Skip to main content
The Journal of Chemical Physics logoLink to The Journal of Chemical Physics
. 2017 May 18;146(19):194106. doi: 10.1063/1.4983391

Gauge invariant theory for super high resolution Raman images

Sai Duan 1, Guangjun Tian 2, Zhen Xie 1, Yi Luo 1,3,1,3,a)
PMCID: PMC5436987  PMID: 28527435

Abstract

The use of a highly localized plasmonic field has enabled us to achieve sub-nanometer resolution of Raman images for single molecules. The inhomogeneous spatial distribution of plasmonic field has become an important factor that controls the interaction between the light and the molecule. We present here a gauge invariant interaction Hamiltonian (GIIH) to take into account the non-uniformity of the electromagnetic field distribution in the non-relativistic regime. The theory has been implemented for both resonant and nonresonant Raman processes within the sum-over-state framework. It removes the gauge origin dependence in the phenomenologically modified interaction Hamiltonian (PMIH) employed in previous studies. Our calculations show that, in most resonant cases, the Raman images from GIIH are similar to those from PMIH when the origin is set to the nuclear charge center of the molecule. In the case of nonresonant Raman images, distinct differences can be found from two different approaches, while GIIH calculations provide more details and phase information of the images. Furthermore, the results from GIIH calculations are more stable with respect to the computational parameters. Our results not only help to correctly simulate the resonant and nonresonant Raman images of single molecules but also lay the foundation for developing gauge invariant theory for other linear and nonlinear optical processes under the excitation of non-uniform electromagnetic field.

I. INTRODUCTION

Light-matter interaction governs all optical spectroscopies and is also the reason why we can see objects. Among diverse types of spectroscopies, emission spectroscopy is one of the most suitable techniques for optical imaging because it can avoid the background noise from incident sources. For example, fluorescence microscope has been used to visualize gene expression as well as protein distribution by using fluorescent proteins.1

Because of the breakdown of the diffraction limit,2 the resolution of near-field based imaging techniques is superior than that of far-field based imaging techniques, although the latter can be significantly improved with the help of “super-resolution” microscopies.3–5 In this context, tip-enhanced Raman spectroscopy (TERS),6–9 which takes the advantage of the hugely enhanced localized surface plasmon, is a promising technique to obtain high resolution images. By precisely controlling the surface plasmons, the resolution of TERS images has been significantly improved.10–17 For instance, under the ultra-high vacuum and low temperature conditions, the TERS images for a porphyrin derivative adsorbed on the silver surface with sub-nanometer resolution have been reported,16 which opens the door for exploring the inner molecular structures by optical means. In this case, the spatial distribution of the plasmonic field has been considered as the decisive factor for such amazingly high resolution.18,19 It also brings in a great challenge to the conventional response theory, in which the light is considered to be uniformly interacted with the molecule.20–27 However, without position dependence, it is not possible to generate any images from the TERS measurements.

To take the non-uniformity of the spatially confined plasmon (SCP) into account, we introduced the amplitude distribution of SCP as the electromagnetic (EM) field in the interaction Hamiltonian.19,28,29 Specifically, the conventional interaction Hamiltonian H^=𝐄^i𝝁^ was modified as H^=Mg(𝐫,𝐑T)𝐄^i𝝁^, where M is the enhancement factor, g is the normalized amplitude distribution function depending on the position of the TERS tip (RT), and 𝝁^ and 𝐄^i are the operators for electron position (r) and incident electric field, respectively. Using this phenomenologically modified interaction Hamiltonian (PMIH), we could nicely reproduce the experimental observations when the size of the SCP was set to around 20 Å.19 It is worth noting that such a small size of the SCP is consistent with advanced EM field simulations30–33 and further confirmed by recent experiments.34 We also found that, because of the dominance of Franck-Condon contributions, calculated resonant Raman images for the porphyrin derivative are not sensitive to the vibrational modes,19 which is in good agreement with the experimental observations.16 To enhance the vibrational dependent Herzberg-Teller contributions, nonresonant Raman processes were considered with the same Hamiltonian. The calculated nonresonant Raman images for water clusters adsorbed on the gold surface demonstrated that they are not only vibrationally resolved but also resemble the corresponding vibrational modes.28 As a result, tip-enhanced nonresonant Raman images could be a practical technique for the visualization of molecular vibrational modes in real space.28

So far, the theoretical frameworks for both resonant and nonresonant Raman images are based on PMIH. It is noted that, when the gauge origin moves by an arbitrary vector a, the transition matrix element between initial state |i and final state |f of PMIH would have an additional term of M𝐄0𝐚i|g|f, where E0 represents the amplitude of the electric field.35 Therefore, a gauge problem in PMIH arises from nonzero value of matrix element i|g|f, which is the characteristic feature of the non-uniform EM field. Although the gauge problem could be circumvented by the “common-origin” method if the field is approximately truncated by the field gradient,36 the high-order terms always have important contributions in Raman images.37 Thus, in this work, we derive a more general gauge invariant interaction Hamiltonian (GIIH) from the rigorous quantization of EM field for SCP to completely eliminate the gauge problem in the PMIH. Our numerical simulations show that, in resonant conditions, the GIIH gives similar results to those from the PMIH when the origin is set to the nuclear charge center of the adsorbates. On the other hand, for nonresonant Raman images, GIIH results highlight the pattern phases that reveal more vibrational information. We also find that the results from GIIH are more stable than those from the Coulomb’s gauge with respect to the choices of the computational parameters. Within the sum-over-state framework, a complete theory for Raman images is thus established. It should be emphasized that GIIH is general for the interaction between non-uniform EM field and matters, which could find applications in the description of other linear and nonlinear optical processes.

II. METHODOLOGY

The non-relativistic electronic Hamiltonian of a N-electron molecule in a general EM field could be expressed as25,38

H^e=i=1N12𝝅^i2i=1NA=1MAZAriA+i<jN1rij, (1)

where i and A are the indices for electron and nucleus, respectively, 𝝅^=𝐩^+𝐀^ is the generalized momentum operator,25,38–40 ZA represents the charge of nuclei, MA is the number of nuclei, riA and rij are the distances between nucleus-electron and electron-electron, respectively. Here, 𝐩^=ı is the momentum operator of electrons and 𝐀^ is the operator of vector potential. Compared with the electronic Hamiltonian without EM field (the unperturbed Hamiltonian H^0), through simple algebra, we have the (time-dependent) perturbation Hamiltonian25,38

H^=12iN(𝐩^i𝐀^+𝐀^𝐩^i), (2)

where the high order 𝐀^2 term is neglected. We should emphasize that, in proper gauge, the scale potential is set to zero.41 It is well known that the magnitude of magnetic field is much smaller than that of the vector potential.36 Moreover, in the minimal coupling Hamiltonian, magnetic fields are always associated with the spin operator.25,38 In our simulations, the spin of Raman processes is conserved. Thus, magnetic fields can be neglected even for Raman processes in an inhomogeneous EM field.36

For a general EM field, the operator related to the vector potential can be written as42,43

𝐀^(𝐫,t)=2πωV(𝐟𝐤(𝐫)a^eıωt+H.c.), (3)

where ω is the frequency of the light, V is the system volume, fk is the mode function that represents the spatial distribution of the field, k is the wave vector, a^ is the annihilation operator for photon, and H.c. is the Hermitian conjugate. Here the quantization of EM field is particularly necessary for spontaneous emission that is involved in Raman processes as shown in Fig. 1. Notice that, for specific nanostructures, the distribution function fk can be determined by the classical Maxwell’s equations41 if the quantum correction for tunneling is adopted.44 According to the energy conservation law in classical electromagnetic theory for metallic nanostructures without considering losses,45,46 the following normalization condition should be fulfilled:

d𝐫12[(ωϵ)ω+μ0ϵ0(ωμ)ω|τ|2|μ|2]|𝐟𝐤(𝐫)|2=V, (4)

where ϵ0 (ϵ) is the permittivity of free space (relative permittivity), μ0 (μ) is the permeability of free space (relative permeability), and τ is the ratio of magnetic induction to electric field. Therefore, substituting Eq. (3) into Eq. (2), the general adiabatic interaction Hamiltonian can be rewritten as a harmonic perturbation

H^(𝐫,t)=V^(𝐫)e(ıω+γ)t+H.c., (5)

where

V^(𝐫)=2πωVi12(𝐩^i𝐟𝐤+𝐟𝐤𝐩^i)a^ (6)

is time independent. Here γ is a positive infinitesimal that ensures H^(𝐫,)=0.47

FIG. 1.

FIG. 1.

Diagrams of resonant (type I) and nonresonant (type II) terms. |i, |r, and |f represent molecular initial, intermediate, and final states, respectively. ω and ωs are frequencies of incident and scattering lights. The number of incident (scattering) light-quantum is also labeled.

Once the interaction Hamiltonian is determined, the Raman intensity could be easily calculated by the second-order time-dependent perturbation theory. It is well-known that Raman scattering has two terms as shown in Fig. 1.24,48,49 Specifically, the resonant term (type I) represents that the initial state |i|nω first excites to the intermediate state |r|(n1)ω by absorbing an incident light-quantum. Then, the intermediate state |r|(n1)ω relaxes to the final state |f|(n1)ω1ωs by emitting a scattering light-quantum. Here, |i, |r, and |f represent molecular initial, intermediate, and final states, respectively, |nω is the state for n photon with frequency of ω, and ω (ωs) is the frequency of incident (scattering) light. The nonresonant term (type II) in Fig. 1 corresponds to that the initial state first relaxes to the intermediate state |r|nω1ωs by emitting a scattering photon and then excites to |f|(n1)ω1ωs by absorbing an incident photon.

It is noted that, in both terms, the plasmonic field is only responsible for the absorption and the vacuum field is responsible for the spontaneous emission. For the plasmonic field, we assume fk = Mg, while for the scattering field, 𝐟𝐤=eı𝐤𝐫1. We focus on the scattering light that can be experimentally detected in the far field. Thus, the dipole approximation is a good approximation for scattering and can be used in our current simulations.29 Again, here M is the enhancement factor and g is the normalized amplitude distribution function. Therefore, according to Eq. (5), for the ρσ-component of the polarizability, we have

Hσ,ω=2πωVMV^σa^ωe(ıω+γ)t+H.c.,Hρ,ωs=2πωsVV^ρa^ωse(ıωs+γ)t+H.c., (7)

where

V^σ=12iN(p^i,σgσ+gσp^i,σ),V^ρ=iNp^i,ρ (8)

are the time-independent operators and ρ (σ) represents the corresponding Cartesian coordinate. Here the annihilation (a^) and creation (a^) operators highlight the absorption and emission processes, respectively.

Following the standard second-order time-dependent perturbation theory,40,48,50 the differential Raman scattering cross section could be obtained (see the Appendix for details). In brief, the total second-order perturbation coefficient for Raman is

cfi(2)(t)=2πMVnωωseı(ωfiω+ωs)t+2γtωfiω+ωs2ıγ(ωfi), (9)

where n is the photon number of incident light, ωfi is the frequency difference between molecular states |f and |i,and

(ωfi)=r(f|V^ρ|rr|V^σ|iωriωıγ+f|V^σ|rr|V^ρ|iωri+ωsıγ). (10)

Thus, the transition rate becomes

Γfi(2)=(2π)3M2nV2ωωs|(ωfi)|2δ(ωfiω+ωs) (11)

because γ is a positive infinitesimal. If the density of final state is considered, we have

dΓfi(2)dΩ=nωωs3M2Md2FPVc3|1ωωs(ωfi)|2, (12)

where Ω is the solid angle, Md is the directional radiation pattern factor owing to the nanostructure,27 FP is the Purcell factor51 that accounts for the enhancement of spontaneous emission, and c is the speed of light. Here the Purcell factor is necessary, which leads to the well-known |E|4 scaling in the surface-enhanced Raman scattering.29 Finally, the differential scattering cross section for Raman processes under plasmonic field can be calculated as

dσfidΩ=ωωs3M2Md2FPc4|𝜶fi|2, (13)

where, comparing with the cross section of classical oscillating dipole,21,48 we can define the general polarizability derived from Eq. (7) as

αfi,ρσ=1ωωsr(f|V^ρ|rr|V^σ|iωriωıγ+f|V^σ|rr|V^ρ|iωri+ωsıγ). (14)

Here Eq. (10) is used.

It should be emphasized that, in free space, we have V^σ=ip^i,σ and V^ρ=ip^i,ρ. Thus, Eq. (14) becomes exactly the original Dirac’s expression for polarizability shown in Ref. 48 if the small term ıγ is neglected. Notice that, from the Dirac’s expression to the more popular Kramers and Heisenberg’s expression,52 the relationship V^σ=ı[H^0,μ^σ] plays an essential role.48 However, for the cases with plasmonic field, the corresponding relationship may be invalid. Thus, we have to keep the general polarizability Eq. (14) for Raman images from GIIH.

In practical calculations, the Born-Oppenheimer approximation53 is adopted. As a result, analogy to the elegant Albrecht’s Raman theory,20,26 the general polarizability for resonant Raman images from GIIH could be expressed by

αfi,ρσ=Aρσ+Bρσ, (15)

where

Aρσ=1ωωs[Ψg|V^ρ|ΨrΨr|V^σ|Ψgvr=0υf|υrυr|υiωerυr:egυiωıγ]+NRT,Bρσ=1ωωsk[Ψg|V^ρ|ΨrQkΨr|V^σ|Ψgυr=0υf|Qk|υrυr|υiωerυr:egυiωıγ+Ψg|V^ρ|ΨrΨr|V^σ|ΨgQkυr=0υf|υrυr|Qk|υiωerυr:egυiωıγ]+NRT. (16)

Here |Ψg and |Ψr are the electronic ground and resonant excited states, Qk is the vibrational normal mode, |υi and |υf are the initial and final vibrational states in |Ψg, |υr is the vibrational state in |Ψr, ωerυr:egυi is the frequency difference between |Ψr|υr and |Ψg|υi, and NRT is the nonresonant term. On the other hand, for nonresonant Raman processes, the NRT will come into play.20,26 Using the state-to-state mapping relationship between the Albrecht’s theory and the perturbation theory,54 the general polarizability becomes

αfi,ρσk=αρσeffQkυf|Qk|υi, (17)

where

αρσeff=1ωωsr(Ψg|V^ρ|ΨrΨr|V^σ|ΨgΔErgω+Ψg|V^σ|ΨrΨr|V^ρ|ΨgΔErg+ωs). (18)

Here ΔErg represents the vertical excitation energy between |Ψg and |Ψr and ıγ is neglected. It should be stressed that, owing to the multimode nature of vibrations, i.e., |υ=|υ1υ2υN (where N is the number of vibrational modes), we can calculate the required vibrational integral in Eq. (17) as υf|Qk|υi=δ1f,1iδ2f,2iυkf|Qk|υkiδNf,Ni. Thus, the normal mode “k” is naturally selected out for harmonic potential energy surfaces in nonresonant Raman cases. Substituting Eqs. (15) and (17) into Eq. (13), the cross section from GIIH could be readily calculated.

For clear comparison, the polarizability expressions from PMIH are also listed. For resonant Raman images, we have

Aρσr=[Ψg|μ^ρ|ΨrΨr|μ^σgσ|Ψgυr=0υf|υrυr|υiωerυr:egυiωıγ]+NRT,Bρσr=k[Ψg|μ^ρ|ΨrQkΨr|μ^σgσ|Ψgυr=0υf|Qk|υrυr|υiωerυr:egυiωıγ+Ψg|μ^ρ|ΨrΨr|μ^σgσ|ΨgQkυr=0υf|υrυr|Qk|υiωerυr:egυiωıγ]+NRT, (19)

and for nonresonant Raman images

αρσeff,r=r(Ψg|μ^ρ|ΨrΨr|μ^σgσ|ΨgΔErgω+Ψg|μ^σgσ|ΨrΨr|μ^ρ|ΨgΔErg+ωs). (20)

III. COMPUTATIONAL DETAILS

To illustrate the algorithm for the GIIH and compare its results with those from the PMIH, we chose a single meso-tetrakis(3,5-di-tert-butylphenyl)porphyrin (H2TBPP) molecule adsorbed on the Ag(111) surface16,19 and water clusters on the Au(111) surface28 for the simulations of resonant and nonresonant Raman images, respectively. All calculation details could be found in our previous studies19,28 as well as the supplementary material. Briefly, geometrical optimizations and vibrational calculations were performed at the periodic boundary conditions implemented as in the Vienna ab initio simulation package.55 Then, the optimized geometries as well as the vibrational modes were extracted for electronic structure calculations of ground and excited states by the gaussian suite of program.56

For vibrational integrals in Eq. (16), the linear coupling model57 was used for the excited state potential energy surfaces. Therefore, all vibrational integrals and energy differences could be calculated by the DynaVib program from the vertical energy derivatives extracted from time-dependent density functional theory (TDDFT) calculations.58 In addition, all other quantities for the electronic part in Eqs. (16) and (18), including transition velocity dipole moments, vertical excitation energies as well as their derivatives, could be directly extracted from calculations by the gaussian program except the integral

I=Ψr|V^σ|Ψg. (21)

According to the experimental setups,16 evaluation of the zz component is adequate for Raman images. For practical simulations, we expanded the z component of the amplitude distribution of the SCP by a set of Gaussian functions, i.e.,

gz=Dl,m,nαcα,Dlmngα,Dlmn, (22)

where gα,Dlmn is a Gaussian function localized at the center rD with exponent α, which can be written as

gα,Dlmn=(xxD)l(yyD)m(zzD)neα(𝐫𝐫D)2 (23)

and cα,Dlmn is the corresponding coefficient. Thus, the integral

Ilmn=Ψr|V^zlmn|Ψg (24)

should be evaluated. It should be emphasized that the superscript “lmn” relates to the angular momentum of Cartesian Gaussian functions (see Eq. (23)).

In the framework of TDDFT, using the coefficients of excitation (X) and de-excitation (Y), we have59

Ilmn=ia(Xiarϕa|υ^zlmn|ϕi+Yiarϕi|υ^zlmn|ϕa), (25)

where ϕi and ϕa are the occupied and unoccupied molecular orbitals (MOs), respectively, and the definition

υ^zlmn=12(p^zgα,Dlmn+gα,Dlmnp^z). (26)

Expanding MOs by atomic orbitals, the required integral between arbitrary MOs ϕp and ϕq could be

Ipqlmn=χηcχpcηq(KLDχKDηLgχK|υ^zlmn|gηL), (27)

where cχp and cηq are the MO coefficients, gχK and gηL are the primitive Cartesian Gaussian functions, and DχK and DηL are the corresponding contraction coefficients. Therefore, the integral with primitive Cartesian Gaussian functions

IKLlmn=gχK|υ^zlmn|gηL (28)

has to be calculated. It is noted that if

IKL,rlmn=gχK|zgα,Dlmn|gηL (29)

is evaluated instead, results from the PMIH are obtained.

Owing to the relationship60

zgαlmn=2αzgαlm(n+1)+ngαlm(n1), (30)

IKLlmn could be converted to a linear combination of the three-center overlap integrals. Here gαlmn is the general expression of both the gα,Dlmn in amplitude distribution function and primitive Cartesian Gaussian functions, meanwhile, αz is the z component of the exponent. Practically, the integral Eq. (28) could be factored as

IKLlmn=IKL,xlIKL,ymIKL,zn=la|lD|lbma|mD|mbna|v^(nD)|nb, (31)

where subscripts a, b, and D represent the different centers of Gaussian functions, and v^(nD) is the n related component of v^zlmn. To calculate Eq. (31), both direct and recursive methods could be used.60 Here we only present the efficient recursive method. Specifically, for IKL,xl and IKL,ym (here only taking IKL,xl, for example), we have60

la+1|lD|lb=(Gx𝐫χ,x)la|lD|lb+la2αxTotla1|lD|lb+lb2αxTotla|lD|lb1+lD2αxTotla|lD1|lb (32)

with

0a,x|0D,x|0b,x=exp(αxTotGx2αχK𝐫χ,x2αηL𝐫η,x2αx𝐫D,x2)×παxTot, (33)

where 𝐫χ (𝐫η) is the center of atomic basis set χ (η), and

αxTot=αχK+αηL+αx,Gx=αχK𝐫χ,x+αηL𝐫η,x+αx𝐫D,xαxTot. (34)

It is noted that the l and m components in IKL,rlmn could also be calculated by Eq. (32) and both are gauge origin independent. For IKL,zn, the recurrence formula reads

na+1|υ^(nD)|nb=(Gz𝐫χ,z)na|υ^(nD)|nb+na2αzTotna1|υ^(nD)|nb+nb2αzTotna|v^(nD)|nb1+nD2αzTotna|υ^(nD1)|nb+ı2αηL+αz2αzTotna|nD|nb (35)

with

0a,z|υ^(nD+1)|0b,z=(Gz𝐫D,z)0a,z|υ^(nD)|0b,z+nD2αzTot0a,z|υ^(nD1)|0b,zı(0.52αηL+αz2αzTot)0a,z|nD,z|0b,z (36)

and

0a,z|v^(0D,z)|0b,z=ı[2αηL(Gz𝐫η,z)+αz(Gz𝐫D,z)]×0a,z|0D,z|0b,z, (37)

where 0a,z|0D,z|0b,z as well as αzTot and Gz could be calculated from Eqs. (33) and (34) by replacing the subscript x with z. It should be emphasized that all distances in the evaluation of Eq. (35) are relative values, which manifests that the integral IKLlmn is gauge origin independent. On the other hand, for n related component of IKL,rlmn, we have the recurrence formula29

na+1|znD|nb=(Gz𝐫χ,z)na|znD|nb+na2αzTotna1|znD|nb+nb2αzTotna|znD|nb1+nD2αzTotna|z(nD1)|nb+12αzTotna|nD|nb, (38)

with

0a,z|z0D,z|0b,z=Gz0a,z|0D,z|0b,z. (39)

Because of the gauge origin dependence of Gz and the nonzero value of ϕp|gα,Dlmn|ϕq, IKL,rlmn as well as the results from the PMIH are thus also gauge origin dependent.

In practical simulations, as the same in our previous studies, we use one s-type Gaussian function for gz.19,28 Specifically, we set the full width at half maximum for the x and y components to 20 Å and its theoretical limit18,45 of 1 Å for resonant19 and nonresonant28 Raman image calculations, respectively. Meanwhile, the full width at half maximum for the z component was set to 5 Å.61 For resonant Raman images, the two near degenerate excited states around the experimental excitation energy for each isomer are accounted equally in the current simulations. For nonresonant cases, the wavelength of the incident light was set to 632.8 nm. All Raman images were calculated by the “First-Principles Approaches for Surface and Tip Enhanced Raman Scattering (FASTERS)” program.62

IV. RESULTS AND DISCUSSIONS

We first consider the resonant Raman images for H2TBPP adsorbed on the Ag(111) surface16 from three different adsorption configurations,19 namely, concave, plane, and convex, respectively. All calculated results from both GIIH and PMIH are depicted in Fig. 2. It should be emphasized that, for different interaction Hamiltonians, the near field around the molecule is the same. The only difference is that we artificially chose different origins in our simulations.

FIG. 2.

FIG. 2.

Calculated resonant Raman images for concave, plane, and convex configurations (from left to right) of H2TBPP adsorbed on the Ag(111) surface from (a) gauge invariant interaction Hamiltonian (GIIH) and (b) phenomenologically modified interaction Hamiltonian (PMIH). In (a) the optimized structures extracted from Ref. 19 are shown in the top panel; the experimentally observed Raman images from Ref. 16 are shown as insert to the top left panel. In (b) the origins were set to the plasmonic center (PC), nuclear charge center of adsorbates (NC), and the topmost layer of the Ag(111) surface (SL) from top to bottom, respectively. The solid lines represent the skeleton of H2TBPP.

We could immediately notice that the calculated images from GIIH are similar to those from PMIH with the origin set to the nuclear charge center of adsorbates (labeled as NC hereafter) for concave and plane configurations. Specifically, four bright corner lobes with less contrast for the bottom-left corner lobe and two (top and bottom) bright ribbons are predicted by both GIIH and PMIH. On the other hand, for the convex configuration, the image from GIIH shows three bright lobes, which are significantly different from the oval-shaped pattern predicted by the PMIH with the origin set to NC but similar to that with the origin set to the plasmonic center (labeled as PC hereafter). It is noted that only the concave configuration can nicely reproduce the experimental observations, which is consistent with the results from PMIH with origin at NC.19 It is also noted that the concave configuration is the most stable one on the Ag(111) surface.19 For instance, its energy is around 2.9 eV lower than that of the convex configuration.

From Fig. 2, we could find that the results from the PMIH are indeed origin dependent, especially for the convex configuration. Specifically, when the origin is set to the PC, three bright lobes are highlighted, which is similar to the result from GIIH as mentioned before. When the origin is set to the topmost metallic surface layer (labeled as SL hereafter), a slight deformation of the oval-shaped pattern is observed. Meanwhile, for the concave and plane configurations, the origin dependence of Raman images is relatively weak with minor contrast changes of bottom-left and top patterns for concave and plane configurations, respectively.

Then we move on to the nonresonant cases. All calculated nonresonant images for water monomer adsorbed on the Au(111) surface28 from both GIIH and PMIH are depicted in Fig. 3. In contrast to the resonant case, all nonresonant Raman images including bending mode (v2) as well as symmetric (v1) and asymmetric (v3) O–H stretching modes from GIIH are significantly different from their counterparts from PMIH when the origin is set to NC. Specifically, the patterns associated with hydrogen atoms in the v1 and v3 images become obscure, while, the patterns near to the oxygen atom are highlighted. It should be emphasized that the phases (sign of αeff/Qk in Eq. (17)) of each pattern are important. In the v1 image, the phase of pattern associated with oxygen is different from that with hydrogens. In addition, in the v3 image, the phases of four patterns are the same as that of the dxy atomic orbital. The phases suggest the constructive and destructive coherence of O–H stretching in the v1 and v3 images, which is consistent with the normal modes. The images for bending mode (v2) from GIIH and PMIH with origin at NC both give a bright ribbon pattern associated with two hydrogen atoms. However, the ball-shaped pattern from PMIH splits into two weak patterns along the outside area of O–H bonds. Comparing the results from PMIH and GIIH, we find that the simulation results from PMIH, for example, with origin at NC shown in Fig. 3(b), may lead to the missing of vibrational phase information. The phase information could exist in experiments and would be the important characteristic feature of different vibrational modes.

FIG. 3.

FIG. 3.

Calculated nonresonant Raman images for bending (v2), symmetric stretching (v1), and asymmetric stretching (v3) modes (from left to right) of water monomer adsorbed on the Au(111) surface from (a) gauge invariant interaction Hamiltonian (GIIH) and (b) phenomenologically modified interaction Hamiltonian (PMIH). In (a) the vibrational modes extracted from Ref. 28 are shown as top panel; the values are calculated frequencies in cm1. In (b) the origins were set to the plasmonic center (PC), nuclear charge center of adsorbates (NC), and the topmost layer of the Au(111) surface (SL) from top to bottom, respectively. The solid lines represent the skeleton of water monomer.

Figure 3(b) shows that, for PMIH results, v3 images are not sensitive to the origin position. However, v1 and v2 images exhibit significant origin dependence. For instance, when the origin is set to PC, the two patterns associated with O–H bonds in v1 image are obscure and a new bright pattern associated with the oxygen atom emerges. For the v2 image, the contrast of ribbon associated with the two hydrogen atoms becomes lower and the size of the bell-shaped pattern associated with the oxygen is reduced when the origin is set to either PC or SL.

For larger water clusters adsorbed on the Au(111) surface, because the images of O–H stretching modes are simple and highly localized (see Figs. S1-S3 in the supplementary material for details), only the stretching modes for hydrogen bonds as well as the bending modes are shown in Fig. 4. Calculated results show that GIIH and PMIH can give similar patterns for the O–H stretching modes (see Figs. S1-S3 in the supplementary material for details).

FIG. 4.

FIG. 4.

Calculated nonresonant Raman images for hydrogen bonding stretching modes (vH for dimer; vHE and vHA for trimer) and bending modes (v2s and v2a for dimer; v2E and v2A for trimer) of water dimer (left) and trimer (right) adsorbed on the Au(111) surface from ((a) and (b)) gauge invariant interaction Hamiltonian (GIIH) and ((c) and (d)) phenomenologically modified interaction Hamiltonian (PMIH). In ((a) and (b)) the vibrational modes extracted from Ref. 28 are shown as top panel; the values are calculated frequencies in cm1. In ((c) and (d)) the origins were set to the plasmonic center (PC), nuclear charge center of adsorbates (NC), and the topmost layer of the Au(111) surface (SL) from top to bottom, respectively. The solid lines represent the skeleton of water dimer and trimer.

For water dimer, the image calculated from GIIH for the hydrogen bond stretching mode (vH) displays a bright pattern between the two water molecules. The image for antisymmetric mixing of v2 modes (v2a) has three bright patterns associated with the top water molecule, while, the image for symmetric mixing (v2s) has one bright pattern associated with the bottom water molecule. All these images are similar to those from PMIH with origin at NC, except that the patterns of the top water molecule are obscure in the image of v2s. When the origin is set to PC or SL, the images from PMIH again show some differences. For instance, the moderate pattern between the two water molecules in vH image becomes obscure. In addition, for the v2s image, the contrast of three patterns associated with the top water is lower, meanwhile, the pattern associated with oxygen in the bottom water is highlighted.

We now turn our attention to the case of water trimer. The calculated pattern for the degenerate asymmetric modes of hydrogen bonds stretching (vHE) from GIIH is similar to that from PMIH with origin at NC. It is noted that the calculated image for symmetric stretching of hydrogen bonds (vHA) from GIIH exhibits the constructive coherence, which leads to the bright center of the trimer ring. This constructive coherence is also observed in the PMIH result with origin at NC.28 However, the contrast of the bright center is higher in the PMIH result. For the bending modes, the GIIH images for the degenerate asymmetric modes (v2E) and the symmetric modes (v2A) both show bright ribbon patterns associated with each water monomer. However, the phases of the patterns reveal the destructive coherence between v2 images for water monomer in the v2E image, while the constructive coherence in the v2A image. This is opposite to the results from PMIH with origin at NC, where the v2E and v2A images exhibit the constructive and destructive coherence, respectively.28 It is noted that, when the origin is set to PC or SL, the coherence would be significantly reduced in PMIH results.

In some cases, the Coulomb’s gauge25,39,41,46

div𝐀(𝐫,t)=0 (40)

is usually adopted. Using the mathematic relationship41 (𝐚f)=𝐚f+f𝐚, where a and f are the arbitrary vector and scale functions, we can rewrite Eq. (26) as

v^zlmn=gα,Dlmnp^z+w^zlmn. (41)

Here we have the operator definition for gradient contribution

w^zlmn=ı2gα,Dlmnz. (42)

In the current simulations, the gradient contribution in general is not zero. Therefore, in the Coulomb’s gauge, we should subtract the gradient contribution from the whole GIIH.

To evaluate the gradient contribution, the integral

IKL,wlmn=la|lD|lbma|mD|mbna|w^(nD)|nb (43)

has to be calculated. The recurrence formula for the only unknown integral in Eq. (43) reads

na+1|w^(nD)|nb=(Gz𝐫χ,z)na|w^(nD)|nb+na2αzTotna1|w^(nD)|nb+nb2αzTotna|w^(nD)|nb1+nD2αzTotna|w^(nD1)|nb+ıαz2αzTotna|nD|nb (44)

with

0a,z|w^(nD+1)|0b,z=(Gz𝐫D,z)0a,z|w^(nD)|0b,z+nD2αzTot0a,z|w^(nD1)|0b,zı(0.5αz2αzTot)0a,z|nD,z|0b,z (45)

and

0a,z|w^(0D,z)|0b,z=ıαz(Gz𝐫D,z)0a,z|0D,z|0b,z. (46)

It is noteworthy that IKL,wlmn is again gauge origin independent.

The calculated resonant Raman images for H2TBPP adsorbed on the Ag(111) surface and nonresonant Raman images for water monomer adsorbed on the Au(111) surface from the Coulomb’s gauge as well as the gradient contribution have been depicted in Fig. 5 as examples. For concave and convex configurations, the results from the Coulomb’s gauge are similar to their counterparts from the GIIH as well as the gradient contributions. On the other hand, the image for plane configuration from the Coulomb’s gauge shows a bright bottom-right corner lobe, which is of significant difference from the result from GIIH, whereas, the gradient contribution for the plane configuration displays two bright patterns that are similar to the result from GIIH. For the nonresonant cases, the results from the Coulomb’s gauge nicely reproduce their GIIH counterparts for all vibrational modes.

FIG. 5.

FIG. 5.

Calculated (a) resonant Raman images for concave, plane, and convex configurations (from left to right) of H2TBPP adsorbed on the Ag(111) surface and (b) nonresonant Raman images for bending (v2), symmetric stretching (v1), and asymmetric stretching (v3) modes (from left to right) of water monomer adsorbed on the Au(111) surface from the Coulomb’s gauge (CG, top) and the gradient contribution (GC, bottom). The values represent the corresponding scale factor and the scale of the contrast is the same as that in Figs. 2 and 3.

It is noted that, in Fig. 5, the gradient contribution is important in resonant images, while, it is negligible in nonresonant images. This should be attributed to the different confinement parameters used for resonant and nonresonant images. Specifically, for resonant images, αx,y is smaller than αz, while for nonresonant images, αx,y is larger. It is anticipated that smaller αz will reduce the importance of the gradient contribution. Therefore, it is beneficial to consider the case of αz=0. In this case, the gradient contribution is zero and the Coulomb’s gauge is automatically fulfilled. Now, the distribution of plasmon becomes a Gaussian cylinder and the tip position is no longer important.

All calculated Raman images with αz=0 are shown in Fig. 6. The calculated resonant images for concave and convex are similar with those from finite αz. However, for the plane configuration, the significant shape changes of two patterns are observed. For nonresonant cases, the images with αz=0 are the same as those from finite αz. By comparing all results with different αz, we find that the images from the whole GIIH are more stable with respect to the choice of αz than those from the Coulomb’s gauge.

FIG. 6.

FIG. 6.

Calculated (a) resonant Raman images for concave, plane, and convex configurations (from left to right) of H2TBPP adsorbed on the Ag(111) surface and (b) nonresonant Raman images for bending (v2), symmetric stretching (v1), and asymmetric stretching (v3) modes (from left to right) of water monomer adsorbed on the Au(111) surface from gauge invariant interaction Hamiltonian with αz set to zero.

V. CONCLUSIONS

We have derived and implemented the gauge invariant algorithm from the GIIH for both resonant and nonresonant Raman images, where the gauge origin dependence in PMIH is eliminated. The similarity and difference between the results of GIIH and PMIH have been carefully examined. The GIIH based theory is ready for simulating Raman images of different molecular systems and can also be further extended to describe other linear and nonlinear optical processes under the electromagnetic field with inhomogeneous spatial distribution in general.

SUPPLEMENTARY MATERIAL

See supplementary material for the detailed computational parameters and other nonresonant Raman images for water dimer and trimer adsorbed on the Au(111) surface.

ACKNOWLEDGMENTS

This work was supported by the National Natural Science Foundation of China (Grant No. 21421063), the “Strategic Priority Research Program” of the Chinese Academy of Sciences (Grant No. XDB01020200), and Swedish Research Council (VR). The Swedish National Infrastructure for Computing (SNIC) is acknowledged for computer time.

APPENDIX: DERIVATION OF GENERAL RAMAN POLARIZABILITY

In time-dependent perturbation theory, the time-dependent coefficients for state |m with different orders could be calculated by the Dyson series, i.e.,40,50

cmi(0)(t)=δmi,cmi(1)(t)=ıtm|H(t)|ieıωmitdt,cmi(2)(t)=ırtm|H(t)|reıωmrtcri(1)(t)dt, (A1)

where |i and |r are the initial and intermediate states, respectively. Using Eq. (7) and according to Fig. 1, we can find the first-order time-dependent coefficient for type I as

cI,ri(1)(t)=ı2πωVMt(n1)ω|r|V^σa^ω|i|nωeı(ωriω)t+γtdt=ı2πnωVMr|V^σ|ieı(ωriω)t+γtı(ωriω)+γ, (A2)

and for type II

cII,ri(1)(t)=ı2πωsVtnω1ωs|r|V^ρa^ωs|i|nω0ωseı(ωri+ωs)t+γtdt=ı2πωsVr|V^ρ|ieı(ωri+ωs)t+γtı(ωri+ωs)+γ. (A3)

Therefore, the desired second-order coefficient for type I is

cI,fi(2)(t)=2πMVnωωsrr|V^σ|iı(ωriω)+γteı(ωfr+ωriω+ωs)t+2γt×(n1)ω1ωs|f|V^ρa^ωs|r|(n1)ω0ωsdt=2πMVnωωsrf|V^ρ|rr|V^σ|iωriωıγeı(ωfiω+ωs)t+2γtωfiω+ωs2ıγ, (A4)

and for type II

cII,fi(2)(t)=2πMVnωωsrr|V^ρ|iı(ωri+ωs)+γteı(ωfr+ωri+ωsω)t+2γt×(n1)ω1ωs|f|V^σa^ω|r|nω1ωsdt=2πMVnωωsrf|V^σ|rr|V^ρ|iωri+ωsıγeı(ωfiω+ωs)t+2γtωfiω+ωs2ıγ, (A5)

where the relationship ωfi=ωfr+ωri is used. Therefore, adding cI,fi(2) and cII,fi(2), the total second-order coefficient can be obtained as shown in Eq. (9).

By definition, the second-order transition rate from initial state to final state is39,40

Γfi(2)=ddt|cfi(2)(t)|2=(2π)2M2nV2ωωs|(ωfi)|24γe4γt(ωfiω+ωs)2+(2γ)2. (A6)

Here the definition of Eq. (10) is used. Furthermore, using the relationship

limγ04γe4γt(ωfiω+ωs)2+(2γ)2=2πδ(ωfiω+ωs), (A7)

Eq. (A6) could be rewritten as Eq. (11). It is noted that the density state of spontaneous emission is continuum and can be expressed as27,40,51

d3k=Vωs2Md2FP(2πc)3dΩdωs. (A8)

As a result, we have

Γfi(2)=dΩdωsVωs2Md2FP(2πc)3(2π)3M2nV2ωωs|(ωfi)|2δ(ωfiω+ωs), (A9)

which leads to Eq. (12). Because the flux of incident light is

J=ncV, (A10)

according to the definition of cross section

dσfidΩ=dΓfi(2)dΩJ, (A11)

the general differential scattering cross section for Raman processes Eq. (13) is obtained.

REFERENCES

  • 1.Chalfie M., Tu Y., Euskirchen G., Ward W. W., and Prasher D. C., Science 263, 802 (1994). 10.1126/science.8303295 [DOI] [PubMed] [Google Scholar]
  • 2.Betzig E. and Trautman J. K., Science 257, 189 (1992). 10.1126/science.257.5067.189 [DOI] [PubMed] [Google Scholar]
  • 3.Hell S. W. and Wichmann J., Opt. Lett. 19, 780 (1994). 10.1364/ol.19.000780 [DOI] [PubMed] [Google Scholar]
  • 4.Betzig E., Patterson G. H., Sougrat R., Lindwasser O. W., Olenych S., Bonifacino J. S., Davidson M. W., Lippincott-Schwartz J., and Hess H. F., Science 313, 1642 (2006). 10.1126/science.1127344 [DOI] [PubMed] [Google Scholar]
  • 5.Rust M. J., Bates M., and Zhuang X., Nat. Methods 3, 793 (2006). 10.1038/nmeth929 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Stöckle R. M., Suh Y. D., Deckert V., and Zenobi R., Chem. Phys. Lett. 318, 131 (2000). 10.1016/s0009-2614(99)01451-7 [DOI] [Google Scholar]
  • 7.Anderson M. S., App. Phys. Lett. 76, 3130 (2000). 10.1063/1.126546 [DOI] [Google Scholar]
  • 8.Hayazawa N., Inouye Y., Sekkat Z., and Kawata S., Opt. Commun. 183, 333 (2000). 10.1016/s0030-4018(00)00894-4 [DOI] [Google Scholar]
  • 9.Pettinger B., Picardi G., Schuster R., and Ertl G., Electrochem. Jpn. 68, 942 (2000). [Google Scholar]
  • 10.Hartschuh A., Sánchez E. J., Xie X. S., and Novotny L., Phys. Rev. Lett. 90, 095503 (2003). 10.1103/physrevlett.90.095503 [DOI] [PubMed] [Google Scholar]
  • 11.Anderson N., Hartschuh A., Cronin S., and Novotny L., J. Am. Chem. Soc. 127, 2533 (2005). 10.1021/ja045190i [DOI] [PubMed] [Google Scholar]
  • 12.Steidtner J. and Pettinger B., Phys. Rev. Lett. 100, 236101 (2008). 10.1103/physrevlett.100.236101 [DOI] [PubMed] [Google Scholar]
  • 13.Ichimura T., Fujii S., Verma P., Yano T., Inouye Y., and Kawata S., Phys. Rev. Lett. 102, 186101 (2009). 10.1103/physrevlett.102.186101 [DOI] [PubMed] [Google Scholar]
  • 14.Yano T.-A., Verma P., Saito Y., Ichimura T., and Kawata S., Nat. Photonics 3, 473 (2009). 10.1038/nphoton.2009.74 [DOI] [Google Scholar]
  • 15.Stadler J., Schmid T., and Zenobi R., Nano Lett. 10, 4514 (2010). 10.1021/nl102423m [DOI] [PubMed] [Google Scholar]
  • 16.Zhang R., Zhang Y., Dong Z. C., Jiang S., Zhang C., Chen L. G., Zhang L., Liao Y., Aizpurua J., Luo Y., Yang J. L., and Hou J. G., Nature 498, 82 (2013). 10.1038/nature12151 [DOI] [PubMed] [Google Scholar]
  • 17.Chen C., Hayazawa N., and Kawata S., Nat. Commun. 5, 3312 (2014). 10.1038/ncomms4312 [DOI] [PubMed] [Google Scholar]
  • 18.Atkin J. M. and Raschke M. B., Nature 498, 44 (2013). 10.1038/498044a [DOI] [PubMed] [Google Scholar]
  • 19.Duan S., Tian G., Ji Y., Shao J., Dong Z. C., and Luo Y., J. Am. Chem. Soc. 137, 9515 (2015). 10.1021/jacs.5b03741 [DOI] [PubMed] [Google Scholar]
  • 20.Albrecht A. C., J. Chem. Phys. 34, 1476 (1961). 10.1063/1.1701032 [DOI] [Google Scholar]
  • 21.Neugebauer J., Reiher M., Kind C., and Hess B. A., J. Comput. Chem. 23, 895 (2002). 10.1002/jcc.10089 [DOI] [PubMed] [Google Scholar]
  • 22.Masiello D. J. and Schatz G. C., Phys. Rev. A 78, 042505 (2008). 10.1103/physreva.78.042505 [DOI] [Google Scholar]
  • 23.Lee S.-Y., Zhang D., McCamant D. W., Kukura P., and Mathies R. A., J. Chem. Phys. 121, 3632 (2004). 10.1063/1.1777214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Sakurai J. J., Advanced Quantum Mechanics (Addison-Wesley Publishing Company, Reading, MA, 1967). [Google Scholar]
  • 25.Scully M. O. and Zubairy M. S., Quantum Optics, 1st ed. (Cambridge University Press, Cambridge, New York, 1997). [Google Scholar]
  • 26.Long D. A., The Raman Effect: A Unified Treatment of the Theory of Raman Scattering by Molecules (Wiley, Chichester, New York, 2002). [Google Scholar]
  • 27.Le Ru E. and Etchegoin P., Principles of Surface-Enhanced Raman Spectroscopy: And Related Plasmonic Effects, 1st ed. (Elsevier, Amsterdam, Boston, 2009). [Google Scholar]
  • 28.Duan S., Tian G., and Luo Y., Angew. Chem., Int. Ed. 55, 1041 (2016). 10.1002/anie.201508218 [DOI] [PubMed] [Google Scholar]
  • 29.Duan S., Tian G., and Luo Y., J. Chem. Theory Comput. 12, 4986 (2016). 10.1021/acs.jctc.6b00592 [DOI] [PubMed] [Google Scholar]
  • 30.Sun M., Zhang Z., Chen L., Sheng S., and Xu H., Adv. Opt. Mater. 2, 74 (2014). 10.1002/adom.201300296 [DOI] [Google Scholar]
  • 31.Meng L., Yang Z., Chen J., and Sun M., Sci. Rep. 5, 9240 (2015). 10.1038/srep09240 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Zhang C., Chen B.-Q., and Li Z.-Y., J. Phys. Chem. C 119, 11858 (2015). 10.1021/acs.jpcc.5b02653 [DOI] [Google Scholar]
  • 33.Barbry M., Koval P., Marchesin F., Esteban R., Borisov A. G., Aizpurua J., and Sánchez-Portal D., Nano Lett. 15, 3410 (2015). 10.1021/acs.nanolett.5b00759 [DOI] [PubMed] [Google Scholar]
  • 34.Jiang S., Zhang Y., Zhang R., Hu C., Liao M., Luo Y., Yang J., Dong Z., and Hou J. G., Nat. Nanotech. 10, 865 (2015). 10.1038/nnano.2015.170 [DOI] [PubMed] [Google Scholar]
  • 35.Luber S. and Reiher M., Chem. Phys. 346, 212 (2008). 10.1016/j.chemphys.2008.01.046 [DOI] [Google Scholar]
  • 36.Chulhai D. V. and Jensen L., J. Phys. Chem. C 117, 19622 (2013). 10.1021/jp4062626 [DOI] [Google Scholar]
  • 37.Hu W., Duan S., and Luo Y., Wiley Interdiscip. Rev.: Comput. Mol. Sci. 7, e1293 (2017). 10.1002/wcms.1293 [DOI] [Google Scholar]
  • 38.Jensen F., Introduction to Computational Chemistry (John Wiley & Sons, Chichester, England, Hoboken, NJ, 2007). [Google Scholar]
  • 39.Shankar R., Principles of Quantum Mechanics, 2nd ed. (Plenum Press, New York, 1994). [Google Scholar]
  • 40.Zettili N., Quantum Mechanics: Concepts and Applications, 1st ed. (Wiley, Chichester, New York, 2001). [Google Scholar]
  • 41.Jackson J. D., Classical Electrodynamics, 3rd ed. (Wiley, New York, 1998). [Google Scholar]
  • 42.Tame M. S., McEnery K. R., Özdemir Ş. K., Lee J., Maier S. A., and Kim M. S., Nat. Phys. 9, 329 (2013). 10.1038/nphys2615 [DOI] [Google Scholar]
  • 43.Parkins A., “Single-atom transistors for light” in Fundamentals of Picoscience (CRC Press, 2013), pp. 635–654. [Google Scholar]
  • 44.Zhu W., Esteban R., Borisov A. G., Baumberg J. J., Nordlander P., Lezec H. J., Aizpurua J., and Crozier K. B., Nat. Commun. 7, 11495 (2016). 10.1038/ncomms11495 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kong J. A., Electromagnetic Wave Theory, 2nd ed. (Wiley, New York, 1990). [Google Scholar]
  • 46.Archambault A., Marquier F., Greffet J.-J., and Arnold C., Phys. Rev. B 82, 035411 (2010). 10.1103/physrevb.82.035411 [DOI] [Google Scholar]
  • 47.Olsen J. and Jørgensen P., J. Chem. Phys. 82, 3235 (1985). 10.1063/1.448223 [DOI] [Google Scholar]
  • 48.Dirac P. A. M., Proc. R. Soc. London, Ser. A 114, 710 (1927). 10.1098/rspa.1927.0071 [DOI] [Google Scholar]
  • 49.Lee S.-Y. and Heller E. J., J. Chem. Phys. 71, 4777 (1979). 10.1063/1.438316 [DOI] [Google Scholar]
  • 50.Dyson F. J., Phys. Rev. 75, 1736 (1949). 10.1103/physrev.75.1736 [DOI] [Google Scholar]
  • 51.Purcell E. M., Phys. Rev. 69, 681 (1946). [Google Scholar]
  • 52.Kramers H. A. and Heisenberg W., Z. Phys. 31, 681 (1925). 10.1007/bf02980624 [DOI] [Google Scholar]
  • 53.Born M. and Oppenheimer R., Ann. Phys. 389, 457 (1927). 10.1002/andp.19273892002 [DOI] [Google Scholar]
  • 54.Gong Z.-Y., Tian G., Duan S., and Luo Y., J. Chem. Theory Comput. 11, 5385 (2015). 10.1021/acs.jctc.5b00761 [DOI] [PubMed] [Google Scholar]
  • 55.Kresse G. and Furthmüller J., Phys. Rev. B 54, 11169 (1996). 10.1103/physrevb.54.11169 [DOI] [PubMed] [Google Scholar]
  • 56.Frisch M. J., Trucks G. W., Schlegel H. B., Scuseria G. E., Robb M. A., Cheeseman J. R., Scalmani G., Barone V., Mennucci B., Petersson G. A., Nakatsuji H., Caricato M., Li X., Hratchian H. P., Izmaylov A. F., Bloino J., Zheng G., Sonnenberg J. L., Hada M., Ehara M., Toyota K., Fukuda R., Hasegawa J., Ishida M., Nakajima T., Honda Y., Kitao O., Nakai H., Vreven T., J. A. Montgomery, Jr., Peralta J. E., Ogliaro F., Bearpark M., Heyd J. J., Brothers E., Kudin K. N., Staroverov V. N., Kobayashi R., Normand J., Raghavachari K., Rendell A., Burant J. C., Iyengar S. S., Tomasi J., Cossi M., Rega N., Millam J. M., Klene M., Knox J. E., Cross J. B., Bakken V., Adamo C., Jaramillo J., Gomperts R., Stratmann R. E., Yazyev O., Austin A. J., Cammi R., Pomelli C., Ochterski J. W., Martin R. L., Morokuma K., Zakrzewski V. G., Voth G. A., Salvador P., Dannenberg J. J., Dapprich S., Daniels A. D., Farkas Ö., Foresman J. B., Ortiz J. V., Cioslowski J., and Fox D. J., gaussian 09, Revision A.02, Gaussian, Inc., Wallingford, CT, 2009. [Google Scholar]
  • 57.Macak P., Luo Y., and Ågren H., Chem. Phys. Lett. 330, 447 (2000). 10.1016/s0009-2614(00)01096-4 [DOI] [Google Scholar]
  • 58.Tian G., Duan S., Zhang G.-P., Hu W., and Luo Y., Phys. Chem. Chem. Phys. 17, 23007 (2015). 10.1039/c5cp02736e [DOI] [PubMed] [Google Scholar]
  • 59.Pedersen T. B. and Hansen A. E., Chem. Phys. Lett. 246, 1 (1995). 10.1016/0009-2614(95)01036-9 [DOI] [Google Scholar]
  • 60.Obara S. and Saika A., J. Chem. Phys. 84, 3963 (1986). 10.1063/1.450106 [DOI] [Google Scholar]
  • 61.Qiu X. H., Nazin G. V., and Ho W., Science 299, 542 (2003). 10.1126/science.1078675 [DOI] [PubMed] [Google Scholar]
  • 62.Duan S., Tian G., and Luo Y., First-principles Approaches for Surface and Tip Enhanced Raman Scattering (FASTERS), version 1.0, Royal Institute of Technology, Sweden, 2016. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

See supplementary material for the detailed computational parameters and other nonresonant Raman images for water dimer and trimer adsorbed on the Au(111) surface.


Articles from The Journal of Chemical Physics are provided here courtesy of American Institute of Physics

RESOURCES