) images were obtained with a JEOL JSM 6700F field-emission microscope.

High-resolution transmission electron microscopy (TEM) was conducted with a 2010 FEG-TEM

microscope operating at an acceleration voltage of 200 kV. For single-particle optical characterization,

diluted dispersions of clusters were spin-coated onto Lovins Field Finder Microslides (cat. 72266-01,

Electron Microscopy Sciences) and localized and imaged with the SEM. Dark-field imaging and

spectroscopy of the Au clusters was carried out on an inverted optical microscope (Nikon Eclipse TE-

2000) equipped with an Acton SpectraPro 2150i monochromator and a Princeton Instruments Pixis 1024

charge coupled device (CCD). Samples were illuminated by white light from a 100 W tungsten lamp

through a dark-field condenser (NA = 0.8). The scattered light was collected with a 40× objective (LD

Plan-NEOFLUAR, NA = 0.6) and reflected to the entrance slit of the monochromator for imaging and

spectroscopy. Scattering spectra from individual Au clusters were corrected by subtracting background

spectra taken from the adjacent regions containing no Au nanoparticles. SERS experiments were

conducted in a micro-Renishaw InVia Reflex system. The spectrograph uses a high resolution grating

Page 11: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

(1200 grooves cm- 1 for the NIR) with additional band-pass filter optics, a confocal microscope, a 2D-

CCD camera and an automatized stage of 100 nm spatial resolution. Excitation was carried out by

separately using laser lines at 632 nm and 785 nm. Benzenethiol was adsorbed in gas phase over the

whole surface of the films by casting a drop of BT (0.1 M in ethanol) in a Petri box where the film was

also contained. Surfaces were then mapped using the Renishaw StreamLine accessory with a step size of

1.2 µm in the same areas where SEM and dark-field data were collected. Acquisition times were set to

360 ms with power at the sample of 0.2 mW.

Theoretical calculations. Electromagnetic simulations were carried out using a multiple elastic-

scattering of multipolar expansions (MESME) method,[30] in which the fields are expressed in terms of

multipoles centered around each of the particles, and multiple scattering is carried out between them

following a highly convergent Lanczos iterative procedure. The far-field amplitude was computed from

the self-consistent expansion coefficients to generate the extinction cross-sections of Fig. 3, whereas the

near fields were obtained by direct evaluation of the multipole expansion sums. Gold spheres of 50 nm in

diameter were considered, with the metal described through its measured dielectric function[26] and

including a 0.6 nm coating of refractive index 1.3 to effectively represent the linking copolymer layer.

The gold-to-gold gap distance was set to 1.2 nm in all cases. The clusters were placed in a medium of

refractive index 1.47 in Fig. 3, and in air in Fig. 4.

Page 12: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Figure S1. Spectra for clusters in air: Same as Fig. 3b of the main paper for particles placed in air. The plasmon resonances are blueshifted away from the wavelengths used for SERS in this work (dotted arrows).

Additional Theoretical Discussion

A tight-binding analysis of the plasmon resonances in the clusters produces revealing results to understand the

long-wavelength features in the extinction spectra:

Figure S2. Near-field maps for a CN=2 cluster in oil and air upon excitation with ligth from 633 to 800 nm. The plots

show ecuatorial cuts with the external field along the gap.

We use the +− dipole mode confined to the gap of the dimer in oil as the unit wave function of our tight-binding

analysis, at wavelength λ~720 nm (energy E0~1.7 eV).

Page 13: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Figure S3. Near-field maps for a CN=3 cluster in oil and air upon excitation with ligth from 633 to 800 nm. The plots

show ecuatorial cuts of the three spheres.

The trimer consists of three gaps, with an inter-gap interaction ∆. This leads to modes of energies determined by

the diagonalization of the interaction matrix, consisting of zeros in the diagonal and with all off-diagonal elements

set to ∆. This produces a symmetric mode with energy E0+2∆ and equal weights in all gaps, as well as two

degenerate modes of energy E0−∆. The symmetric mode does not have a net dipole moment, and thus, it should not

couple efficiently to the external light. It is actually absent from the calculations shown in Fig. 3b of the main

paper. In contrast, the degenerate modes have a dipole moment and are responsible for the prominent long-

wavelength feature, which is blueshifted in the trimer with respect to the dimer, thus suggesting an attractive

interaction between gaps (∆<0), consistent with the +− gap-mode charge structure. The symmetric mode shows up

in the spectra at its expected position (redshifted with respect to the dimer by twice as much as the blueshift of the

bright modes), presumably as a result of asymmetries in the actual cluster.

Page 14: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Figure S4. Near-field maps for a CN=4 cluster in oil and air upon excitation with ligth from 633 to 800 nm. The plots

show ecuatorial cuts of the three lower spheres.

Figure S5. Near field maps for a CN=4 cluster in oil and air upon excitation with ligth from 633 to 800 nm. Vertical

cuts.

From a similar analysis of the CN=4 cluster, we find two sets of three degenerate modes at energies E0−2∆ and

E0+2∆ (six modes in total, as we have six gaps in tetrahedral cluster). The former are the only ones that display a

net dipole, and thus, they produce a dominant spectral feature, which is correctly redshifted with respect to the

dimer by twice as much as the trimer plasmon in the calculated spectrum of Fig. 3b of the main paper. The

experimental spectrum for the tetrahedron (Fig. 3a of main paper) is already showing some signatures of distortion

produced by the immersion oil.

Figure S6. Near field maps for a CN=5 cluster in oil and air upon excitation with ligth from 633 to 800 nm. Ecuatorial

cuts of the three central spheres.

Page 15: Andreas Fery,* and Ramn A. Alvarez-Puebla* · 2017-12-11 · Angewandte Chemie Nanoparticle Clusters DOI: 10.1002/anie.201207019 Organized Plasmonic Clusters with High Coordination

Figure S7. Near field maps for a CN=5 cluster in oil and air upon excitation with ligth from 633 to 800 nm. Vertical

cuts.

A similar analysis can be easily performed for higher CN clusters. This analysis, and the agreement with

experiment for low CN, indicates that the gap modes form a good basis set to describe the optical

behavior of the clusters. Higher CN clusters seem to be deformed by the immersion oil, and thus direct

comparison with experiment is not possible, but the tight-binding model produces very good results

compared with full MESME calculations. This study also demonstrates that the gaps act rather

independently under off-resonance illumination (i.e., for the conditions used to acquire ours SERS spectra

with the clusters in air), and thus, the number of gaps is the main magnitude that determines the SERS

enhancement factor of a given cluster.