Atomistic Multiscale Modeling of Colloidal Plasmonic Nanoparticles

Luca Nicoli Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy.    Sveva Sodomaco Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy.    Piero Lafiosca Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy.    Tommaso Giovannini Department of Physics, University of Rome Tor Vergata, Via della Ricerca Scientifica 1, 00133, Rome, Italy [email protected]    Chiara Cappelli [email protected] Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy.
Abstract

A novel fully atomistic multiscale classical approach to model the optical response of solvated real-size plasmonic nanoparticles (NPs) is presented. The model is based on the coupling of the Frequency Dependent Fluctuating Charges and Fluctuating Dipoles (ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ), specifically designed to describe plasmonic substrates, and the polarizable Fluctuating Charges (FQ) classical force field to model the solvating environment. The resulting ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approach accounts for the interactions between the radiation and the NP, as well as with the surrounding solvent molecules, by incorporating mutual interactions between the plasmonic substrate and solvent. ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is validated against reference TD-DFTB/FQ calculations, demonstrating remarkable accuracy, particularly in reproducing plasmon resonance frequency shifts for structures below the quantum-size limit. The flexibility and reliability of the approach are also demonstrated by simulating the optical response of homogeneous and bimetallic NPs dissolved in pure solvents and solvent mixtures.

1 Introduction

In the past decades, colloidal nanoparticles (NPs), i.e. NPs dissolved in solution, have gained significant interest due to their applications in many technological contexts, such as sensing,1 biomedical applications,2 optoelectronics,3 and energy conversion.4 By choosing different precursors, reducing agents, solvents, and cap** agents, nanostructured materials can be synthesized with a fine control of shape and size.5 Size, shape, chemical composition, and the solvent can indeed affect the plasmon resonance frequency (PRF), i.e. the maximum of the NP absorption spectrum.6 Such a feature is the basis of a particular class of sensors, which exploit the shift of the PRF upon change of the local refractive index (RI) of the solvent in which the plasmonic NPs are dissolved. Such devices have been widely employed in biosensing,7, 8 where maximizing the induced PRF shift as a function of RI is crucial to enhance sensitivity.1, 9

Rationalizing solvent effects on the PRF of colloidal NPs is particularly challenging from a theoretical point of view. In fact, the optical properties of colloidal NPs result from the interplay of complex phenomena originating under the action of the external electric field, such as the appearance of a localized surface plasmon (LSP) excitation and the polarization of the solvent electron cloud. Solvent effects on PRFs result from a delicate balance between NP-solvent electrostatic (and polarization) interactions, charge-transfer effects, and the possible alteration of the plasmon decaying channels.10

In principle, a proper description of all possible NP-solvent effects would require an ab initio treatment of the whole system. However, first principle approaches become rapidly unfeasible due to their unfavorable scaling as a function of the system’s size, thus hindering the simulation of realistic systems. For these reasons, commonly exploited theoretical approaches to simulate plasmonic colloidal NPs are rooted in classical physics,11, 12, 13, 14, 15 generally making use of the classical Mie theory11, the Boundary Element Method (BEM),12 the Discrete Dipole Approximation (DDA),13 and the Finite Difference Time Domain (FDTD).14 However, all these approaches are based on approximated descriptions of both the NP and the solvent, which is generally modeled as a continuum dielectric characterized by its specific permittivity ε𝜀\varepsilonitalic_ε. In addition, such approaches do not retain the atomistic nature of the system.

In this paper, we propose a novel multiscale method where both the plasmonic NP and the solvent are treated at full atomistic level. In particular, we employ the fully atomistic electromagnetic model called Frequency Dependent Fluctuating Charges Fluctuating Dipoles (ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ)16 to describe the optical response of plasmonic NPs. Such a model is remarkably versatile and can be applied to NPs of any shape16 and chemical composition,17 even at the quantum size limit (<<< 5 nm).16 The solvent is modeled employing the polarizable Fluctuating Charges (FQ) force field18, 19, 20, which is specifically designed to model the polarization of the electron cloud of molecular systems. For this reason, it has been widely exploited in the context of computational spectroscopy of solvated systems.21, 22, 23, 24

The two approaches, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ and FQ, are coupled in a multiscale fashion so that the resulting ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ model accounts for the mutual electrostatic interaction between the solvent and the plasmonic NP, allowing for the modeling of the optical properties of generic colloidal plasmonic NPs. ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is also coupled to classical molecular dynamics (MD) simulations, which are exploited to sample the NP-solvent phase space. Therefore, differently from previous methods,11, 11, 12, 13, 14 the dynamical aspects of the solvation phenomenon, which are crucial to properly model solvent effects on spectroscopy, are taken into account.25, 23, 21

Note that, if solvent molecules are not adsorbed on the surface of the NP, the main solvent effect on plasmonic properties is the local RI variation of the medium surrounding the nanomaterial. This leads to a modification of the local optical field, generally causing a PRF redshift.26, 10 Such effect is harnessed in many colorimetric-sensors for the detection of specific biomolecular analytes.7, 8, 27, 28, 29, 30, 11, 31, 32, 33, 34 In this work, we model the physical processes that lead to plasmon shift upon change of the local RI of the embedding medium.

The paper is organized as follows: first, the novel ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approach is presented after the theoretical foundations of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ and FQ methods are recalled. Then, the computational protocol is presented and ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is validated in comparison with reference Time-Dependent Density Functional Tight Binding/Fluctuating Charges (TD-DFTB/FQ35) results. The versatility and robustness of the approach are demonstrated by studying real-size homogeneous and bimetallic NPs dissolved in pure solvent or solvent mixtures. A summary and the main conclusions of the work end the manuscript.

2 Theory

Refer to caption
Figure 1: Pictorial view of the multiscale scheme employed to develop ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ.

In this section, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ and FQ are briefly recalled and the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approach for modeling the optical properties of colloidal plasmonic NPs is presented (see Fig. 1).

2.1 Plasmonic nanoparticle: the atomistic-electromagnetic ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ model

ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ models the plasmonic NP atomistically. Each atom is endowed with a set of complex frequency-dependent atom-centered charges 𝐪(ω)𝐪𝜔\mathbf{q}(\omega)bold_q ( italic_ω ) (ω𝜔\omegaitalic_ωFQs), and dipoles 𝝁(ω)𝝁𝜔\bm{\mu}(\omega)bold_italic_μ ( italic_ω ) (ω𝜔\omegaitalic_ωFμ𝜇\muitalic_μs) (see Fig. 1), accounting for intraband and interband mechanisms, respectively. Charges are obtained by solving the following equation of motion, obtained by modulating a Drude-like conduction mechanics with quantum tunneling:36, 37

iωqi(ω)=j(Ajnj(1fji(lij))1/τjiω+Aini(1fij(lij))1/τiiω)ϕi(ω)ϕj(ω)liji𝜔subscript𝑞𝑖𝜔subscript𝑗subscript𝐴𝑗subscript𝑛𝑗1subscript𝑓𝑗𝑖subscript𝑙𝑖𝑗1subscript𝜏𝑗i𝜔subscript𝐴𝑖subscript𝑛𝑖1subscript𝑓𝑖𝑗subscript𝑙𝑖𝑗1subscript𝜏𝑖i𝜔subscriptitalic-ϕ𝑖𝜔subscriptitalic-ϕ𝑗𝜔subscript𝑙𝑖𝑗-\mathop{}\!\mathrm{i}\omega q_{i}(\omega)=\sum_{j}\left(\frac{A_{j}n_{j}(1-f_% {ji}(l_{ij}))}{1/\tau_{j}-\mathop{}\!\mathrm{i}\omega}+\frac{A_{i}n_{i}(1-f_{% ij}(l_{ij}))}{1/\tau_{i}-\mathop{}\!\mathrm{i}\omega}\right)\frac{\phi_{i}(% \omega)-\phi_{j}(\omega)}{l_{ij}}- roman_i italic_ω italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( divide start_ARG italic_A start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( 1 - italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ( italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) ) end_ARG start_ARG 1 / italic_τ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - roman_i italic_ω end_ARG + divide start_ARG italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 1 - italic_f start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) ) end_ARG start_ARG 1 / italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - roman_i italic_ω end_ARG ) divide start_ARG italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) - italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG (1)

In eq.1 Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the atomic effective area, the electron density, and the relaxation time associated with the intraband scattering events of the i𝑖iitalic_i-th atom, respectively. Quantum tunneling effects are expressed in terms of a Fermi-like dam** function fij(lij)subscript𝑓𝑖𝑗subscript𝑙𝑖𝑗f_{ij}(l_{ij})italic_f start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ), which exponentially damps the charge exchange between the atoms (lijsubscript𝑙𝑖𝑗l_{ij}italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the distance between i𝑖iitalic_i-th and j𝑗jitalic_j-th atoms). ϕi(ω)subscriptitalic-ϕ𝑖𝜔\phi_{i}(\omega)italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) is the chemical potential of i𝑖iitalic_i-th atom, which reads:

ϕi(ω)=Viq(ω)+Viμ(ω)+Viext(ω)subscriptitalic-ϕ𝑖𝜔subscriptsuperscript𝑉𝑞𝑖𝜔subscriptsuperscript𝑉𝜇𝑖𝜔subscriptsuperscript𝑉ext𝑖𝜔\phi_{i}(\omega)=V^{q}_{i}(\omega)+V^{\mu}_{i}(\omega)+V^{\mathrm{ext}}_{i}(\omega)italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) = italic_V start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + italic_V start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + italic_V start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) (2)

where Viextsubscriptsuperscript𝑉ext𝑖V^{\mathrm{ext}}_{i}italic_V start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the electric potential associated with the optical radiation, whereas Viqsubscriptsuperscript𝑉𝑞𝑖V^{q}_{i}italic_V start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Viμsubscriptsuperscript𝑉𝜇𝑖V^{\mu}_{i}italic_V start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the electric potentials induced by charges and dipoles on the i𝑖iitalic_i-th atomic site (see Ref.16 for more details).

The plasmonic properties of alkali metals and graphene in the Pauli-blocking regime are correctly described by ω𝜔\omegaitalic_ωFQs which properly models intraband mechanisms.36, 37, 38 However, when the interband absorption energy threshold is comparable to the plasmon resonance frequency (PRF), such as in noble metal nanoparticles, interband transitions become relevant to the decaying mechanism.39 To model this process, an additional complex frequency-dependent dipole 𝝁(ω)𝝁𝜔\bm{\mu}(\omega)bold_italic_μ ( italic_ω ) is assigned to each atom:

𝝁i(ω)subscript𝝁𝑖𝜔\displaystyle\bm{\mu}_{i}(\omega)bold_italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) =αiIB(ω)𝐄itot(ω)absentsuperscriptsubscript𝛼𝑖𝐼𝐵𝜔subscriptsuperscript𝐄tot𝑖𝜔\displaystyle=\alpha_{i}^{IB}(\omega)\ \mathbf{E}^{\mathrm{tot}}_{i}(\omega)= italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I italic_B end_POSTSUPERSCRIPT ( italic_ω ) bold_E start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) (3)
=αiIB(ω)[𝐄iq(ω)+𝐄iμ(ω)+𝐄iext(ω)]absentsuperscriptsubscript𝛼𝑖𝐼𝐵𝜔delimited-[]subscriptsuperscript𝐄𝑞𝑖𝜔subscriptsuperscript𝐄𝜇𝑖𝜔subscriptsuperscript𝐄ext𝑖𝜔\displaystyle=\alpha_{i}^{IB}(\omega)\left[\mathbf{E}^{q}_{i}(\omega)+\mathbf{% E}^{\mu}_{i}(\omega)+\mathbf{E}^{\mathrm{ext}}_{i}(\omega)\right]= italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I italic_B end_POSTSUPERSCRIPT ( italic_ω ) [ bold_E start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + bold_E start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + bold_E start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) ] (4)

where αiIB(ω)superscriptsubscript𝛼𝑖𝐼𝐵𝜔\alpha_{i}^{IB}(\omega)italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I italic_B end_POSTSUPERSCRIPT ( italic_ω ) is the interband frequency-dependent polarizability of the i𝑖iitalic_i-th atom, and 𝐄itot(ω)subscriptsuperscript𝐄tot𝑖𝜔\mathbf{E}^{\mathrm{tot}}_{i}(\omega)bold_E start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) is the total electric field acting on the i𝑖iitalic_i-th dipole, accounting for charge-dipole (𝐄qsuperscript𝐄𝑞\mathbf{E}^{q}bold_E start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT), dipole-dipole (𝐄μsuperscript𝐄𝜇\mathbf{E}^{\mu}bold_E start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT), and dipole-field (𝐄extsuperscript𝐄ext\mathbf{E}^{\text{ext}}bold_E start_POSTSUPERSCRIPT ext end_POSTSUPERSCRIPT) interactions (see Ref. 16 for more details). For homogeneous materials, αiIB(ω)superscriptsubscript𝛼𝑖𝐼𝐵𝜔\alpha_{i}^{IB}(\omega)italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I italic_B end_POSTSUPERSCRIPT ( italic_ω ) is determined from the frequency-dependent bulk-permittivity, whereas in the case of multimetallic systems, the interband polarizability of the i𝑖iitalic_i-th atom is expressed as a function of the local composition of the system. 16, 17

The ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ charge-dipole coupled equations can be recast in the following set of complex linear equations:

(𝐀(ω)𝐓qq𝐙(ω)𝐀(ω)𝐓qμ𝐓μq𝐓μμ𝐙IB(ω))(𝐪𝝁)=(𝐀(ω)𝐕ext𝐄ext)matrix𝐀𝜔superscript𝐓qq𝐙𝜔𝐀𝜔superscript𝐓q𝜇superscript𝐓𝜇qsuperscript𝐓𝜇𝜇superscript𝐙IB𝜔matrix𝐪𝝁matrix𝐀𝜔superscript𝐕extsuperscript𝐄ext\begin{pmatrix}\mathbf{A}(\omega)\mathbf{T}^{\text{qq}}-\mathbf{Z}(\omega)&% \mathbf{A}(\omega)\mathbf{T}^{\text{q}\mu}\\ \mathbf{T}^{\mathrm{\mu q}}&\mathbf{T}^{\mathrm{\mu\mu}}-\mathbf{Z}^{\mathrm{% IB}}(\omega)\end{pmatrix}\begin{pmatrix}\mathbf{q}\\ \bm{\mu}\end{pmatrix}=\begin{pmatrix}-\mathbf{A}(\omega)\mathbf{V}^{\mathrm{% ext}}\\ -\mathbf{E}^{\mathrm{ext}}\end{pmatrix}( start_ARG start_ROW start_CELL bold_A ( italic_ω ) bold_T start_POSTSUPERSCRIPT qq end_POSTSUPERSCRIPT - bold_Z ( italic_ω ) end_CELL start_CELL bold_A ( italic_ω ) bold_T start_POSTSUPERSCRIPT q italic_μ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL bold_T start_POSTSUPERSCRIPT italic_μ roman_q end_POSTSUPERSCRIPT end_CELL start_CELL bold_T start_POSTSUPERSCRIPT italic_μ italic_μ end_POSTSUPERSCRIPT - bold_Z start_POSTSUPERSCRIPT roman_IB end_POSTSUPERSCRIPT ( italic_ω ) end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL bold_q end_CELL end_ROW start_ROW start_CELL bold_italic_μ end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL - bold_A ( italic_ω ) bold_V start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL - bold_E start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) (5)

where 𝐀(ω)𝐀𝜔\mathbf{A}(\omega)bold_A ( italic_ω ) is a frequency-dependent matrix containing NP chemical and geometrical parameters, while 𝐙(ω)𝐙𝜔\mathbf{Z}(\omega)bold_Z ( italic_ω ), and 𝐙IB(ω)superscript𝐙IB𝜔\mathbf{Z}^{\mathrm{IB}}(\omega)bold_Z start_POSTSUPERSCRIPT roman_IB end_POSTSUPERSCRIPT ( italic_ω ) are diagonal matrices. 𝐓qqsuperscript𝐓qq\mathbf{T}^{\mathrm{qq}}bold_T start_POSTSUPERSCRIPT roman_qq end_POSTSUPERSCRIPT, 𝐓qμsuperscript𝐓q𝜇\mathbf{T}^{\mathrm{q\mu}}bold_T start_POSTSUPERSCRIPT roman_q italic_μ end_POSTSUPERSCRIPT, 𝐓μqsuperscript𝐓𝜇q\mathbf{T}^{\mathrm{\mu q}}bold_T start_POSTSUPERSCRIPT italic_μ roman_q end_POSTSUPERSCRIPT and 𝐓μμsuperscript𝐓𝜇𝜇\mathbf{T}^{\mathrm{\mu\mu}}bold_T start_POSTSUPERSCRIPT italic_μ italic_μ end_POSTSUPERSCRIPT are charge-charge, charge-dipole, dipole-charge, and dipole-dipole interaction kernels, respectively (see Ref. 40 for more details).

When the linear system in Eq. 5 is solved, charges and dipoles modeling the optical intra- and interband response of the plasmonic NP are obtained. From such variables, the NP complex polarizability and the absorption cross-section σabs(ω)superscript𝜎abs𝜔\sigma^{\mathrm{abs}}(\omega)italic_σ start_POSTSUPERSCRIPT roman_abs end_POSTSUPERSCRIPT ( italic_ω ) can be computed (See Ref. 16 and Ref. 17 for further details about the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ model).

2.2 Solvent: the Fluctuating Charges model (FQ)

The physics governing the optical response of solvent systems is utterly different from that of plasmonic materials. When external electric fields are applied, solvent molecules might experience several phenomena ranging from electronic transitions to molecular rovibrations depending on the external field frequency. In the following, we focus only on solvents that are transparent in the spectral region where the plasmonic nanostructure absorbs light. Note that this is generally the case for noble metal nanoparticles, whose PRF falls within the visible range (400 - 700 nm).41, 42 In this way, the external fields exciting the plasmons have energies lower than those of the electronic transitions of the solvent molecules but sufficiently high to quench rovibrational effects completely. Thus, the interaction with the external field only yields the polarization of the solvent’s electron cloud. This is modeled by using the Fluctuating Charges (FQ) force field,18, 19, 20 which has been extensively used in computational quantum chemistry for the modeling molecules in solution.21, 43, 23 Within FQ, each solvent atom is endowed with a charge 𝐪~~𝐪\tilde{\mathbf{q}}over~ start_ARG bold_q end_ARG whose value is not fixed but can vary as a response to the external electric potential (see Fig. 1). Such charge fluctuation is governed by the electronegativity equalization principle (EEP), 44, 45 which states that at equilibrium, each atom has the same electronegativity, i.e. the negative of the chemical potential, as reported by Parr.46 The FQ energy, i.e. the energy required to create a partial charge on each atom is generally written as a second-order Taylor expansion in the charges. For a polyatomic system, this can be written as:18

U(𝐪~)=α,i(χαiq~αi+12ηαiq~αi2+V~αitotq~αi)+β,k<α,iq~βkTβk,αiq~q~q~αi𝑈~𝐪subscript𝛼𝑖subscript𝜒𝛼𝑖subscript~𝑞𝛼𝑖12subscript𝜂𝛼𝑖subscriptsuperscript~𝑞2𝛼𝑖subscriptsuperscript~𝑉𝑡𝑜𝑡𝛼𝑖subscript~𝑞𝛼𝑖subscriptformulae-sequence𝛽𝑘𝛼𝑖subscript~𝑞𝛽𝑘subscriptsuperscript𝑇~𝑞~𝑞𝛽𝑘𝛼𝑖subscript~𝑞𝛼𝑖U(\tilde{\mathbf{q}})=\sum_{\alpha,i}\left(\chi_{\alpha i}\tilde{q}_{\alpha i}% +\frac{1}{2}\eta_{\alpha i}\tilde{q}^{2}_{\alpha i}+\tilde{V}^{tot}_{\alpha i}% \tilde{q}_{\alpha i}\right)+\sum_{\beta,k<\alpha,i}\tilde{q}_{\beta k}T^{% \tilde{q}\tilde{q}}_{\beta k,\alpha i}\tilde{q}_{\alpha i}italic_U ( over~ start_ARG bold_q end_ARG ) = ∑ start_POSTSUBSCRIPT italic_α , italic_i end_POSTSUBSCRIPT ( italic_χ start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_η start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT + over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_β , italic_k < italic_α , italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_β italic_k end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β italic_k , italic_α italic_i end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT (6)

where Greek and Roman indices run over molecules and atoms, respectively. χαisubscript𝜒𝛼𝑖\chi_{\alpha i}italic_χ start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT is the electronegativity of the i𝑖iitalic_i-th atom of the α𝛼\alphaitalic_α-th molecule, and 𝐓q~q~superscript𝐓~𝑞~𝑞\mathbf{T}^{\tilde{q}\tilde{q}}bold_T start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT is the charge-charge interaction kernel40 of which the diagonal elements ηαisubscript𝜂𝛼𝑖\eta_{\alpha i}italic_η start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT are chemical hardnesses, representing self-interaction polarization term. Both electronegativity and chemical hardness are well rooted in Conceptual Density Functional Theory 47 and are the only free parameters defining the model. Moreover, 𝐕~totsuperscript~𝐕𝑡𝑜𝑡\tilde{\mathbf{V}}^{tot}over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT is the total external electric potential on each atomic site, which in the FQ case is the potential associated with the external optical field (𝐕~tot=𝐕~extsuperscript~𝐕𝑡𝑜𝑡superscript~𝐕𝑒𝑥𝑡\tilde{\mathbf{V}}^{tot}=\tilde{\mathbf{V}}^{ext}over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT = over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_e italic_x italic_t end_POSTSUPERSCRIPT). To constrain the charge of each molecule Qαtotsubscriptsuperscript𝑄𝑡𝑜𝑡𝛼Q^{tot}_{\alpha}italic_Q start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT to a constant, a set of Lagrange multipliers λαsubscript𝜆𝛼\lambda_{\alpha}italic_λ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is exploited. The energy functional in Eq. 6 thus becomes:

F(𝐪~,𝝀)=U(𝐪~)+αλα(i(q~αi)Qαtot)𝐹~𝐪𝝀𝑈~𝐪subscript𝛼subscript𝜆𝛼subscript𝑖subscript~𝑞𝛼𝑖subscriptsuperscript𝑄𝑡𝑜𝑡𝛼F(\tilde{\mathbf{q}},\bm{\lambda})=U(\tilde{\mathbf{q}})+\sum_{\alpha}\lambda_% {\alpha}\left(\sum_{i}(\tilde{q}_{\alpha i})-Q^{tot}_{\alpha}\right)italic_F ( over~ start_ARG bold_q end_ARG , bold_italic_λ ) = italic_U ( over~ start_ARG bold_q end_ARG ) + ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ) - italic_Q start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) (7)

From the minimization of Eq. 7 with respect to the variables (𝐪~~𝐪\tilde{\mathbf{q}}over~ start_ARG bold_q end_ARG and 𝝀𝝀\bm{\lambda}bold_italic_λ), the FQ polarization equations are obtained, implying that the electronic degrees of freedom of the solvent instantaneously rearrange without energy dissipation.18 Such an approximation is also valid for optical fields in the visible range and for transparent solvents.48, 21, 22 When the total electric field is monochromatic at frequency ω𝜔\omegaitalic_ω, the FQ polarization equations in the frequency domain read as follows (See Sec. S1.1 in the Supporting Information SI  for further details):

(𝐓q~q~𝟏𝟏t𝟎)(𝐪~(ω)𝝀(ω))=(𝐕~ext(ω)0)matrixsuperscript𝐓~𝑞~𝑞1superscript1𝑡0matrix~𝐪𝜔𝝀𝜔matrixsuperscript~𝐕𝑒𝑥𝑡𝜔0\begin{pmatrix}\mathbf{T}^{\tilde{q}\tilde{q}}&\mathbf{1}\\ \mathbf{1}^{t}&\mathbf{0}\end{pmatrix}\begin{pmatrix}\tilde{\mathbf{q}}(\omega% )\\ \bm{\lambda}(\omega)\end{pmatrix}=\begin{pmatrix}-\tilde{\mathbf{V}}^{ext}(% \omega)\\ 0\end{pmatrix}( start_ARG start_ROW start_CELL bold_T start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL bold_1 end_CELL end_ROW start_ROW start_CELL bold_1 start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL over~ start_ARG bold_q end_ARG ( italic_ω ) end_CELL end_ROW start_ROW start_CELL bold_italic_λ ( italic_ω ) end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL - over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_e italic_x italic_t end_POSTSUPERSCRIPT ( italic_ω ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARG ) (8)

where 𝟏1\mathbf{1}bold_1 is a rectangular matrix containing the blocks associated with the Lagrange multipliers. By solving Eq. 8, the atom-centered FQ charges, modeling the instantaneous polarization of the electronic cloud of the solvent under the application of an external monochromatic field, are obtained.

2.3 Optical response of a plasmonic nanoparticle in solution: the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ model

ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ for describing the nanostructure’s response and FQ for modeling solvent polarization are coupled in a multiscale fashion. In the resulting ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approach, mutual polarization effects between the solvent and the metal NP are introduced (see Fig. 1). To this end, we include the solvent-induced electric potential 𝐕q~(ω)superscript𝐕~𝑞𝜔\mathbf{V}^{\tilde{q}}(\omega)bold_V start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT ( italic_ω ) and electric field 𝐄q~(ω)superscript𝐄~𝑞𝜔\mathbf{E}^{\tilde{q}}(\omega)bold_E start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT ( italic_ω ) acting on the NP’s charges 𝐪(ω)𝐪𝜔\mathbf{q}(\omega)bold_q ( italic_ω ) and dipoles 𝝁(ω)𝝁𝜔\bm{\mu}(\omega)bold_italic_μ ( italic_ω ) respectively. Thus, the total chemical potential and field (Eq. 2 and Eq. 4 respectively) acting on each atomic site now read:

ϕi(ω)subscriptitalic-ϕ𝑖𝜔\displaystyle\phi_{i}(\omega)italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) =Viq(ω)+Viμ(ω)+Viq~(ω)+Viext(ω)absentsubscriptsuperscript𝑉𝑞𝑖𝜔subscriptsuperscript𝑉𝜇𝑖𝜔subscriptsuperscript𝑉~𝑞𝑖𝜔subscriptsuperscript𝑉ext𝑖𝜔\displaystyle=V^{q}_{i}(\omega)+V^{\mu}_{i}(\omega)+V^{\tilde{q}}_{i}(\omega)+% V^{\mathrm{ext}}_{i}(\omega)= italic_V start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + italic_V start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + italic_V start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + italic_V start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) (9)
𝐄itot(ω)subscriptsuperscript𝐄𝑡𝑜𝑡𝑖𝜔\displaystyle\mathbf{E}^{tot}_{i}(\omega)bold_E start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) =𝐄iq(ω)+𝐄iμ(ω)+𝐄iq~(ω)+𝐄iext(ω)absentsubscriptsuperscript𝐄𝑞𝑖𝜔subscriptsuperscript𝐄𝜇𝑖𝜔subscriptsuperscript𝐄~𝑞𝑖𝜔subscriptsuperscript𝐄ext𝑖𝜔\displaystyle=\mathbf{E}^{q}_{i}(\omega)+\mathbf{E}^{\mu}_{i}(\omega)+\mathbf{% E}^{\tilde{q}}_{i}(\omega)+\mathbf{E}^{\mathrm{ext}}_{i}(\omega)= bold_E start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + bold_E start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + bold_E start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) + bold_E start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) (10)

The electric potential and field induced by the FQ solvent charges 𝐪~(ω)~𝐪𝜔\tilde{\mathbf{q}}(\omega)over~ start_ARG bold_q end_ARG ( italic_ω ) on the NP’s atomic sites read:

Viq~(ω)subscriptsuperscript𝑉~𝑞𝑖𝜔\displaystyle V^{\tilde{q}}_{i}(\omega)italic_V start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) =α,kTi,αkqq~q~αk(ω)absentsubscript𝛼𝑘subscriptsuperscript𝑇𝑞~𝑞𝑖𝛼𝑘subscript~𝑞𝛼𝑘𝜔\displaystyle=\sum_{\alpha,k}T^{q\tilde{q}}_{i,\alpha k}\tilde{q}_{\alpha k}(\omega)= ∑ start_POSTSUBSCRIPT italic_α , italic_k end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT italic_q over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_α italic_k end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_k end_POSTSUBSCRIPT ( italic_ω ) (11)
𝐄iq~(ω)subscriptsuperscript𝐄~𝑞𝑖𝜔\displaystyle\mathbf{E}^{\tilde{q}}_{i}(\omega)bold_E start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω ) =α,k𝐓i,αkμq~q~αk(ω)absentsubscript𝛼𝑘subscriptsuperscript𝐓𝜇~𝑞𝑖𝛼𝑘subscript~𝑞𝛼𝑘𝜔\displaystyle=\sum_{\alpha,k}\mathbf{T}^{\mu\tilde{q}}_{i,\alpha k}\tilde{q}_{% \alpha k}(\omega)= ∑ start_POSTSUBSCRIPT italic_α , italic_k end_POSTSUBSCRIPT bold_T start_POSTSUPERSCRIPT italic_μ over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_α italic_k end_POSTSUBSCRIPT over~ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_α italic_k end_POSTSUBSCRIPT ( italic_ω ) (12)

𝐓qq~superscript𝐓𝑞~𝑞\mathbf{T}^{q\tilde{q}}bold_T start_POSTSUPERSCRIPT italic_q over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT and 𝐓μq~superscript𝐓𝜇~𝑞\mathbf{T}^{\mu\tilde{q}}bold_T start_POSTSUPERSCRIPT italic_μ over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT are the Coulomb interaction kernels between the FQ solvent charges and ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ plasmonic charges and dipoles respectively, i.e.:

Tijqq~superscriptsubscript𝑇𝑖𝑗𝑞~𝑞\displaystyle T_{ij}^{q\tilde{q}}italic_T start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT =1|𝐫ij|absent1subscript𝐫𝑖𝑗\displaystyle=\frac{1}{|\mathbf{r}_{ij}|}= divide start_ARG 1 end_ARG start_ARG | bold_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT | end_ARG (13)
𝑻ijμq~superscriptsubscript𝑻𝑖𝑗𝜇~𝑞\displaystyle\bm{T}_{ij}^{\mu\tilde{q}}bold_italic_T start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT =riTijqq~absentsubscriptsubscript𝑟𝑖superscriptsubscript𝑇𝑖𝑗𝑞~𝑞\displaystyle=\mathbf{\nabla}_{r_{i}}T_{ij}^{q\tilde{q}}= ∇ start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q over~ start_ARG italic_q end_ARG end_POSTSUPERSCRIPT (14)

where 𝐫ijsubscript𝐫𝑖𝑗\mathbf{r}_{ij}bold_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the distance between atoms i𝑖iitalic_i and j𝑗jitalic_j.

To account for mutual polarization effects, the total potential 𝐕~totsuperscript~𝐕tot\tilde{\mathbf{V}}^{\mathrm{tot}}over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT acting on the solvent atomic sites includes the potential generated by the NP’s charges 𝐕~q(ω)superscript~𝐕𝑞𝜔\tilde{\mathbf{V}}^{q}(\omega)over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT ( italic_ω ) and dipoles 𝐕~μ(ω)superscript~𝐕𝜇𝜔\tilde{\mathbf{V}}^{\mu}(\omega)over~ start_ARG bold_V end_ARG start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT ( italic_ω ):

V~αitot(ω)=V~αiq(ω)+V~αiμ(ω)subscriptsuperscript~𝑉𝑡𝑜𝑡𝛼𝑖𝜔subscriptsuperscript~𝑉𝑞𝛼𝑖𝜔subscriptsuperscript~𝑉𝜇𝛼𝑖𝜔\tilde{V}^{tot}_{\alpha i}(\omega)=\tilde{V}^{q}_{\alpha i}(\omega)+\tilde{V}^% {\mu}_{\alpha i}(\omega)over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_t italic_o italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ( italic_ω ) = over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ( italic_ω ) + over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ( italic_ω ) (15)

where the electric potential and field induced by the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ charges and dipoles on the i𝑖iitalic_i-th solvent atom of the α𝛼\alphaitalic_α-th molecule read:

V~αiq(ω)subscriptsuperscript~𝑉𝑞𝛼𝑖𝜔\displaystyle\tilde{V}^{q}_{\alpha i}(\omega)over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ( italic_ω ) =kTαi,kq~qqk(ω)absentsubscript𝑘subscriptsuperscript𝑇~𝑞𝑞𝛼𝑖𝑘subscript𝑞𝑘𝜔\displaystyle=\sum_{k}T^{\tilde{q}q}_{\alpha i,k}q_{k}(\omega)= ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i , italic_k end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) (16)
V~αiμ(ω)subscriptsuperscript~𝑉𝜇𝛼𝑖𝜔\displaystyle\tilde{V}^{\mu}_{\alpha i}(\omega)over~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT ( italic_ω ) =k𝐓αi,kq~μ𝝁k(ω)absentsubscript𝑘subscriptsuperscript𝐓~𝑞𝜇𝛼𝑖𝑘subscript𝝁𝑘𝜔\displaystyle=\sum_{k}\mathbf{T}^{\tilde{q}\mu}_{\alpha i,k}\bm{\mu}_{k}(\omega)= ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT bold_T start_POSTSUPERSCRIPT over~ start_ARG italic_q end_ARG italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_i , italic_k end_POSTSUBSCRIPT bold_italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) (17)

Note that in Eq. 15 we neglect local field effects (i.e. V~αiext(ω)=0superscriptsubscript~𝑉𝛼𝑖ext𝜔0\tilde{{V}}_{\alpha i}^{\mathrm{ext}}(\omega)=0over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT ( italic_ω ) = 0).

The coupled ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ and FQ equations can thus be recast in a linear system defining the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ master equations:

(𝐀(ω)𝐓qq𝐙(ω)𝐀(ω)𝐓qμ𝐀(ω)𝐓qq~𝟎𝐓qμ𝐓μμ𝐙IB(ω)𝐓μq~𝟎𝐓q~q,t𝐓q~μ,t𝐓q~q~𝟏𝟎𝟎𝟏t𝟎)(𝐪𝝁𝐪~𝝀)=(𝐀(ω)𝐕ext(ω)𝐄ext(ω)00)matrix𝐀𝜔superscript𝐓qq𝐙𝜔𝐀𝜔superscript𝐓q𝜇𝐀𝜔superscript𝐓q~q0superscript𝐓𝑞𝜇superscript𝐓𝜇𝜇superscript𝐙IB𝜔superscript𝐓𝜇~q0superscript𝐓~qq𝑡superscript𝐓~q𝜇𝑡superscript𝐓~q~q100superscript1𝑡0matrix𝐪𝝁~𝐪𝝀matrix𝐀𝜔superscript𝐕ext𝜔superscript𝐄ext𝜔00\begin{pmatrix}\mathbf{A}(\omega)\mathbf{T}^{\mathrm{qq}}-\mathbf{Z}(\omega)&% \mathbf{A}(\omega)\mathbf{T}^{\mathrm{q\mu}}&\mathbf{A}(\omega)\mathbf{T}^{% \mathrm{q\tilde{q}}}&\mathbf{0}\\ \mathbf{T}^{q\mu}&\mathbf{T}^{\mathrm{\mu\mu}}-\mathbf{Z}^{\mathrm{IB}}(\omega% )&\mathbf{T}^{\mathrm{\mu\tilde{q}}}&\mathbf{0}\\ \mathbf{T}^{\mathrm{\tilde{q}q},t}&\mathbf{T}^{\mathrm{\tilde{q}\mu},t}&% \mathbf{T}^{\mathrm{\tilde{q}\tilde{q}}}&\mathbf{1}\\ \mathbf{0}&\mathbf{0}&\mathbf{1}^{t}&\mathbf{0}\end{pmatrix}\begin{pmatrix}% \mathbf{q}\\ \bm{\mu}\\ \tilde{\mathbf{q}}\\ \bm{\lambda}\end{pmatrix}=\begin{pmatrix}-\mathbf{A}(\omega)\mathbf{V}^{% \mathrm{ext}}(\omega)\\ -\mathbf{E}^{\mathrm{ext}}(\omega)\\ 0\\ 0\end{pmatrix}( start_ARG start_ROW start_CELL bold_A ( italic_ω ) bold_T start_POSTSUPERSCRIPT roman_qq end_POSTSUPERSCRIPT - bold_Z ( italic_ω ) end_CELL start_CELL bold_A ( italic_ω ) bold_T start_POSTSUPERSCRIPT roman_q italic_μ end_POSTSUPERSCRIPT end_CELL start_CELL bold_A ( italic_ω ) bold_T start_POSTSUPERSCRIPT roman_q over~ start_ARG roman_q end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_T start_POSTSUPERSCRIPT italic_q italic_μ end_POSTSUPERSCRIPT end_CELL start_CELL bold_T start_POSTSUPERSCRIPT italic_μ italic_μ end_POSTSUPERSCRIPT - bold_Z start_POSTSUPERSCRIPT roman_IB end_POSTSUPERSCRIPT ( italic_ω ) end_CELL start_CELL bold_T start_POSTSUPERSCRIPT italic_μ over~ start_ARG roman_q end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_T start_POSTSUPERSCRIPT over~ start_ARG roman_q end_ARG roman_q , italic_t end_POSTSUPERSCRIPT end_CELL start_CELL bold_T start_POSTSUPERSCRIPT over~ start_ARG roman_q end_ARG italic_μ , italic_t end_POSTSUPERSCRIPT end_CELL start_CELL bold_T start_POSTSUPERSCRIPT over~ start_ARG roman_q end_ARG over~ start_ARG roman_q end_ARG end_POSTSUPERSCRIPT end_CELL start_CELL bold_1 end_CELL end_ROW start_ROW start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_1 start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL bold_q end_CELL end_ROW start_ROW start_CELL bold_italic_μ end_CELL end_ROW start_ROW start_CELL over~ start_ARG bold_q end_ARG end_CELL end_ROW start_ROW start_CELL bold_italic_λ end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL - bold_A ( italic_ω ) bold_V start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT ( italic_ω ) end_CELL end_ROW start_ROW start_CELL - bold_E start_POSTSUPERSCRIPT roman_ext end_POSTSUPERSCRIPT ( italic_ω ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARG ) (18)

By solving Eq.18, the charges and dipoles defining the response of the nanostructure as modified by the presence of the solvent are obtained. This allows the simulation of the optical response of plasmonic substrates with arbitrary shape and chemical composition embedded in a generic solvent or solvent mixture (See Sec. S1.2 in the SI).

3 Computational Protocol

Refer to caption
Figure 2: Graphical scheme of the computational protocol employed to compute ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption spectra of colloidal plasmonic NPs (see also Secs. S2.1-S2.4 in the SI).

In this work, we apply ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ to the calculation of the absorption cross-section of plasmonic nanoparticles in solution. To reproduce the experimental PRF shift induced by solvent effects, the dynamic nature of the solvation phenomena needs to be properly described. To this end, we adapt the protocol designed for molecular systems in solution21 to the specific case of colloidal plasmonic NPs.

The protocol can be divided into four main steps (see Fig. 2):

  1. (i)

    Geometry generation: The geometry of isolated plasmonic NPs is generated by using an in-house code that employs the Atomic Simulation Environment (ASE) Python module v. 3.17 49 (see Fig. 2a and Sec. S2.1 in the SI)

  2. (ii)

    Molecular Dynamics: The nanostructure is solvated. To sample the NP-solvent phase-space, we perform a classical Molecular Dynamics (MD) simulation of the solvated colloidal system by using the GROMACS software package (version 2020.4) and a suitable force-field50 (see Fig. 2b and Sec. S2.2 in the SI).

  3. (iii)

    Extraction of structures: From the MD trajectory, we extract 25 uncorrelated representative structures of the whole system. The number of structures ensures convergence of the spectral signal (see Sec. S2.4.2 in the SI). For each structure, we retain all solvent molecules that are at most 15 Å  from the NP surface, resulting in a spherical droplet (see Fig. 2c) and Sec. S2.4.1 in the SI).

  4. (iv)

    ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ spectral calculations: For each spherical droplet, absorption cross sections are computed at the ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ level. The overall spectroscopic response is then recovered as the average over all structures (see Fig. 2d and Sec. S2.3 in the SI). All ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ calculations are performed by employing a stand-alone Fortran 95 package.

4 Results and Discussion

In this section, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is applied to compute the absorption properties of noble metal NPs in solution. First, the model is validated by reproducing reference data. Then, it is employed to study solvent effects on realistic homogeneous and bimetallic NPs dissolved in a pure solvent or a solvent mixture, showcasing the potential and flexibility of the approach.

4.1 Model Validation

The ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approach is validated against vacuo-to-water PRF shifts of a small silver spherical-like cluster composed of 164 atoms (Ag164). ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ values are compared to polarizable Time-Dependent Density Functional Tight Binding/Fluctuating Charges (TD-DFTB/FQ) calculations.35 The advantageous computational scaling of TD-DFTB/FQ, which is reached through the approximation of two-electron interactions, makes it capable of handling larger systems than Time-Dependent Density Functional Theory-based methods (TD-DFT) while preserving accuracy.35 The initial geometry of Ag164 is taken from Ref. 51: from that structure, MD simulations in aqueous solution are run, a single random snapshot is extracted and cut in a spherical droplet containing water (WAT) molecules within 10 Å from the surface of the NP, resulting in a total number of 807 water molecules. Such distance is chosen to account for the most relevant NP-water interactions (See Sec. S2.4.1 in the SI). The absorption spectrum of Ag164 in vacuo (i.e. by removing all water molecules from the snapshot) and in aqueous solution (Ag164/WAT807) are computed by using ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ and DFTB, and ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ and DFTB/FQ, respectively (see Sec. S2.5 in the SI for more details on the DFTB/FQ calculations). To evaluate quantum effects at the NP-solvent interface, we also performed DFTB/DFTBWAT/FQ calculations, where the first solvation shell is treated at the DFTB level of theory (109 water molecules), whereas the remaining 698 WAT molecules are described at the FQ level. ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ parameters are recovered from Ref. 16. The FQ WAT molecules are modeled by using the parameters reported in Ref. 18 (WAT1, see Tab. S1 in Sec. S2.3 in the SI). Note that additional calculations were performed by using the FQ parameters reported in Refs. 52, 53, 54 (see Sec. S3 in the SI).

In Fig. 3, we report the PRF shifts (ΔEPRF=EVACEWATΔsuperscript𝐸PRFsuperscript𝐸𝑉𝐴𝐶superscript𝐸𝑊𝐴𝑇\Delta E^{\text{PRF}}=E^{VAC}-E^{WAT}roman_Δ italic_E start_POSTSUPERSCRIPT PRF end_POSTSUPERSCRIPT = italic_E start_POSTSUPERSCRIPT italic_V italic_A italic_C end_POSTSUPERSCRIPT - italic_E start_POSTSUPERSCRIPT italic_W italic_A italic_T end_POSTSUPERSCRIPT) in eV calculated at the DFTB/FQ level (panel A), ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ (panel B), and DFTB/DFTBWAT/FQ (panel C).

Refer to caption
Figure 3: Vacuo-to-water PRF shifts (in eV) computed by using DFTB/FQ, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, and DFT/DFTBWAT/FQ levels of theory. A graphical depiction of the structure used for each calculation is presented and the elements are colored according to the level of theory used (orange-DFTB, gray-ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ, blue-FQ). In the green inset a zoom of the DFTB/DFTBWAT/FQ structure is reported to highlight the presence of water molecules treated at the DFTB level of theory (orange).

All methods predict a very similar vacuo-to-water PRF redshift. The sign of the shift is expected due to the increase of the refractive index of the NP surrounding medium, as also supported by experimental observations.55, 56 Notably, the hybrid DFTB/FQ and the fully classical ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ approaches predict the same solvatochromic shift (64 meV). This is particularly remarkable because the dimension of the studied NP falls within the quantum size region, where quantum effects are expected to play a major role. Our results highlight the accuracy of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ in describing such structures.57 When the first solvation shell of water molecules is described at the DFTB level (DFTB/DFTBWAT/FQ), the predicted solvatochromic shift decreases by similar-to\sim 10 meV (53 meV). Such reduction can be attributed to the inclusion of purely quantum NP-solvent interactions, mainly related to Pauli repulsion effects, 58, 59, 35 which are not taken into account by both DFTB/FQ and ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ. Remarkably, the computational cost associated with ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is negligible as compared to DFTB-based methods (see Tab. S4 in the SI). Indeed, the favorable computational scaling of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ provides a substantial 99.4 % speed-up of the calculation with respect to DFTB/FQ, thus representing an effective, reliable, and cost-effective alternative to state-of-the-art ab initio methods.

4.2 Homogeneous colloidal NPs

The favorable computational scaling of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ opens up to computing the plasmonic response of large NPs in solution. We first consider homogeneous silver and gold spherical NPs (diameter = 5 nm, 3851 atoms) in aqueous solution, as modeled by using the WAT1 parameterization18. The average number of WAT molecules in the FQ region is 6464 (19392 atoms). Computed absorption cross-sections both in vacuo (VAC) and in water (WAT) are plotted in Fig. 4A-B, respectively, together with the corresponding vacuo-to-water PRF shifts (ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ, in nm). The spectra of both systems are characterized by a main plasmonic band centered at about 360 nm (Ag NPs) and 570 nm (Au NPs). Au peaks are broader than Ag bands, in agreement with previous observations.60, 61 In the studied region (300-800 nm), the spectra of Au systems are also characterized by the presence of a broad and intense band associated with interband absorption.16 When dissolved in solution, the main spectral features are maintained. However, we note that the plasmonic peak increases in intensity and slightly redshifts, in line with experimental data.9, 62, 55, 56 For Ag, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ predicts a red-shift of the plasmonic peak larger than Au NPs, again in agreement with experiments.9, 62, 55, 56 Such shift is due to the refractive index sensitivity (RIS) of the LSP, a quantity that is commonly exploited in plasmonic colorimetric sensors.63, 29, 64, 30 RIS is generally expressed as ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ (in nm) over refractive index unit - ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ(nm)/RIU.

Refer to caption
Figure 4: ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption spectra of (A) Ag3851 and (B) Au3851 in vacuo (VAC) and water (WAT1). ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ is the vacuo-to-water solvatochromic shift in nm. All spectra are normalized to the corresponding maximum in vacuo.

The computed ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ RIS values are 15.0 nm/RIU for Ag and 7.9 nm/RIU for Au, while the experimentally measured RIS are 120-160 nm/RIU (Ag) and 70 nm/RIU (Au).63, 29, 64, 30 Therefore, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absolute RIS values are systematically lower than experimental ones for Ag and Au NPs, however ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ can nicely reproduce the experimental Ag/Au ratio (similar-to\sim 2). The discrepancy can be related to the FQ parameters employed for modeling water, which have not been specifically tuned to describe solvated NPs.18 Nevertheless, although experimental absolute RIS values are underestimated by a factor of 10, the capability of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ to match the Ag/Au sensitivity ratio is a remarkable feature of the model, which can be used to rationalize the optical behavior of plasmonic nanostructures with high RIS, with potential applications in sensor design.

4.3 Au@Ag core-shell colloidal NPs

By taking advantage of the favorable computational scaling of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, its atomistic nature, and its capability of correctly reproducing Ag/Au RI sensitivity ratio, the model is challenged to optimize the sensitivity of Au@Ag core-shell spherical NPs. In fact, colorimetric sensors exploiting the PRF shift of noble metals upon change of the local refractive index (RI) are widely used in biosensing 7, 8, 27, 28, 29, 30, 11, 31, 32, 33, 34. Among the various substrates used for this technology, gold-silver core-shell (Au@Ag) NPs are the most employed 65, 66, 67, 68, 69, 70 due to their high RIS.11, 70 By studying such structures, we aim to showcase the potentialities of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, open up to colorimetric LSP sensor design.

We first consider spherical Au@Ag core-shell NPs (diameter = 5 nm, 3851 atoms) characterized by an Au core of increasing size (dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT=2, 3, and 4 nm – see Fig. 5A,B,C), in aqueous solution. To sample the NP-solvent phase space, we perform a classical MD simulation of a single Ag NP in the aqueous solution. We then construct the Ag@Au core-shell bimetallic NP by properly substituting the metal atoms after the MD snapshot’s extraction, which is a physically consistent procedure due to the almost equal lattice constants of the two metals.71 On average, 6464 water molecules are considered in the FQ region and modeled employing the WAT1 parameterization.18

Refer to caption
Figure 5: ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption spectra of core-shell Au@Ag spherical NPs in vacuo (VAC) and in aqueous solution (WAT1) as a function of the diameter of the Au core (dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT: (A) 2.0 nm, (B) 3.0 nm and (C) 4.0 nm). The solvatochromic shifts of “Ag” (ΔλAgΔsubscript𝜆𝐴𝑔\Delta\lambda_{Ag}roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_g end_POSTSUBSCRIPT) and “Au” (ΔλAuΔsubscript𝜆𝐴𝑢\Delta\lambda_{Au}roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT) peaks are reported in nm. All spectra are normalized to the maximum in vacuo.

In Fig. 5, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption spectra of Au@Ag core-shell NPs in vacuo (black) and water (blue) are reported. For all structures, absorption spectra in vacuo and solution feature two main bands at short (similar-to\sim 350 nm) and long wavelengths (> 400 nm), in agreement with previous theoretical and experimental observations.11, 12, 72, 73 Such peaks are located in the spectral region of the plasmon bands of pure Ag and Au spheres (see Fig. 4A,B). For this reason, we call them “Ag” and “Au” peaks, respectively. Their position and intensity strongly depend on the diameter of the Au core both in the gas phase and solution. In particular, by first focusing on vacuo results, for Au-core thickness of 1 nm (Fig. 5A), the “Ag” peak, located around 346 nm, dominates the computed spectrum, while the “Au” band appears as a shoulder. By increasing the Au-core size, the “Ag” peak slightly blueshifts of about 10 nm while decreasing in intensity. On the contrary, the “Au” peak largely redshifts of about 150 nm, and significantly increases in intensity, becoming dominant for Au-core diameters of 4 nm (1 nm Ag-shell thickness). Including the solvent redshifts both “Ag” and “Au” peaks and increases their intensity. The shift differs for each peak and each Au-core diameter. In particular, all “Ag” peaks show lower solvatochromic shifts as compared to pure Ag spheres (Δλ=Δ𝜆absent\Delta\lambda=roman_Δ italic_λ = 5.0 nm see Fig. 4A), and increasing the Au-core size reduces the redshift of the “Ag” peak from ΔλAg=Δsubscript𝜆𝐴𝑔absent\Delta\lambda_{Ag}=roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_g end_POSTSUBSCRIPT = 3.8 nm (dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT = 2 nm) to ΔλAg=Δsubscript𝜆𝐴𝑔absent\Delta\lambda_{Ag}=roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_g end_POSTSUBSCRIPT = 0.4 nm (dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT = 4 nm). The shift of the ”Au” peak is generally larger (e.g. ΔλAu=Δsubscript𝜆𝐴𝑢absent\Delta\lambda_{Au}=roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT = 4.4 nm for dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT = 2 nm) than that of a pure Au sphere (Δλ=Δ𝜆absent\Delta\lambda=roman_Δ italic_λ = 2.6 nm, see Fig. 4B). This suggests that for small NPs (dsimilar-to𝑑absentd\simitalic_d ∼ 5 nm), the “Au” peak RIS can be increased by coating the NP with an Ag layer. Remarkably, our findings align with previous theoretical and experimental observations.11, 67, 70.

To showcase more potentialities of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, we now move to study the effect of alloying the Ag shell on the refractive index sensitivity, for which an atomistic picture is essential.17 Note that, to the best of our knowledge, the use of atomistic modeling in this field has received only a little attention.12, 74, 11

We consider a substantially more complex system composed of a spherical Au@Ag core-shell bimetallic NP (see Fig. 6) of 5 nm of diameter (3851 atoms) with an Au core of diameter (dAusubscript𝑑𝐴𝑢d_{Au}italic_d start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT) of 3 nm, and featuring a 50% Ag 50%Au random bimetallic layer of 1 nm. Such NP is solvated in aqueous solution, exploiting the same procedure discussed above. Fig. 6, shows computed ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption cross-section (σabssuperscript𝜎abs\sigma^{\mathrm{abs}}italic_σ start_POSTSUPERSCRIPT roman_abs end_POSTSUPERSCRIPT) both in vacuo and water. The computed spectrum substantially deviates from the corresponding perfect core-shell system (Fig. 5B). In fact, the “Au” peak is shifted at higher wavelengths (similar-to\sim 600 nm) and dominates the spectrum, while the ”Ag” peak disappears in a dim spectral feature between 350 nm and 500 nm. Remarkably, the computed PRF shift of the gold peak (ΔλAuΔsubscript𝜆𝐴𝑢\Delta\lambda_{Au}roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT) is 8.4 nm, almost double the original Au@Ag core-shell structure (ΔλAu=Δsubscript𝜆𝐴𝑢absent\Delta\lambda_{Au}=roman_Δ italic_λ start_POSTSUBSCRIPT italic_A italic_u end_POSTSUBSCRIPT = 4.4 nm, see Fig. 5B), and more than three times than a pure Au sphere of the same dimensions (Δλ=Δ𝜆absent\Delta\lambda=roman_Δ italic_λ = 2.6 nm, see Fig. 4). Remarkably, the same trend is also obtained for the computed ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ RIS sensitivity, which reaches 25.5 nm/RIU. The obtained results thus suggest that the alloying of the Ag shell of an Au@Ag core shell can potentially improve the refractive index sensitivity of the system by almost a factor of 2.

Refer to caption
Figure 6: ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption cross section (σabssuperscript𝜎𝑎𝑏𝑠\sigma^{abs}italic_σ start_POSTSUPERSCRIPT italic_a italic_b italic_s end_POSTSUPERSCRIPT) of core-shell Au@Ag spherical NP with an Au core diameter of 3 nm and featuring an alloyed (50% Au/Ag) external layer of 1 nm (see right panel) in vacuo (VAC) and aqueous solution (WAT1). ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ indicates the solvatochromic shift (in nm). All spectra are normalized to the maximum in vacuo.

To finally showcase the flexibility of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, we solvate the NP in a 1:1 water-ethanol mixture (WAT-ETH). The even more complex chemical nature of this system enforces the need for a fully atomistic model to simulate its optical properties. To the best of our knowledge, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is the first fully atomistic model capable of calculating the response of a generic multimetallic plasmonic NP embedded in a multicomponent solvent.

Refer to caption
Figure 7: ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ absorption cross section (σabssuperscript𝜎𝑎𝑏𝑠\sigma^{abs}italic_σ start_POSTSUPERSCRIPT italic_a italic_b italic_s end_POSTSUPERSCRIPT) of core-shell Au@Ag spherical NP with an Au core diameter of 3 nm and featuring an alloyed (50% Au/Ag) external layer of 1 nm (see right panel) in vacuo (VAC) and a 1:1 water-ethanol mixture (WAT-ETH). The vacuo-to-mixture solvatochromic shift is given in nm. All spectra are normalized to the maximum in vacuo.

To sample the NP-solvent phase space, we perform a classical MD simulation of a single Ag NP in the WAT-ETH mixture. We then construct the Ag@Au core-shell bimetallic NP by properly substituting the metal atoms after the MD snapshot extraction. The FQ parameterization employed for the water molecules is WAT1, while for ethanol we exploit the parameters proposed in Ref. 54 (see Tab.S1 in the SI). On average 1800 ETH and 1218 WAT molecules are included in the spherically-shaped snapshot (i.e. 19854 atoms in total).

In Fig. 7, the absorption spectrum of the complex bimetallic Au/Ag NP both in vacuo (VAC) and solvated in the 1:1 water-ethanol mixture (WAT-ETH) is graphically depicted, together with the corresponding computed PRF shift (ΔλΔ𝜆\Delta\lambdaroman_Δ italic_λ in nm). Notably, absorption spectra both in vacuo and in solution present a sharp peak around 600 nm which can be assigned to the Au plasmonic dipolar band, and a shoulder at about 350-400 nm associated with the Ag plasmonic absorption. Solvent effects provided by the WAT-ETH mixture lead to a huge enhancement of the peak absorption (almost twice that in vacuo), which is also red-shifted by 16 nm. Considering a refractive index for the WAT-ETH mixture of 1.36,75 a computed RIS of about 44.0 nm/RIU is obtained. Such a value is almost twice that obtained in pure water, highlighting a non-trivial RIS dependence on the solvent composition.

5 Summary and Conclusions

We have presented a novel fully atomistic multiscale classical model, named ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, which is capable of simulating the optical properties of real-size plasmonic colloidal nanoparticles (NPs) of a generic chemical nature. ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is a multiscale model based on the mutual electrostatic interaction between the solvent and the NP. More specifically, the interaction of the plasmonic substrate with the external optical field is modeled employing ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ,16, 17 where each atom is endowed with a frequency-dependent charge and dipole, modeling intraband and interband plasmon decaying mechanisms, respectively. The solvent environment is considered transparent to the optical radiation and its instantaneous polarization is modeled through the polarizable FQ force field.18, 19, 20

ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ has been challenged to reproduce reference TD-DFTB/FQ values, showing an almost perfect match with the TD-DFTB/FQ vacuo-to-water plasmon resonance frequency (PRF) shift of a small silver cluster (Ag164), with a substantial 99.4 % speed-up of the calculation. Then, we have showcased the capabilities of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ by simulating the optical properties of real-size homogeneous Ag and Au spherical NPs (similar-to\sim 5 nm of diameter, 3851 metal atoms with 6464 water molecules), highlighting how the computed ratio between the refractive index sensitivities of Au and Ag NPs matches the experiments. Remarkably, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ can also be used to study the sensitivity of colorimetric LSP sensors, as demonstrated by the chemical substitution of Au atoms in Au@Ag core-shell NPs, which can potentially enhance the sensitivity by a factor 2 similar-to\sim 3. Finally, the flexibility of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is validated by simulating the absorption spectrum of a bimetallic Ag/Au NP solvated in a 1:1 water-ethanol mixture. Remarkably, the model can be applied to any solvent or solvent mixtures, including green solvents,76 pending a reliable parametrization of the FQ force field.54

In conclusion, ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ is the first fully atomistic classical model capable of providing a platform for the calculation of LSP shifts of plasmonic NPs with the accuracy of ab-initio methodologies for systems in the quantum confinement size region, but with a computational cost that consents its application to realistic-sized colloidal NPs. Such a development can potentially pave the way for future in-silico rational design of colorimetric sensors.

Acknowledgments

We gratefully acknowledge the Center for High-Performance Computing (CHPC) at SNS for providing the computational infrastructure.

Supporting Information

Details on ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ, computational details, convergence of ω𝜔\omegaitalic_ωFQFμ𝜇\muitalic_μ/FQ spectra as a function of the radius of the solvent droplet and number of snapshots, MD analysis for spherical NPs (diameter = 5 nm) in solution.

References

  • Mayer and Hafner 2011 Mayer, K. M.; Hafner, J. H. Localized surface plasmon resonance sensors. Chem. Rev. 2011, 111, 3828–3857.
  • McNamara and Tofail 2017 McNamara, K.; Tofail, S. A. Nanoparticles in biomedical applications. Adv. Phys. X 2017, 2, 54–88.
  • Dutta et al. 2020 Dutta, A.; Medda, A.; Patra, A. Recent advances and perspectives on colloidal semiconductor nanoplatelets for optoelectronic applications. J. Phys. Chem. C 2020, 125, 20–30.
  • Hu et al. 2010 Hu, X.; Li, G.; Yu, J. C. Design, fabrication, and modification of nanostructured semiconductor materials for environmental and energy applications. Langmuir 2010, 26, 3031–3039.
  • Yang et al. 2016 Yang, P.; Zheng, J.; Xu, Y.; Zhang, Q.; Jiang, L. Colloidal synthesis and applications of plasmonic metal nanoparticles. Adv. Mater. 2016, 28, 10508–10517.
  • Chen et al. 2009 Chen, H.; Shao, L.; Woo, K. C.; Ming, T.; Lin, H.-Q.; Wang, J. Shape-dependent refractive index sensitivities of gold nanocrystals with the same plasmon resonance wavelength. J. Phys. Chem. C 2009, 113, 17691–17697.
  • Lee et al. 2013 Lee, J.-H.; Kim, B.-C.; Oh, B.-K.; Choi, J.-W. Highly sensitive localized surface plasmon resonance immunosensor for label-free detection of HIV-1. Nanomedicine 2013, 9, 1018–1026.
  • Kim et al. 2018 Kim, J.; Oh, S. Y.; Shukla, S.; Hong, S. B.; Heo, N. S.; Bajpai, V. K.; Chun, H. S.; Jo, C.-H.; Choi, B. G.; Huh, Y. S., et al. Heteroassembled gold nanoparticles with sandwich-immunoassay LSPR chip format for rapid and sensitive detection of hepatitis B virus surface antigen (HBsAg). Biosens. Bioelectron. 2018, 107, 118–122.
  • Chen et al. 2008 Chen, H.; Kou, X.; Yang, Z.; Ni, W.; Wang, J. Shape-and size-dependent refractive index sensitivity of gold nanoparticles. Langmuir 2008, 24, 5233–5237.
  • Foerster et al. 2017 Foerster, B.; Joplin, A.; Kaefer, K.; Celiksoy, S.; Link, S.; Sönnichsen, C. Chemical interface dam** depends on electrons reaching the surface. ACS nano 2017, 11, 2886–2893.
  • Steinbrück et al. 2011 Steinbrück, A.; Stranik, O.; Csaki, A.; Fritzsche, W. Sensoric potential of gold–silver core–shell nanoparticles. Anal. Bioanal. Chem. 2011, 401, 1241–1249.
  • Szántó et al. 2021 Szántó, G.; Csarnovics, I.; Bonyár, A. Numerical investigation of the refractive index sensitivity of Au/Ag core-shell nanostructures for sensing applications. Sens. Bio-Sens. Res. 2021, 32, 100414.
  • Hu et al. 2016 Hu, Y.; Zhang, A.-Q.; Li, H.-J.; Qian, D.-J.; Chen, M. Synthesis, study, and discrete dipole approximation simulation of Ag-Au bimetallic nanostructures. Nanoscale Res. Lett. 2016, 11, 1–9.
  • Hsiao et al. 2015 Hsiao, A.; Gartia, M. R.; Chang, T.-W.; Wang, X.; Khumwan, P.; Liu, G. L. Colorimetric plasmon resonance microfluidics on nanohole array sensors. Sens. Bio-Sens. Res. 2015, 5, 24–32.
  • Yamada 2022 Yamada, A. Computational Analyses of Plasmonics of a Silver Nanoparticle in a Vacuum and in a Water Solution by Classical Electronic and Molecular Dynamics Simulations. J. Phys. Chem. A 2022, 126, 4762–4771.
  • Giovannini et al. 2022 Giovannini, T.; Bonatti, L.; Lafiosca, P.; Nicoli, L.; Castagnola, M.; Illobre, P. G.; Corni, S.; Cappelli, C. Do We Really Need Quantum Mechanics to Describe Plasmonic Properties of Metal Nanostructures? ACS Photonics 2022,
  • Nicoli et al. 2023 Nicoli, L.; Lafiosca, P.; Grobas Illobre, P.; Bonatti, L.; Giovannini, T.; Cappelli, C. Fully atomistic modeling of plasmonic bimetallic nanoparticles: nanoalloys and core-shell systems. Front. Photon. 2023, 4, 1199598.
  • Rick et al. 1994 Rick, S. W.; Stuart, S. J.; Berne, B. J. Dynamical fluctuating charge force fields: Application to liquid water. J. Chem. Phys. 1994, 101, 6141–6156.
  • Rick et al. 1995 Rick, S. W.; Stuart, S. J.; Bader, J. S.; Berne, B. Fluctuating charge force fields for aqueous solutions. J. Mol. Liq. 1995, 65, 31–40.
  • Rick and Berne 1996 Rick, S. W.; Berne, B. Dynamical fluctuating charge force fields: the aqueous solvation of amides. J. Am. Chem. Soc. 1996, 118, 672–679.
  • Giovannini et al. 2020 Giovannini, T.; Egidi, F.; Cappelli, C. Molecular spectroscopy of aqueous solutions: a theoretical perspective. Chem. Soc. Rev. 2020, 49, 5664–5677.
  • Giovannini et al. 2020 Giovannini, T.; Egidi, F.; Cappelli, C. Theory and algorithms for chiroptical properties and spectroscopies of aqueous systems. Phys. Chem. Chem. Phys. 2020, 22, 22864–22879.
  • Gómez et al. 2023 Gómez, S.; Giovannini, T.; Cappelli, C. Multiple facets of modeling electronic absorption spectra of systems in solution. ACS Physical Chemistry Au 2023, 3, 1–16.
  • Giovannini and Cappelli 2023 Giovannini, T.; Cappelli, C. Continuum vs. atomistic approaches to computational spectroscopy of solvated systems. Chem. Commun. 2023, 59, 5644–5660.
  • Mennucci and Corni 2019 Mennucci, B.; Corni, S. Multiscale modelling of photoinduced processes in composite systems. Nat. Rev. Chem. 2019, 3, 315–330.
  • Xia and Halas 2005 Xia, Y.; Halas, N. J. Shape-controlled synthesis and surface plasmonic properties of metallic nanostructures. MRS Bull. 2005, 30, 338–348.
  • Loiseau et al. 2019 Loiseau, A.; Zhang, L.; Hu, D.; Salmain, M.; Mazouzi, Y.; Flack, R.; Liedberg, B.; Boujday, S. Core–shell gold/silver nanoparticles for localized surface plasmon resonance-based naked-eye toxin biosensing. ACS Appl. Mater. Interfaces 2019, 11, 46462–46471.
  • Piliarik et al. 2012 Piliarik, M.; Šípová, H.; Kvasnička, P.; Galler, N.; Krenn, J. R.; Homola, J. High-resolution biosensor based on localized surface plasmons. Opt. Express 2012, 20, 672–680.
  • Mock et al. 2003 Mock, J. J.; Smith, D. R.; Schultz, S. Local refractive index dependence of plasmon resonance spectra from individual nanoparticles. Nano lett. 2003, 3, 485–491.
  • Underwood and Mulvaney 1994 Underwood, S.; Mulvaney, P. Effect of the solution refractive index on the color of gold colloids. Langmuir 1994, 10, 3427–3430.
  • Rycenga et al. 2011 Rycenga, M.; Cobley, C. M.; Zeng, J.; Li, W.; Moran, C. H.; Zhang, Q.; Qin, D.; Xia, Y. Controlling the synthesis and assembly of silver nanostructures for plasmonic applications. Chem. Rev. 2011, 111, 3669–3712.
  • Haes and Van Duyne 2004 Haes, A. J.; Van Duyne, R. P. A unified view of propagating and localized surface plasmon resonance biosensors. Anal. Bioanal. Chem. 2004, 379, 920–930.
  • Anker et al. 2008 Anker, J. N.; Hall, W. P.; Lyandres, O.; Shah, N. C.; Zhao, J.; Van Duyne, R. P. Biosensing with plasmonic nanosensors. Nat. Mater 2008, 7, 442–453.
  • Stewart et al. 2008 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, 494–521.
  • Lafiosca et al. 2022 Lafiosca, P.; Gómez, S.; Giovannini, T.; Cappelli, C. Absorption properties of large complex molecular systems: the DFTB/fluctuating charge approach. J. Chem. Theory Comput. 2022, 18, 1765–1779.
  • Giovannini et al. 2019 Giovannini, T.; Rosa, M.; Corni, S.; Cappelli, C. A classical picture of subnanometer junctions: an atomistic Drude approach to nanoplasmonics. Nanoscale 2019, 11, 6004–6015.
  • Giovannini et al. 2020 Giovannini, T.; Bonatti, L.; Polini, M.; Cappelli, C. Graphene plasmonics: Fully atomistic approach for realistic structures. J. Phys. Chem. Lett. 2020, 11, 7595–7602.
  • Bonatti et al. 2022 Bonatti, L.; Nicoli, L.; Giovannini, T.; Cappelli, C. In silico design of graphene plasmonic hot-spots. Nanoscale Adv. 2022, 4, 2294–2302.
  • Maier et al. 2007 Maier, S. A., et al. Plasmonics: fundamentals and applications; Springer, 2007; Vol. 1.
  • Giovannini et al. 2019 Giovannini, T.; Puglisi, A.; Ambrosetti, M.; Cappelli, C. Polarizable QM/MM approach with fluctuating charges and fluctuating dipoles: the QM/FQFμ𝜇\muitalic_μ model. J. Chem. Theory Comput. 2019, 15, 2233–2245.
  • Litjens et al. 1999 Litjens, R. A.; Quickenden, T. I.; Freeman, C. G. Visible and near-ultraviolet absorption spectrum of liquid water. Appl. Opt. 1999, 38, 1216–1223.
  • Sani and Dell’Oro 2016 Sani, E.; Dell’Oro, A. Spectral optical constants of ethanol and isopropanol from ultraviolet to far infrared. Opt. Mater. 2016, 60, 137–141.
  • Nicoli et al. 2022 Nicoli, L.; Giovannini, T.; Cappelli, C. Assessing the quality of QM/MM approaches to describe vacuo-to-water solvatochromic shifts. J. Chem. Phys. 2022, 157.
  • Mortier et al. 1985 Mortier, W. J.; Van Genechten, K.; Gasteiger, J. Electronegativity equalization: application and parametrization. J. Am. Chem. Soc. 1985, 107, 829–835.
  • Sanderson 1951 Sanderson, R. An interpretation of bond lengths and a classification of bonds. Science 1951, 114, 670–672.
  • Parr 1983 Parr, R. G. Density functional theory. Annu. Rev. Phys. Chem. 1983, 34, 631–656.
  • Geerlings et al. 2003 Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual density functional theory. Chem. Rev. 2003, 103, 1793–1874.
  • Cappelli 2016 Cappelli, C. Integrated QM/polarizable MM/continuum approaches to model chiroptical properties of strongly interacting solute–solvent systems. Int. J. Quantum Chem. 2016, 116, 1532–1542.
  • Larsen et al. 2017 Larsen, A. H. et al. The atomic simulation environment—a Python library for working with atoms. J. Phys.: Condens. Matter 2017, 29, 273002.
  • Abraham et al. 2015 Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1, 19–25.
  • Liu et al. 2020 Liu, Z.; Alkan, F.; Aikens, C. M. TD-DFTB study of optical properties of silver nanoparticle homodimers and heterodimers. J. Chem. Phys. 2020, 153.
  • Carnimeo et al. 2015 Carnimeo, I.; Cappelli, C.; Barone, V. Analytical gradients for MP 2, double hybrid functionals, and TD-DFT with polarizable embedding described by fluctuating charges. J. Comput. Chem. 2015, 36, 2271–2290.
  • Giovannini et al. 2019 Giovannini, T.; Lafiosca, P.; Chandramouli, B.; Barone, V.; Cappelli, C. Effective yet reliable computation of hyperfine coupling constants in solution by a QM/MM approach: Interplay between electrostatics and non-electrostatic effects. J. Chem. Phys. 2019, 150, 124102.
  • Ambrosetti et al. 2021 Ambrosetti, M.; Skoko, S.; Giovannini, T.; Cappelli, C. Quantum mechanics/fluctuating charge protocol to compute solvatochromic shifts. Journal of Chemical Theory and Computation 2021, 17, 7146–7156.
  • Zhang et al. 2016 Zhang, C.; Sun, L.-D.; Yan, C.-H. Noble metal plasmonic nanostructure related chromisms. Inorganic Chemistry Frontiers 2016, 3, 203–217.
  • Lee et al. 2011 Lee, Y. H.; Chen, H.; Xu, Q.-H.; Wang, J. Refractive index sensitivities of noble metal nanocrystals: the effects of multipolar plasmon resonances and the metal type. J. Phys. Chem. C 2011, 115, 7997–8004.
  • Liebsch 1993 Liebsch, A. Surface-plasmon dispersion and size dependence of Mie resonance: silver versus simple metals. Phys. Rev. B 1993, 48, 11317.
  • Giovannini et al. 2019 Giovannini, T.; Ambrosetti, M.; Cappelli, C. Quantum confinement effects on solvatochromic shifts of molecular solutes. J. Phys. Chem. Lett. 2019, 10, 5823–5829.
  • Amovilli and Floris 2020 Amovilli, C.; Floris, F. M. On the effect of solute-solvent Pauli repulsion on n→ π𝜋\piitalic_π* transition for acrolein in water solution. Phys. Chem. Liq. 2020, 58, 281–289.
  • Hottin et al. 2013 Hottin, J.; Wijaya, E.; Hay, L.; Maricot, S.; Bouazaoui, M.; Vilcot, J.-P. Comparison of gold and silver/gold bimetallic surface for highly sensitive near-infrared SPR sensor at 1550 nm. Plasmonics 2013, 8, 619–624.
  • Dengler et al. 2012 Dengler, S.; Kübel, C.; Schwenke, A.; Ritt, G.; Eberle, B. Near-and off-resonant optical limiting properties of gold–silver alloy nanoparticles for intense nanosecond laser pulses. J. Opt. 2012, 14, 075203.
  • Ghosh et al. 2004 Ghosh, S. K.; Nath, S.; Kundu, S.; Esumi, K.; Pal, T. Solvent and ligand effects on the localized surface plasmon resonance (LSPR) of gold colloids. J. Phys. Chem. B 2004, 108, 13963–13971.
  • Malinsky et al. 2001 Malinsky, M. D.; Kelly, K. L.; Schatz, G. C.; Van Duyne, R. P. Chain length dependence and sensing capabilities of the localized surface plasmon resonance of silver nanoparticles chemically modified with alkanethiol self-assembled monolayers. J. Am. Chem. Soc. 2001, 123, 1471–1482.
  • Nath and Chilkoti 2004 Nath, N.; Chilkoti, A. Label free colorimetric biosensing using nanoparticles. J. Fluoresc. 2004, 14, 377–389.
  • Mao et al. 2017 Mao, K.; Yang, Z.; Li, J.; Zhou, X.; Li, X.; Hu, J. A novel colorimetric biosensor based on non-aggregated Au@ Ag core–shell nanoparticles for methamphetamine and cocaine detection. Talanta 2017, 175, 338–346.
  • Dong et al. 2013 Dong, P.; Wu, Y.; Guo, W.; Di, J. Plasmonic biosensor based on triangular Au/Ag and Au/Ag/Au core/shell nanoprisms onto indium tin oxide glass. Plasmonics 2013, 8, 1577–1583.
  • Sun et al. 2014 Sun, L.; Li, Q.; Tang, W.; Di, J.; Wu, Y. The use of gold-silver core-shell nanorods self-assembled on a glass substrate can substantially improve the performance of plasmonic affinity biosensors. Microchim. Acta 2014, 181, 1991–1997.
  • Hao et al. 2014 Hao, J.; Xiong, B.; Cheng, X.; He, Y.; Yeung, E. S. High-throughput sulfide sensing with colorimetric analysis of single Au–Ag core–shell nanoparticles. Anal. Chem. 2014, 86, 4663–4667.
  • Guo et al. 2016 Guo, Y.; Wu, J.; Li, J.; Ju, H. A plasmonic colorimetric strategy for biosensing through enzyme guided growth of silver nanoparticles on gold nanostars. Biosens. Bioelectron. 2016, 78, 267–273.
  • Fu et al. 2012 Fu, Q.; Zhang, D.; Yi, M.; Wang, X.; Chen, Y.; Wang, P.; Ming, H. Effect of shell thickness on a Au–Ag core–shell nanorods-based plasmonic nano-sensor. J. Opt. 2012, 14, 085001.
  • Haynes 2014 Haynes, W. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, 2014.
  • Chen et al. 2012 Chen, Y.; Wu, H.; Li, Z.; Wang, P.; Yang, L.; Fang, Y. The study of surface plasmon in Au/Ag core/shell compound nanoparticles. Plasmonics 2012, 7, 509–513.
  • Peña-Rodríguez and Pal 2011 Peña-Rodríguez, O.; Pal, U. Au@ Ag core–shell nanoparticles: efficient all-plasmonic Fano-resonance generators. Nanoscale 2011, 3, 3609–3612.
  • Ma et al. 2015 Ma, Y.-W.; Zhang, L.-H.; Wu, Z.-W.; Yi, M.-F.; Zhang, J.; Jian, G.-S. The study of tunable local surface plasmon resonances on Au-Ag and Ag-Au core-shell alloy nanostructure particles with DDA method. Plasmonics 2015, 10, 1791–1800.
  • Oelke and Arnold 1936 Oelke, W.; Arnold, R. The Refractive Indices of Alcohol-Water Mixture at 25º C. Proceedings of the Iowa Academy of Science. 1936; pp 175–176.
  • Atilhan and Aparicio 2018 Atilhan, M.; Aparicio, S. Molecular dynamics simulations of metal nanoparticles in deep eutectic solvents. J. Phys. Chem. C 2018, 122, 18029–18039.