Plasmonics (2016) 11:493–501 DOI 10.1007/s11468-015-0075-3 Analyzing Photothermal Heat Generation Efficiency in a Molecular Plasmonic Silver Nanomatryushka Dimer Arash Ahmadivand 1 & Nezih Pala 1 Received: 2 June 2015 / Accepted: 10 September 2015 / Published online: 22 September 2015 # Springer Science+Business Media New York 2015 Abstract We report on the numerically and analytically investigated plasmonic and photothermal responses of a nanomatryushka structure composed of silver concentric nanoshells which exhibited strong plasmon resonance localization in the optical frequencies. Illuminating an isolated silver nanomatryushka in an aqueous system, we calculated the photothermal response of the structure and quantified the absorbed optical power and generated photothermal heat. In addition, it is shown that a couple of nanomatryushka structures as a symmetric molecular dimer in weak and strong coupling regimes are able to support strong plasmon resonances in the visible to the near-infrared region. Utilizing strong near-field coupling in the metallic nanostructures and hybridization of plasmons, and also employing silver as a highly absorptive material at the visible spectrum, we increased the energy dissipation per unit volume almost three orders of magnitude in comparison to the other analogous subwavelength structures. Employing numerical methods, we showed that a symmetric metallic nanomatryushka dimer is able to generate enough photothermal heat which could result in a remarkable amount of temperature change (ΔT= 140 K) at the picosecond time scale. According to hybridization theory, the symmetric dimer is able to support strong bonding and antibonding plasmon resonant modes. Utilizing concentric nanoshells with high geometrical tunability facilitates using all of the surfaces and center of nanoparticles to generate heat with a large temperature change within a short * Arash Ahmadivand [email protected] 1 Department of Electrical and Computer Engineering, Florida International University, 10555 W Flagler St., Miami, FL 33174, USA relaxation time. This understanding opens new avenues to utilize simple nanoparticle orientations to generate significant heat power in an extremely short time scale for cancer therapy, photothermal therapy, and biological applications. Keywords Plasmonics . Nanomatryushka dimer . Photothermal heat . Plasmon hybridization . Numerical modeling Introduction Nanomatryushka (NM) was introduced as one of the strategic nanoparticle (NP) arrangements to characterize the plasmon resonance hybridization theory in simple and complex colloidal NP aggregates in molecular dimensions [1–3]. As a pioneer work, Prodan et al. [3] have investigated the plasmon response of a gold (Au) NM during strong localization of plasmon resonances across the visible spectrum to the nearinfrared region (NIR). In terms of physical properties, a NM consists of two concentric metallic nanoshells embedded and filled with a dielectric material (e.g., SiO2), entirely. The attractive behavior of a NM originates from the controllability of plasmon resonances in a wide range of optical spectrum by changing its geometrical and chemical properties. Having large number of independent and tunable geometrical parameters, this configuration yields a remarkable geometrical tunability, which can be exploited to enhance the electromagnetic (EM) field localization and hybridization in subwavelength dimensions. In terms of chemical characteristics, Au has widely been used for designing and fabricating of NM units for various purposes [1–4]. As another conventional noble metal, silver (Ag) has been proposed as a suitable material for plasmonic applications in the visible to the NIR, which is wellknown for its significant light absorption across the wide 494 range of spectrum [5, 6]. These features make the Ag as a proper choice in designing molecular clusters and structures to utilize in bio-chemical sensing [7], light harvesting and photovoltaic devices [8, 9], and surface-enhanced Raman spectroscopy (SERS) [10]. Considering the light-matter interaction effect in nanoscale regime, a simple metallic NM can be tailored to support strong bonding and antibonding plasmon resonances. It is wellknown that a bright mode can be excited via light-matter interactions directly with a net dipole moment, whereas an antibonding mode has no net dipole moment and, therefore, cannot be excited by regular light-matter interactions [11, 12]. Adjusting two equivalent nanoparticle structures in a close proximity to each other results in the formation of a molecular dimer. In a cluster-type structure, the appeared bonding and antibonding modes of each one of particles (here is NM) can be interfered with each other in destructive and constructive regimes. In the retarded (quasi-static) limit, a weak and destructive interference between opposite modes (bonding and antibonding modes) gives rise to the formation of pronounced Fano resonant dips in the extinction profile. Therefore, tremendous localization of plasmon resonances at the Fano dip frequency is expected, and such a huge amount of hot electron energy can be controlled effectively based on the behavior of Fano resonant mode [13, 14]. Additionally, intense localized surface plasmon resonance (LSPR) modes at the resonance frequency (ωLSPR) can result in scattering and absorption of optical power at this point, while the ratio of scattering to absorption cross sections depends on the structural and chemical properties of a metallic nanostructure. By careful selection of these parameters, the scattering and absorption efficiencies can be controlled effectively. On the other hand, conversion of light into usable photothermal heat energy in an extremely short time scale in plasmonic components has been considered as one of fundamental mechanisms in cancer therapy [15, 16], photothermal therapy and imaging [17–20], and vapor generation in the aqueous systems [21, 22]. For instance, in cancer therapy applications, illuminating metallic NPs with an external infrared laser delivers optical power to a targeted zone as a treatment mechanism [23]. Therefore, developing a process to design noble metallic (e.g., Ag, Au, Cu) structures to harvest large amount of incident light helps to fabricate numerous efficient and practical nanoscale devices for clinical applications. Moreover, the shape and size of NPs, their orientation, and surrounding medium have fundamental impacts on the enhancement of the light into heat generation efficiency. Very recently, Toroghi et al. [24] have studied the photothermal response of compositional (Au-Ag-Au) nanospheres as a heterogenous molecular trimer, wherein the proposed nanostructure shows twofold of absorption and heat generation in a finite and isolated subwavelength structure. In terms of lightmatter interactions, an oscillating free electron gas Plasmonics (2016) 11:493–501 interchanges the heat energy with the metallic part of the structure via electron-photon scattering in few picoseconds. The result of this process is generation of photothermal heat energy in subwavelength dimensions from the absorbed power [25, 26]. Due to the ultra-fast conversion of the absorbed photons into heat power and vice versa (cooling) in nanoscale structures, they have a significant potential for realizing beneficial devices for photothermal clinical applications. This mechanism can be understood by considering the photothermal effect in plasmonic structures, wherein the strong localization and hybridization of resonant modes can be ensued by magnifying the heating energy in a molecular system. Most of the earlier works have utilized cascaded plasmon resonances to enhance the photothermal effect as well as intensifying the EM field for heat generation with photothermal heat variations in the range of ~100 K [24, 27]. In this study, we utilize a couple of molecular silver nanomatryushka (Ag-NM) units in a dimer orientation with the experimentally determined Palik constants [28] to generate significant photothermal heat to achieve as high as possible photothermal heat via hybridization of plasmon resonances. Assuming a controllable thermal atmosphere (a liquefied environment) as the ambience for our nanostructure, we numerically calculate extinction spectra to determine the produced heat energy in the light into heat conversion process. Moreover, by comparing the temperature of the produced heat for an isolated Ag-NM and a couple of Ag-NM units (dimer), we demonstrate the effect of structural dimensions on the light into heat conversion mechanism by using Mie [29], plasmon hybridization [3, 11], and discrete dipole approximation (DDA) theories [30]. Theory Heat generation in nanoscale dimensions in an extremely short time in a fluid medium includes numerous events, which make the entire process tremendously complex to analyze theoretically and numerically [31–33]. The photothermal response and produced temperature in a light-matter interaction in subwavelength dimensions can be calculated by considering the physical and optical properties of NP as well as the characteristics of the ambient in analytical computations [34, 35]. Using this scenario, one would be able to determine the amount of incident light absorption by a plasmonic nanostructure and the efficiency of light into heat conversion. Currently, there are two major methods that can be employed to study the photothermal response and heat generation for the presented Ag-NM structure. The general mathematical model to describe the relation between temperature changes (ΔT) and absorption coefficient can be understood by using time-dependent Fourier law equation for thermal Plasmonics (2016) 11:493–501 495 conductivity as q(t)=−κ∇T(t) W m−2 [36]. Consequently, using the latest equation, for the quasi-static EM field we have [37]: ∂ΔT 1 C abs ð∇:κÞ∇ΔT þ ¼ ρNM cNM ∂t ρNM cNM ð1Þ where κ is the thermal conductivity (W m−1K−1),cNM is the specific heat capacity of a sample NM (J kg−1K−1), t is the illumination time, ΔT is the temperature changes in a specific time scale (K), ρNM is the density of the nanosize structure (kg m−3), and C abs is the absorption coefficient for the Ag-NM unit that can be obtained by calculating the absorption spectra under illumination by a light source or laser pump exposure. More than the described parameters in the differential equation above, also, the intensity of the incident EM field (laser source) plays a major role in the light into heat conversion efficiency. The major problem correlating with the proposed method above (Eq. 1) is the independency of the equation to the volume of the structure [37]. It should be underlined that the volume of metallic nanostructures has significant influence in light absorption and plasmon resonance localization. This problem can be mitigated by including the volume of a NP in the calculation of the absorption coefficient and photothermal heat generation. It is shown that the equation above (Eq. 1) can be written in a simple and new form as below [24], which directly depends on the physical and chemical characteristics of the proposed AgNM unit: ΔT C abs ð2Þ ¼ cNM V NM ρNM φ where φ is the optical flux of the incident light source pulse (J m−2) (determined by the light source setting) and V NM is the volume of a NM unit in Bm3.^ Therefore, determining accurate geometrical dimensions for the proposed Ag-NM and illuminating the structure with strong EM field could help generate remarkable heat during lightmatter interaction. In the current work, the behavior of phonons correlating with the NP (hot object) and medium (cool object) and their transport without appreciable interaction is used to characterize the heat generation process, wherein mean free path of these phonons is on the same order with the dimensions of the NPs. As an important parameter in photothermal computations, the absorption cross section can be defined by integration of the absorbed optical power at the LSPR peak position over the volume of the Ag-NM and then the obtained value is divided by the simulation settings as below: Z ffi . rffiffiffiffiffi ε0 ⊥ 2 C abs ¼ 2 PLSPR dV nm ð3Þ E μ0 inc NM Where PLSPR is the optical power absorbed by the Ag-NM unit(s) based on the ohmic losses at the resonance frequency, nm is the refractive index of the surrounding medium, ε0and μ0 are the permittivity and permeability of the vacuum or free space, andĒ⊥incis the amplitude of the incident transverse electric field emitted by a laser source. The intensity of the incoming electric field is set to 1 mW μm−2 in all of the simulations. More precisely, increasing both volume and the specific heat capacity causes to the formation of undesired reduction in the temperature of the produced photothermal heat (see Eq. 2). Therefore, providing a method to absorb large amounts of incident optical energy with a small volume of NP and also using strong localization of plasmon resonances in a few nanometer spot lead to the formation of remarkable thermal power. In nanoplasmonic structures, the resonance frequency/ wavelength (λLSPR) has a significant impact on the behavior of a nanoscale structure. In the quasi-static limit for a nanoscale structure [36], the absorption coefficient at the resonance frequency is given by the following: λ LSPR C abs ¼ 18πV NM n3m λLSPR ImfεNM g 0 ð4Þ 0 0 is the complex permittivity of mewhere εNM ¼ εNM þ εNM tallic NM. A simple NM facilitates excitation of surface plasmon resonances by using all of the outer and inner surfaces and central spots of the NPs and hence allows for producing thermal power with a significant temperature change. Also, the geometrical tunability of a NM provides high efficiency in optical power absorption, making it superior to simple nanoparticles such as nanosphere, rod, disk, and shell. Figure 1a illustrates three- and two-dimensional artworks for the examined Ag-NM with the description of geometrical parameters that is filled and surrounded by a glass (SiO2) host with the permittivity of ε=3.9, and the entire symmetric AgNM is completely immersed in a fluid system includes water (H2O) with the refractive index of n=1.33, analogous to biological applications. The heat energy generation (Qh) can be computed by considering the absorption cross-sectional coefficient at the resonance positions and the intensity of the incident spectrum as following equation [35]: Qλh LSPR ¼ 2 1 LSPR nm c0 ε0 E ⊥inc C λabs 2 ð5Þ where c0is the velocity of light in the free space (m s−1), and the absorption coefficient is correlated with the dissipated power coefficient at the LSPR wavelength. It is already verified that in Eq. 5, the term 0.5nc0ε0|E⊥inc|2refers to the properties of the propagating electric field in a free space ambient [36, 37]. Thus, calculation of the optical power absorption for the Ag-NM is enough for thermal heat quantification. To determine the absorption efficiency for an Ag-NM unit, we examined the optical response by modifying the size of NM 496 Plasmonics (2016) 11:493–501 Fig. 1 a A schematic diagram of the Ag-NM with the description of the geometrical dimensions. b Extinction spectra for the Ag-NM under transverse electric polarization mode excitation. Inset is the schematic picture of the fluid system. c Twodimensional snapshots for the field intensity |E| in the NM for the absorption peaks at 470, 620, and 835 nm under illumination in a liquid system. Our numerical studies are performed by employing finite-difference time-domain (FDTD) method (Lumerical FDTD Solutions 2015 package), and the fundamental parameters are set as follows: the spatial cell sizes are set to dx =dy =dz =0.1 nm, with 200,000 numbers of cells, and perfectly matched layers (PMLs) as the boundary condition with 128 layers. In addition, considering numerical stability for the employed subwavelength structures, simulation time step is set to the 0.02 fs according to the Courant stability condition. The light source is a Gaussian electric pulse with a pulse length of 2.6533 fs, offset time of 7.5231 fs, the waist radius of the incident beam is 600 nm, and the divergence angle is set to 9.04°. Results and Discussion The photothermal response of an Ag-NM in single and dimer orientations can be achieved by using a theoretical model as described already in the BTheory^ section. To this end, we used Ag [38], SiO2 [39], and water [40] with the thermal conductivity of (κ) of 429, 0.92, and 0.609 W m−1 K−1, respectively. In addition, to define the spectral response for the proposed NM, we used a modified version of Drude model [41]: εAl ¼ ε∞ − ω2B ωðω þ iγ Þ ð6Þ where ε∞ is the high-frequency response (~5.7), ωB is the bulk plasma frequency (1.3736×1016 rad/s), and γ is the damping frequency (2.7335×1015 Hz), in which these parameters are set to experimentally determined values by Murata et al. [42] to calculate the effective dielectric function for an Ag-NP. Figure 1b shows the numerically calculated extinction crosssectional profile for an isolated Ag-NM with the default size of 35/75/95/115 nm in the wavelength range of λ=400– 1300 nm. To provide a detailed work, we examined the influence of all variable structural parameters (Ra/Rb/Rc/Rd) on the absorption efficiency. Using these geometries, the net and effective volume of a concentric structure composed of two nanoshells is given by the following: VNM = πh{(R2b −R2a) −(R2d −R2c )}. The height of NPs are set to h=80 nm in all of the future analysis. In the extinction diagram, three distinct peaks are observed, which are correlated with the dipolar and multipolar plasmon resonant modes at λ=470, 620, and 835 nm. This broad absorption characteristic including three maxima in the visible to the NIR is the advantageous feature of the proposed nanostructure. This absorption feature facilitates for heat generation in a wide range of spectrum with minor differences in efficiencies. The inset in Fig. 1b is a schematic for a liquid system under exposure by a laser diode, while nanoparticles are immersed thoroughly in an aqueous system. As it is expected, due to the inherent properties of Ag material, the absorption behavior of the Ag-NM is dominant in comparison to the scattering diagram in the extinction spectra. Figure 1c shows two-dimensional snapshots of the electric Plasmonics (2016) 11:493–501 497 field intensity (|E|) and plasmon resonance excitation and hybridization inside the isolated Ag-NM at the plasmon resonance maxima under transverse polarization excitation, where the role of the space (filled with dielectric substance) between two nanoshells is obvious in all snapshots. To examine the effect of geometrical sizes on the absorption profile, we compared the behavior of an Ag-NM in Fig. 2a by calculating normalized absorption profile. For the proposed Ag-NM with the size of 35/75/95/115 nm, we observed three distinct shoulders correlating with the absorption peaks at the visible to the NIR spectra (λ~480, 650, and 840 nm). By increasing the size of the inner shell (45/85/100/115 nm), we experienced a destructive reduction in the light absorption, and accordingly, the absorption peaks blue-shifted to the shorter spectra. In contrast, for 45/85/110/135 nm, we increased the size of the outer nanoshell which gives rise to the formation of strong red-shift in the position of absorption peaks to the longer wavelengths, where this shift includes a significant increase in the normalized absorption amplitude. This absorption enhancement originates from the formation of bright modes during light-matter interactions within the outer shell and leads to support strong resonances by our metallic NM. Indeed, the exterior and interior nanoshells interact with the incident Gaussian beam, and the result of this interference is the formation of bonding plasmon modes in the energy continuum. Due to the small distance between two concentric nanoshells (<10 nm), dipolar and multipolar modes interfering with each other includes possible destructive interferences between opposite modes (bonding↔ antibonding). The result of this destructive interference is significant intensification in the absorption coefficient. The other reason for the remarkable power absorption is related to the contribution of all of the surfaces of the NM structure in the excitation process. Figure 2b compares the light absorption (nW nm−3) for the proposed NM with two different geometries. The amount of the absorbed power (P) by both NP systems is quite large due to the Fig. 2 a Normalized absorption spectra of the Ag-NM with different dimensions under transverse electric polarization excitation, b power absorption snapshots for two Ag-NMs different dimensions, c a comparison of the numerically calculated thermal power (Qh) generated by Ag-NMs different dimensions, and d numerically calculated scattering spectra for NM dimer for four different geometrical sizes 498 Plasmonics (2016) 11:493–501 plasmon resonances localization and absorption at the resonance positions. This absorption spectra can be utilized efficiently in photothermal heat generation, and the produced heat (Qh) correlated with the LSPR frequency/wavelength can be estimated by using the modified version of Eq. 5, as: ¼ QTotal h 2 X λ 1 LSPR nm c0 ε0 E⊥inc C abs 2 λ ð7Þ LSPR where λLSPR is the LSPR wavelength for an Ag-NM under transverse polarization excitation. Figure 2c compares the computed entire photothermal heat for a single Ag-NM with different geometrical sizes. The most important and effective parameter in the thermal heat generation spectroscopy is the resultant temperature variations. Employing Eq. 2, and using the heat capacity of Ag cNM =233 J kg−1K−1 [43], we calculated the temperature change per optical fluence rate ΔT/φ=45 K μJ−1 mm2 for an Ag-NM which gives temperature changes as ΔT=90 K for a pulse fluence of φ=20 nJ mm−2. Comparing the obtained temperature changes for an isolated Ag-NM unit with the other types of NPs such as nanospheres and nanorods in a liquid system [24, 27, 44, 45], we proved a superior behavior for the examined NMs. By far, we calculated the photothermal heat for an isolated Ag-NM; however, this photothermal heat generation can be further enhanced by forming a dimer cluster using the examined NM. Figure 2d displays the scattering cross-sectional profiles for the proposing dimer composed of similar NM units with four different geometrical dimensions. As obvious as it is, for the structure with following geometries: D 1 = (45/85/110/ 135) nm, D2 =(35/75/95/115) nm, and D3 =(40/80/100/ 115) nm, a distinct asymmetric Fano dip is performed around λ~800 nm. However, this minimum can be even enhanced using different geometries. Locating a couple of NM units with the following dimensions: D4 =(45/85/ 110/135) nm in a few nanometer distance from each other (with the gap spacing of 10 nm), we induced an asymmetric and deep Fano minimum including a redshift to the NIR spectra. Figure 3a illustrates a schematic picture of the Ag-NM dimer (Ag-NMD) with the offset gap distance of Dg. Using the earlier mechanism to illuminate an isolated Ag-NM, we examined the plasmon and photothermal responses of an Ag-NMD in a liquid system. With the placement of two equivalent Ag-NM units close to each other, we defined the effect of gap distance alterations on the thermal heat generation numerically. It is well-accepted that the intensity and strength of localized plasmon resonances can be changed significantly by modifying the interparticle distance [46, 47]. In addition, it is shown that for high values of Dg, the coupling strength is remarkably weak Fig. 3 a A three-dimensional schematic diagram for an Ag-NMD. b–d Numerically calculated scattering and absorption spectra for an Ag-NMD under transverse electric polarization excitation and hybridization cannot be achieved due to missing of the required energy for dipolar and multipolar Plasmonics (2016) 11:493–501 499 Fig. 4 a–c Two-dimensional snapshots of the field intensity |E| inside the Ag-NMD for the absorption peaks. d–f Power absorption spots snapshots for an Ag-NMD under transverse polarization excitation interactions [48, 49]. Therefore, for small gap distances, we expect strong coupling between dipolar and multipolar plasmon resonant modes, called strong coupling regime. Strong localization of plasmon resonances in a molecular cluster composed of metallic NPs can be resulted with a large scattering of EM field. Formation of bonding and antibonding plasmon modes and placement of two metallic NPs in a close proximity to each other Fig. 5 Numerically calculated thermal power (Qh) for an Ag-NMD with three different offset gap distances leads to intensification of plasmon modes (hybridization mechanism). In this regime, a symmetry cancelation in the structural properties directly gives rise to the formation of new collective antibonding modes and a destructive interference between bonding and antibonding modes result in a Fano resonance dip as a hotspot [49–51]. Our proposed Ag-NMD is a symmetric structure and formation of Fano resonances in such a symmetric molecular arrangement is challenging. However, recently, Ahmadivand et al. [52] proposed a method to obtain Fano dips in an Al/Al2O3 NM dimer in free space conditions. In a simple dimer consisting of a couple of complex NM units, the outer shell supports the bright mode via light-matter interactions and the inner shell provides the required multipolar dark modes to generate Fano dip via a weak and destructive interference between dark and bright plasmon energies. For our Ag nanostructure, due to the strong absorption property in the visible to the NIR spectra, we expect a distinct Fano minimum as a hotspot in the scattering profile. Figure 3(b), (c), and (d) exhibit the scattering and absorption diagrams for an Ag-NMD with different offset gap distances Dg =25/10/5 nm. Due to superior behavior of Ag in comparison to Al for plasmonics purposes, effective support of strong plasmon resonances and interactions between the opposite modes are possible. In the calculated scattering profiles, we obtained a noticeable minimum around λ~800 nm by reducing the size 500 of offset gap distance between two NP arrangements. Bringing two NMs closer to each other causes to the formation of a significant hotspot including a subtle redshift in the position of Fano dip. The peak of the calculated normalized absorption profile is observed at the same position with the Fano resonances and originates from the strong absorption of plasmon resonances by two Ag-NMs. Comparing the absorption figures for Ag-NMD with an isolated Ag-NM, a noticeable enhancement in the absorption efficiency is achieved with a broadening of the absorption peak. Figure 4a–c are numerically calculated simulation snapshots for the plasmon resonance excitation and hybridization in an AgNMD in an aqueous system for different gap distances (Dg) from 25 to 5 nm, which resembles a transmutation of weak coupling of resonant modes to the strong coupling regime. Figure 4d–f depicts the power absorbed by Ag-NM units under illumination by a laser pump. Increasing the intensity of the coupling of the plasmon resonances (hybridization regime) gives rise to a dramatic power absorption enhancement in a molecular system that is in accordance with the absorption and scattering profiles (see Fig. 3b–d). Quantifying generated photothermal heat for the studied Ag-NMD with different gap sizes can help to understand the effect of hybridization on the absorption of the incident light and photothermal heat generation. Using the mechanism that has been presented in the earlier section, we determined the photothermal heat power (Qh) as demonstrated in Fig. 5. The amount of photothermal power increases by the transmutation from weak to strong coupling regime. The obtained thermal power for the dimer structure is almost twofold higher than the that of the previous regime and even higher than that of the analogous dimers and trimers. Finally, we computed the temperature change ratio per optical fluence of the incident pulse (J m-2) as ΔT/φ= 70 K μJ-1 mm2 for the dimer. Assuming a Gaussian pulse fluence (laser pump setting) of φ =20 nJ mm-2, the temperature variations that can be obtained by an Ag-NMD colloidal structure is quantified as ΔT=140 K. Plasmonics (2016) 11:493–501 nanoparticles in photothermal heat generation. It is verified that hybridization of plasmon resonances in such a symmetric and molecular structure composed of NM units can be resulted by pronounced Fano dip as a hotspot in short time scale. We exploited these unique features to achieve extremely fast heating up to 140 K and fast cooling down to the liquefied environmental temperature. Unlike regular spherical NP aggregates, concentric nanoshells support strong plasmon and Fano resonances in the visible to the NIR spectra during photothermal heating process. The results imply that the studied dimer is a strong candidate for various biological applications that requires short relaxation time in the range of picoseconds. Acknowledgments This work is supported by the NSF CAREER program with the award number: 0955013 and by the Army Research Laboratory (ARL) Multiscale Multidisciplinary Modeling of Electronic Materials (MSME) Collaborative Research Alliance (CRA) (Grant No. W911NF-12-2-0023, Program Manager: Dr. Meredith L. Reed). References 1. 2. 3. 4. 5. 6. 7. 8. 9. Conclusions In conclusion, we showed that generation of considerable photothermal heat energy in an Ag-NM structure in molecular levels is possible and practical in a really short time scale via light to heat conversion mechanism in a liquid medium. Despite of inherent symmetry of the proposed dimer composed of Ag-NMs, we induced strong hotspots by using an absorptive noble metallic material. Utilizing concentric nanoshells facilitates using all of the surfaces and center of the 10. 11. 12. Bardhan R, Mukherjee S, Mirin NA, Levit SD, Nordlander P, Halas NJ (2009) Nanosphere-in-a-nanoshell: a simple nanomatryushka. J Phys Chem C 114:7378–7383 Wu DJ, Cheng Y, Wu XW, Liu XJ (2014) An active metallic nanomatryushka with two similar super-resonances. J Appl Phys 116:013502 Prodan E, Radloff C, Halas NJ, Nordlander P (2003) A hybridization model for the plasmon response of complex nanostructure. Science 302:419–422 Kulkarni V, Prodan E, Nordlander P (2013) Quantum plasmonics: optical properties of a nanomatryushka. Nano Lett 13:5873–5879 Evanoff DD, Chumanov G (2005) Synthesis and optical properties of silver nanoparticles and arrays. Chem Phys Chem Phys 6:1221– 1231 Taleb A, Russier V, Courty A, Pileni MP (1999) Collective optical properties of silver nanoparticles organized in two-dimensional superlattices. Phys Rev B 59:13350 Liu S, Tang Z (2010) Nanoparticle assemblies for biological and chemical sensing. J Mater Chem 20:24–35 Wang P, Huang B, Dai Y, Whangbo MH (2012) Plasmonic photocatalysts: harvesting visible light with noble metal nanoparticles. Phys Chem Chem Phys 14:9813–9825 Bakr OM, Amendola VA, Aikens CM, Wenseleers W, Li R, Dan Negro L, Schatz GC, Stellacci F (2009) Silver nanoparticles with broad multiband linear optical absorption. Angew Chem Int Ed 121:6035–6040 Cao YWC, Jin R, Mirkin CA (2002) Nanoparticles with Raman spectroscopic fingerprints for DNA and RNA detection. Science 297:1536–1540 Fan JA, Wu C, Bao K, Bao J, Bardhan R, Halas NJ, Manoharan VN, Nordlander P, Shvets G, Capasso F (2010) Self-assembled plasmonic nanoparticle clusters. Science 328:1135–1138 Ahmadivand A, Golmohammadi S (2014) Electromagnetic plasmon propagation and coupling through the gold nanoring heptamers: a route to design optimized telecommunication photonic nanostructures. Appl Opt 53:3832–3840 Plasmonics (2016) 11:493–501 13. Brown LV, Sobhani H, Lassiter JB, Nordlander P, Halas NJ (2010) Heterodimers: plasmonic properties of mismatched nanoparticle pairs. ACS Nano 4:819–832 14. Verellen N, Van Dorpe P, Huang C, Lodewijks K, Vandenbosch GAE, Lagae L, Moshchalkov VV (2011) Plasmon line shaping using nanocrosses for high sensitivity localized surface plasmon resonance sensing. Nano Lett 11:391–397 15. El-Sayed IH, Huang X, El-Sayed MA (2006) Selective laser photothermal therapy of epithelial carcinoma using anti-EGFR antibody conjugated gold nanoparticles. Cancer Lett 239:129–135 16. Huang X, Qian W, El-Sayed IH, El-Sayed MA (2007) The potential use of the enhanced nonlinear properties of gold nanospheres in photothermal cancer therapy. Lasers Surg Med 39:747–753 17. Lapotko D, Lukianova E, Potapnev M, Aleinikova O, Oraevsky A (2006) Methods, of laser activated nano-thermolysis for elimination of tumor cells. Cancer Lett 239:36–45 18. Huang X, Jain PK, El-Sayed IH, El-Sayed MA (2007) Plasmonic photothermal therapy (PPTT) using gold nanoparticles. Lasers Med Sci 23:217–228 19. Munidasa M, Mandelis A (1991) Photopyrelectric thermal-wave tomography of aluminum with ray-optic reconstruction. J Opt Soc Am A 8:1851–1858 20. Vermeulen P, Cognet L, Lounis B (2014) Photothermal microscopy: optical detection of small absorbers in scattering environments. J Microsc 254:115–121 21. Neumann O, Urban AS, Day J, Lal S, Nordlander P, Halas NJ (2012) Solar vapor generation enabled by nanoparticles. ACS Nano 7:42–49 22. Fang Z, Zhen Y–R, Neumann O, Polman A, Javier Garcia de Abajo F, Nordlander P (2013) Evolution of light-induced vapor generation at a liquid-immersed metallic nanoparticle. Nano Lett 13:1736– 1742 23. Gobin AM, Lee MH, Halas NJ, James WD, Drezek RA, West JL (2007) Near-infrared resonant nanoshells for combined optical imaging and photothermal cancer therapy. Nano Lett 7:1929–1934 24. Toroghi S, Kik P (2014) Photothermal response enhancement in heterogenous plasmon-resonant nanoparticle trimers. Phys Rev B 90:205414 25. Muller EA, Strader ML, John JE, Yang A, Caplins BW, Shearer AJ, Suich DE, Harris CB (2013) Femtosecond electron solvation at the ionic liquid/metal electrode interface. J Am Chem Soc 29:10646– 10653 26. Pusch A, Shadrivov IV, Hess O, Kivshar YS (2013) Self-focusing of femtosecond surface plasmon polaritons. Opt Express 21:1121– 1127 27. Lin H, Yi Z (2012) Double resonance 1-D photonic crystal cavities for single-molecule mid-infrared photothermal spectroscopy: theory and design. Opt Lett 37:1304–1306 28. Palik ED (1991) Handbook of optical constants of solids, 2nd edn. Academic Press, San Diego 29. Jackson JB, Halas NJ (2001) Silver nanoshells: variations in morphologies and optical properties. J Phys Chem B 105:2743–2746 30. Muinonen K, Zubko E, Tyynelä J, Shkuratov YG, Videen G (2007) Light scattering by Gaussian random particles with discrete-dipole approximation. J Quant Spectrosc Radiat Transf 106:360–377 31. Baffou G, Quidant R, Girard C (2009) Heat generation in plasmonic nanostructures: Influence of morphology. Appl Phys Lett 94: 153109 32. Uyar R, Erdogdu F, Marra F (2014) Effect of load volume on power absorption and temperature evolution during radio-frequency of 501 meat cubes: a computational struct. Food Bioprod Process 92: 243–251 33. Baptiste J, Fourier JB (1822) Théorie analytique de la chaleur, Chez Frimin Didot, Pére et fils, 34. Chen G (1996) Nonlocal and noneqilibrium heat conduction in the vicinity of nanoparticles. J Heat Trans T ASME 118:539–545 35. Govorov AO, Richardson HH (2007) Generating heat with metal nanoparticles. Nano Today 2:30–38 36. Baffou G, Quidant R (2013) Thermo-plasmonics: using metallic nanostructures as nano-sources of heat. Lasers Photon Rev 7:171– 187 37. Kim W, Wang R, Majumdar A (2007) Nanostructuring expands thermal limits. Nano Today 2:40–47 38. Lide DR (2004) Handbook of chemistry and physics, 84th ed., CRC Press, US 39. Kittel C (1949) Interpretation of the thermal conductivity of glasses. Phys Rev 75:972 40. Ramires MLV, Nieto de Castro CA, Nagasaka Y, Nagashima A, Assael MJ, Wakeham WA (1995) Standard reference data for the thermal conductivity of water. J Phys Chem Ref Data 24:1377 41. Maier SA (2007) Plasmonics: fundamentals and applications. Springer, NY 42. Murata I, Tanaka H (2010) Surface-wetting effects on the liquidliquid transition of a single-component molecular liquid. Nat Commun 1:1–9 43. Meads PF, Forsythe WR, Giauque WF (1941) The heat capacities and entropies of silver and lead from 15° to 300°K. J Am Chem Soc 63:1902–1905 44. Toroghi S, Lumdee C, Kik PG (2015) Heterogenous plasmonic trimers for enhanced nonlinear optical absorption. Appl Phys Lett 106:103012 45. Herzog JB, Knight MW, Natelson D (2014) Thermoplasmonics: quantifying plasmonic heating in single nanowires. Nano Lett 14: 499–503 46. Ahmadivand A, Karabiyik M, Pala N (2015) Intensifying magnetic dark modes in the antisymmetric plasmonic quadrumer composed of Al/Al2O3 nanodisks with the placement of silicon nanospheres. Opt Commun 338:218–225 47. Lassiter JB, Aizpurua J, Hernandez LI, Brandl DW, Romero I, Lal S, Hafner JH, Nordlander P, Halas NJ (2008) Close encounters between two nanoshells. Nano Lett 8:1212–1218 48. Ahmadivand A, Golmohammadi S (2014) Optimized plasmonic configurations: adjacent and merging regimes between a symmetric couple of Au rod/shell nano-arrangements for LSPR sensing and spectroscopic purposes. J Nanoparticle Res 16:2491 49. Luk’yanchuk B, Zheludev NI, Maier SA, Halas NJ, Nordlander P, Giessen H, Chong CT (2010) The Fano resonance in plasmonic nanostructures and metamaterials. Nat Mater 9:707–715 50. Golmohammadi S, Ahmadivand A (2014) Fano resonances in compositional clusters of aluminum nanodisks at the UV spectrum: a route to design efficient and precise biochemical sensors. Plasmonics 9:1447–1465 51. Nazir A, Panaro S, Zaccaria RP, Liberale C, De Angelis F, Toma A (2014) Fano coil-type resonance for magnetic hot-spot generation. Nano Lett 14:3166–3171 52. Ahmadivand A, Pala S (2015) Localization, hybridization, and coupling of plasmon resonances in an aluminum nanomatryushka, Plasmonics Online first article doi: 10.1007/s11468-014-9868-z
© Copyright 2025 Paperzz