plasmon-driven sub-picosecond breathing of metal...

7
Nanoscale PAPER Cite this: Nanoscale, 2017, 9, 12391 Received 23rd June 2017, Accepted 28th July 2017 DOI: 10.1039/c7nr04536k rsc.li/nanoscale Plasmon-driven sub-picosecond breathing of metal nanoparticlesFranco P. Bonafé, a Bálint Aradi, b Mengxue Guan, c Oscar A. Douglas-Gallardo, a Chao Lian, c Sheng Meng, c,d Thomas Frauenheim b and Cristián G. Sánchez * a We present the rst real-time atomistic simulation on the quantum dynamics of icosahedral silver nano- particles under strong laser pulses, using time dependent density functional theory (TDDFT) molecular dynamics. We identify the emergence of sub-picosecond breathing-like radial oscillations starting immediately after laser pulse excitation, with increasing amplitude as the eld intensity increases. The ultrafast dynamic response of nanoparticles to laser excitation points to a new mechanism other than equilibrium electronphonon scattering previously assumed, which takes a much longer timescale. A sharp weakening of all bonds during laser excitation is observed, thanks to plasmon damping into excited electrons in anti-bonding states. This sudden weakening of bonds leads to a uniform expansion of the nanoparticles and launches coherent breathing oscillations. 1. Introduction For decades, since the invention of femtosecond spectroscopy techniques, scientists have been studying thermalization of electrons excited by laser pulses in metal nanoparticles and thin films. 1 By the end of the 1990s, mechanical oscillations at the picosecond time scale after laser illumination were observed for the first time in semiconductors 2 followed by silver 3 and gold 4 nanoparticles, evidenced by a transmission signal change in pumpprobe measurements. These oscil- lations were quickly found to match the lowest frequency radial mode (or breathing mode) of the particles, modeled as an elastic sphere by classical calculations. Their frequency is inversely proportional to the nanoparticle radius. 5,6 Similar be- havior was observed in tin and gallium nanoparticles. 2 This discovery paved the way to more research in electronphonon dynamics in metal nanoparticles in the following years. 1,714 In particular, experimental studies for monolayer-protected nanoparticles also confirmed relaxation of electronic excitation to vibrational modes, although dierences in features such us excited state lifetime and vibrational periods with respect to uncapped nanoparticles were found. 12,15,16 The proposed explanation, supported by experimental evidence, considers that electrons thermalize in the hundred femtosecond timescale, and then transfer heat to phonons faster than the period of the breathing mode (typically in the order of 12 ps); the consequent increase in lattice tempera- ture increases suddenly the equilibrium radius of the particle, inducing atoms to oscillate around the new potential energy minimum. 5,6,9 Experimental evidence in uncapped nano- particles is consistent with this mechanism, since oscillations could not be observed for small particles (for which the characteristic oscillation time is shorter). 5 However, even though thermalization dynamics are clearly understood, the reason why the oscillations are launched is yet not very clear. In particular, investigations within the small size and short time regime are needed to prove that electronphonon coupling driving an indirect displacive mechanism is the main cause for the oscillations in all size and time regimes. In particular, hot-electron pressure has been reported as responsible for the mechanical vibrations in small particles and for the initial phase of the oscillations. 10,1720 Computational studies have helped to explore the afore- mentioned conditions 2123 but have not yet been able to account for real-time laser induced phenomena. Here, we report the first real-time computational study that shows evidence of breathing oscillations in the sub-picosecond regime, via surface plasmon excitation. Since our method does not explicitly account for dissipation from electrons to Electronic supplementary information (ESI) available. See DOI: 10.1039/ C7NR04536K a INFIQC (CONICET - Universidad Nacional de Córdoba), Departamento de Química Teórica y Computacional, Facultad de Ciencias Químicas, Universidad Nacional de Córdoba, Ciudad Universitaria, 5000 Córdoba, Argentina. E-mail: [email protected] b Bremen Center for Computational Materials Science, Universität Bremen, Otto-Hahn-Alle 1, 28359 Bremen, Germany c Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing, 100190, P. R. China d Collaborative Innovation Center of Quantum Matter, Beijing, 100190, P. R. China This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 1239112397 | 12391 Published on 03 August 2017. Downloaded by Institute of Physics, CAS on 18/11/2017 04:03:03. View Article Online View Journal | View Issue

Upload: others

Post on 31-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

Nanoscale

PAPER

Cite this: Nanoscale, 2017, 9, 12391

Received 23rd June 2017,Accepted 28th July 2017

DOI: 10.1039/c7nr04536k

rsc.li/nanoscale

Plasmon-driven sub-picosecond breathing ofmetal nanoparticles†

Franco P. Bonafé, a Bálint Aradi,b Mengxue Guan, c

Oscar A. Douglas-Gallardo, a Chao Lian, c Sheng Meng,c,d

Thomas Frauenheimb and Cristián G. Sánchez *a

We present the first real-time atomistic simulation on the quantum dynamics of icosahedral silver nano-

particles under strong laser pulses, using time dependent density functional theory (TDDFT) molecular

dynamics. We identify the emergence of sub-picosecond breathing-like radial oscillations starting

immediately after laser pulse excitation, with increasing amplitude as the field intensity increases. The

ultrafast dynamic response of nanoparticles to laser excitation points to a new mechanism other than

equilibrium electron–phonon scattering previously assumed, which takes a much longer timescale.

A sharp weakening of all bonds during laser excitation is observed, thanks to plasmon damping into

excited electrons in anti-bonding states. This sudden weakening of bonds leads to a uniform expansion of

the nanoparticles and launches coherent breathing oscillations.

1. Introduction

For decades, since the invention of femtosecond spectroscopytechniques, scientists have been studying thermalization ofelectrons excited by laser pulses in metal nanoparticles andthin films.1 By the end of the 1990s, mechanical oscillations atthe picosecond time scale after laser illumination wereobserved for the first time in semiconductors2 followed bysilver3 and gold4 nanoparticles, evidenced by a transmissionsignal change in pump–probe measurements. These oscil-lations were quickly found to match the lowest frequencyradial mode (or breathing mode) of the particles, modeled asan elastic sphere by classical calculations. Their frequency isinversely proportional to the nanoparticle radius.5,6 Similar be-havior was observed in tin and gallium nanoparticles.2 Thisdiscovery paved the way to more research in electron–phonondynamics in metal nanoparticles in the following years.1,7–14

In particular, experimental studies for monolayer-protectednanoparticles also confirmed relaxation of electronic excitation

to vibrational modes, although differences in features such usexcited state lifetime and vibrational periods with respect touncapped nanoparticles were found.12,15,16

The proposed explanation, supported by experimentalevidence, considers that electrons thermalize in the hundredfemtosecond timescale, and then transfer heat to phononsfaster than the period of the breathing mode (typically in theorder of 1–2 ps); the consequent increase in lattice tempera-ture increases suddenly the equilibrium radius of the particle,inducing atoms to oscillate around the new potential energyminimum.5,6,9 Experimental evidence in uncapped nano-particles is consistent with this mechanism, since oscillationscould not be observed for small particles (for which thecharacteristic oscillation time is shorter).5

However, even though thermalization dynamics are clearlyunderstood, the reason why the oscillations are launched is yetnot very clear. In particular, investigations within the smallsize and short time regime are needed to prove that electron–phonon coupling driving an indirect displacive mechanism isthe main cause for the oscillations in all size and timeregimes. In particular, hot-electron pressure has been reportedas responsible for the mechanical vibrations in small particlesand for the initial phase of the oscillations.10,17–20

Computational studies have helped to explore the afore-mentioned conditions21–23 but have not yet been able toaccount for real-time laser induced phenomena.

Here, we report the first real-time computational study thatshows evidence of breathing oscillations in the sub-picosecondregime, via surface plasmon excitation. Since our method doesnot explicitly account for dissipation from electrons to

†Electronic supplementary information (ESI) available. See DOI: 10.1039/C7NR04536K

aINFIQC (CONICET - Universidad Nacional de Córdoba), Departamento de Química

Teórica y Computacional, Facultad de Ciencias Químicas, Universidad Nacional de

Córdoba, Ciudad Universitaria, 5000 Córdoba, Argentina.

E-mail: [email protected] Center for Computational Materials Science, Universität Bremen,

Otto-Hahn-Alle 1, 28359 Bremen, GermanycBeijing National Laboratory for Condensed Matter Physics and Institute of Physics,

Chinese Academy of Sciences, Beijing, 100190, P. R. ChinadCollaborative Innovation Center of Quantum Matter, Beijing, 100190, P. R. China

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 12391–12397 | 12391

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article OnlineView Journal | View Issue

Page 2: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

phonons, the breathing oscillations we see occur as a quasi-adiabatic process, in the sense that no direct dissipative elec-tron–phonon interaction is included in the model. An expla-nation for the observed results is proposed based on a dis-placive excitation mechanism. The use of a density functionaltight binding method for larger particles is proved to be validby comparing with the results obtained by time-dependentdensity functional theory.

2. Computational methods2.1. Time-dependent density functional theory

Time-dependent density functional theory (TDDFT) calcu-lations were performed using the time-dependent ab initiopackage (TDAP) as implemented in SIESTA.24 This method hasbeen used recently to study dynamics after excitation in severalsystems.25–27 Numerical atomic orbitals with double zeta polar-ization (DZP) were used as the basis set, while the electron–nuclear interactions were described by Troullier–Martinspseudopotentials in the local density approximation (LDA) anda real-space grid equivalent to a plane-wave cutoff of 250 Rywas adopted.

Structures were converged using a standard conjugate-gradi-ent technique until an energy tolerance of 10–6 eV was reached.Forces on each atom were found to be less than 0.01 eV Å−1.

Absorption spectra were computed from real-time propa-gation of the density matrix (DM) ρ of the system after aninitial step perturbation (eqn (1)) to the ground state DM. Thefrequency-dependent polarizability tensor α is obtained withinthe linear response regime by deconvoluting the dipolemoment signal in the time domain, leading to the absorptioncross-section which is directly proportional to the imaginarypart of the polarizability:

EνðtÞ ¼ E0νð1� ΘðtÞÞ: ð1Þ

Non-adiabatic molecular dynamics simulations were per-formed integrating the time-dependent Kohn–Sham equationsas described by Meng and Kaxiras.24 We used a time step of0.05 fs and reached a total simulation time of 800 fs. The laserpulse was described as shown in eqn (2), where the frequencyωp matches the plasmon energy (3.8 eV). The pulse peak timewas set at t0 = 20 fs while the width was chosen to be σ = 6 fs.E0 was set to 0.5 V Å−1:

EðtÞ ¼ E0 cosðωptÞexpð�ðt� t0Þ2=2σ 2Þ: ð2ÞThis highly accurate method could be used to describe the

time-dependent behavior of particles of up to 55 atoms, due toits high computational cost. For larger systems we used themore approximate time-dependent density functional tight-binding method.

2.2. Time-dependent density functional tight-binding

The electronic description of the ground state is achievedusing the self-consistent charge density functional tight-

binding (DFTB) approach, as implemented in the DFTB+package.28 This method is based on a second order expansionof the Kohn–Sham functional around a reference density andhas proven useful for accurate descriptions of the properties ofmetallic nanostructures, semiconductors and large molecules.The Slater–Koster parameters were obtained using the Hyb-0-2parameter set.29,30 The conjugate-gradient method was used tooptimize the structures.

The methodology used to calculate the absorption spectrais similar to the one described in section 2.1, using in thiscase a Dirac-delta-like perturbation. The Liouville–vonNeumann equation for non-orthogonal basis (eqn (3)) is thennumerically integrated using the Leapfrog algorithm. Ineqn (3) Sij = ⟨ϕi|ϕj⟩ is the overlap matrix, H is the Hamiltonianmatrix and i the imaginary unit:

ρ̇ ¼ �iðS�1H½ρ�ρ� ρH½ρ�S�1Þ: ð3Þ

In this case, the polarizability tensor can be calculated byFourier-transforming the dipole moment signal in the timedomain and dividing by the Dirac-delta-like pulse field inten-sity (eqn (4)). A damping factor of 0.1 fs−1 was used to obtainuniform broadening of the peaks. This choice of damping isrelated to the length of the trajectory used as the input in eqn(4) and is only used to obtain a well-resolved plasmon peakand to identify its energy precisely. This methodology hasbeen applied to study the optical properties of a range of mole-cular and nanostructured systems:31–34

αðωÞ ¼ μðωÞE0

: ð4Þ

For the systems under analysis in the present paper, we pro-pagated 20 000 steps using a time step of 0.0048 fs and an elec-tric field intensity of E0 = 0.001 V Å−1, well within the linearresponse regime. Plasmon energies were identified as the lowest-energy peak in the absorption spectra as described elsewhere.34

Non-adiabatic molecular dynamics were performed withinthe Ehrenfest approximation, in which electrons are treated asquantum and nuclei as classical particles that move in a meanfield caused by the instantaneous electronic wave-function. Inthis approximation, non-adiabaticity is described only approxi-mately, and inelastic dissipation from electrons to phonons isneglected.35

This is due to the neglect of quantum fluctuations of theposition and momentum within the Ehrenfest approximation.This approximation can be seen as the first term of a momentexpansion of the coupled equations of motion of the electrondensity matrix and the position and momentum operatorprobability distributions around their average values.36 It canbe shown that spontaneous phonon emission and the descrip-tion of processes such as inelastic current–voltage spec-troscopy (involving the proper description of electron–phononscattering) requires the expansion to be carried out up to thesecond moment of the momentum and position distri-butions.37 In the results shown in this paper this drawback isused as an advantage, as it is discussed in section 3.

Paper Nanoscale

12392 | Nanoscale, 2017, 9, 12391–12397 This journal is © The Royal Society of Chemistry 2017

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online

Page 3: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

The formalism described as follows was derived from for-mulations reported in the literature38,39 and implemented inthe DFTB+ package. The equation of motion of the DM issimilar to the Liouville–von Neumann equation but includingnon-adiabatic coupling terms (eqn (5)):

ρ̇ ¼ �iðS�1H½ρ�ρ� ρH½ρ�S�1Þ � ðS�1Dρþ ρD†S�1Þ ð5Þwhere DkBlA = vA·⟨ϕkB|∇AϕlA⟩ is the non-adiabatic couplingmatrix, vA being the velocity of atom A and ∇AϕlA the gradientof the orbital ϕl (which is localized on atom A) with respect tothe coordinates of A. In our implementation, we approximated⟨ϕkB|∇AϕlA⟩ as the gradient of the overlap matrix with respectto the AB interatomic vector, RAB, for A ≠ B, and 0 otherwise(see eqn (6), where δAB is Kronecker’s delta function):

hϕkB j∇AϕlA i ¼ ∇ABSkBlA ð1� δABÞ: ð6ÞAn external laser field can be applied by adding a time-

dependent term to the Hamiltonian of the form V = −μ(t )·E(t ).For a sin2-shaped pulse, the electric field is determined byusing eqn (7), where ωp matches the plasmon energy for eachcase, and the total duration of the pulse was set to τ = 25 fs forall simulations. We varied E0 exploring the range between0.001 V Å−1 and 0.75 V Å−1. Here Θ(t ) is the Heaviside stepfunction. The pulse envelope function is different from what isshown in section 2.1, but its effect in the dynamics is minor,affecting only its amplitude (see the ESI†):

EðtÞ ¼ E0 sinðωptÞsin2ðπt=τÞð1� ΘðτÞÞ: ð7ÞTime-dependent forces are calculated at each electronic

step and are used to propagate the nuclei simultaneouslyusing the velocity Verlet algorithm. The expression for theforces can be found in the ESI.† Phonon frequencies and dis-placement vectors were calculated using the modes program ofthe same package (see below for explanation of the method).

2.3. Normal mode analysis

The mass-weighted Hessian matrix F, defined in eqn (8), iscomputed by numerical differentiation using either self-con-sistent DFT or DFTB for the energy calculation. Within the har-monic approximation, the normal modes frequencies and thenormalized displacement vectors are the eigenvalues andeigenvectors of F, respectively:

Fij ¼ 1ffiffiffiffiffiffiffiffiffiffiffimimj

p d2ε

dqjdqi: ð8Þ

The eigenvectors are then projected along the radial direc-tion, and the modes with the highest projection are analyzedto find the fundamental radial mode, hereafter called thequasi-breathing mode (QBM), since the pure breathing modeexists only up to n = 13.40

2.4. Model systems

Icosahedral silver nanoparticles (NPs) of three different sizes(number of atoms n = 55, 147, 309) were studied in this work.

Their diameter and plasmon energy as calculated by TDDFTBare summarized in Table 1.

3. Results and discussion

The 55-atom NP, after geometry optimization and having zeroinitial velocities, was illuminated with laser pulses of intensityE0 = 0.5 V Å−1 as described in sections 2.1 and 2.2. The systemevolves freely after this initial perturbation for 800 fs. The NPaverage radius, calculated by averaging the distance of allsurface atoms to the central atom, is shown as a function oftime in Fig. 1 for both methods. An oscillatory behavior isobserved for both simulations, although similarities are onlyqualitative. There is an evident mismatch of the periodbetween the two methods, being ca. 300 fs for TDDFT and ca.390 fs for TDDFTB. By normal mode analysis with DFT andDFTB we found that the QBM periods for each method (304 fsand 393 fs, respectively) are strikingly similar to the onesfound by the TD calculations. The dashed lines in Fig. 1 are

Table 1 Number of atoms n, structure, diameter d and plasmon energyħωp of the silver nanoparticles under study

n 55 147 309

d/nm 1.09 1.64 2.19ħωp/eV 3.21 2.95 2.85

Fig. 1 Radius (solid lines) as a function of time from TDDFT (top) andTDDFTB (bottom). Sine functions showing the calculated QBM areshown (dashed lines). Calculations were performed on an n = 55 Agnanoparticle using a 25 fs laser pulse, with E0 = 0.5 V Å−1.

Nanoscale Paper

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 12391–12397 | 12393

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online

Page 4: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

sin functions with a period equal to the QBM. This confirmedthat the QBM is selectively excited in the system under ana-lysis. Different periods can be attributed to the different hard-nesses of the potential energy surface that is obtained usingthe two approaches. The same fact can account for the differ-ence in amplitude.

TDDFT’s higher accuracy is due to a larger basis set and abetter description of the electronic interaction throughexchange and correlation effects, and it is an excellent tool tostudy these processes. However, for practical limitations onlythe n = 55 NP could be studied using this method. In the pre-vious discussion, we have validated TDDFTB as a validapproach to study this process, with quantitative differences inthe results when compared to TDDFT. As we were mostly inter-ested in trends and elucidating the general mechanism ratherthan reproducing exact results, we used TDDFTB for theremaining part of this work.

Molecular dynamics on the three NPs described in section2.4 were run after illumination with 25 fs-long laser pulses asdescribed in section 2.2, in resonance with plasmon excitation(identified as the lowest-energy peak in the absorption spectra,see the ESI†). The fact that the laser pulse is shorter than inmost experimental studies (where 100–180 fs pulse widths arecommonly used) does not alter the validity of the results, sinceas long as the pulse is shorter than the period of the QBM (asit is explained below), the main effect of a longer pulse (usinga weaker field) would be to start the oscillations later in time,precluding us from studying an adequate time span due to itshigh computational cost.

The electronic energy (mostly the band component)increases quadratically with the laser field.41 In Fig. 2(a) thechanges of the average radii of the NPs are shown as timeevolves. For laser intensities higher than the threshold value of0.1 V Å−1 an oscillatory behavior is observed. The shaded areaindicates the length of the laser pulse. Even when the period isclearly dependent on the size (see below for detailed discus-sion), the amplitude of the change in the radius seems to beindependent of the NP size and radius (note that the scale inthe vertical axis is the same for each row). An interestingfeature is that vibrations start during the laser pulse, implyingthat they are directly launched as a consequence of theelectron excitation.

The change in the lattice temperature with these field inten-sities ranged from 0.15 K to 70 K, while higher temperaturesand oscillation amplitudes were found using higher fieldintensities.†

The observed oscillations show a large radial component byvisual inspection of the animations. To find out whether theQBM is excited, normal modes are computed within the DFTBframework as described in section 2.3 and the QMB period isplotted as a function of the NP diameter as shown in Fig. 2(b).A linear relationship is found, in agreement with the expectedtrend from atomistic and elastic sphere calculations accordingto Lamb’s theory,42 plotted in the same figure. The oscillationperiod observed in the molecular dynamics simulations is alsoshown in the same figure, evidencing good agreement with the

result obtained from the normal mode analysis, emphasizingthe fact that the QBM is excited by the laser pulse. Thesedifferences can be attributed to the anharmonicity in thepotential (discussed below) and excitation of other modes. InFig. 2(b) the QBM period calculated by Sauceda et al.21 bynormal mode analysis using the Gupta potential is alsoplotted for comparison. Even though there is an offsetbetween the trends of about 70 fs, the slopes are very similar.As it has been reported, the slope of the period vs. diameterrelationship is a characteristic of the QBM mode. The offsetcan then arise from systematic differences related to theemployed methodology (i.e. the DFTB method).

Given that our method does not account explicitly for elec-tron–phonon inelastic scattering, which would allow heattransfer to occur, this result is puzzling in light of the mechan-ism reported in the literature.

To understand the origin of such oscillations we investi-gated the band component of the pairwise bond energy for allpairs of atoms A ≠ B, hereafter denoted as EABpb (eqn (9)), as a

Fig. 2 (a) Change in the radius for three particle sizes (columns) andthree laser field intensities (rows). Shaded areas indicate the laser pulse.(b) Period of oscillation as a function of diameter calculated from conti-nuum (Lamb’s) theory (gray line), normal mode analysis within DFTB(blue dots), laser induced oscillations (red triangles) and normal modeanalysis using a Gupta potential, reproduced from Sauceda et al.21

(green squares).

Paper Nanoscale

12394 | Nanoscale, 2017, 9, 12391–12397 This journal is © The Royal Society of Chemistry 2017

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online

Page 5: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

function of time. This quantity is practically zero unless A andB are the nearest neighbors. The self-consistent charge (scc)and repulsive contributions to the energy28 are not consideredsince their variation is much smaller than the bandcontribution:

EABpb ¼

X

μ[A

X

ν[B

ρμνH0μν: ð9Þ

In Fig. 3 the change in Epb for all nearest-neighbor pairs ofatoms and its mean value are plotted for the n = 309 NP withE0 = 0.25 V Å−1. The inset shows the first 50 fs, where all bondenergies increase due to the excitation of electrons caused bythe laser pulse (25 fs). We conclude that such homogeneousweakening of all bonds launches the atomic oscillations.During the first 700 fs the mean bond energy oscillates accom-panying the nuclei but does not decrease to its original value,consistently with the fact that no electron–phonon dissipationmechanism is considered in the simulations. Fig. 3 also showswhich bonds weaken and strengthen the most (top andbottom NP diagrams embedded in the figure, respectively).

The proposed mechanistic picture is then given as follows:as electrons absorb energy they populate the excited states,which are of antibonding character (see Fig. 4 which evidenceselectrons being promoted from the occupied states of the spand d bands to the antibonding states). These excitations arehighly delocalized all over the particle due to the delocalizednature of the plasmon excitation, uniformly weakening allbonds. As a consequence of this, the potential energy surfacechanges (bonds become looser) and the equilibrium radius R0

of the particle increases. The initial expansion is isotropic, and

when the particle reaches R0 it overshoots by inertia of theatoms, contracting afterwards due to the attractive componentof the force, which results in an impulsive excitation of thequasi-breathing mode.

The described mechanism is very similar to the displaciveexcitation of coherent phonons previously reported,43 whichconsiders excited electrons as the driving force for the initialexpansion instead of fast heat transfer from the electrons tothe lattice. This result is also in agreement with the calcu-lations by Lethiec et al.,23 where it was found that, for tetra-hedral silver clusters, breathing mode excitation arises frominternal conversion after excitation rather than from verticalexcitation itself.

This kind of impulsive nuclear motion induced by elec-tronic excitation is caused by the sudden change in the expec-tation value of the force at the expectation value of the posi-tion, a process that is correctly described by the first orderterm in the expansion mentioned in section 2.2.36,37 This isthe case for other impulsive processes such as the induced dis-sociation of molecules described in ref. 25 and 26. This is thereason why displacive excitation is observed in the simulationsdespite that no direct electron–phonon coupling (i.e. quantumscattering) can occur within the used simulation scheme.

Several studies have reported features in the vibrations thatare compatible with the displacive excitation mechanism.Direct coupling processes such as hot-electron pressure show asine response with amplitude increasing with the frequency ofthe mode,9 which is larger for smaller particles. Experimentson non-spherical particles with anisotropic acoustic vibrationshave shown that hot-electron pressure is an important contri-bution to thermal expansion17 and can even excite a particularmode.18

Further support of this mechanism comes from measure-ments in monolayer-protected clusters. High-frequencyvibrations observed in gold clusters of the 1.1–4 nm size rangecapped with hexane thiolate were explained using thisapproach.12 However, unlike our results, the vibrational fre-

Fig. 3 Change in pairwise bond energy for all first-neighbor bonds inthe n = 309 NP as function of time. The inset shows a magnification ofthe first 50 fs of the main plot. The bonds that correspond to somecurves in the 200–300 fs range are shown in the figure.

Fig. 4 Change in the population of states as a function of state energyand time. The density of states is plotted on the left panel. The dashedvertical line at 25 fs indicates the end of the laser pulse.

Nanoscale Paper

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 12391–12397 | 12395

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online

Page 6: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

quency was found to be size-independent and compatible witha local mode of the core gold atoms, not the breathing modeof the whole cluster. This effect may be explained consideringthat ligands induce strong surface disorder,44,45 leaving a crys-talline core (at least for NPs larger than 2 nm) leading to theaforementioned feature which has also been seen in other rele-vant studies.15,16

The validity of the displacive excitation mechanism wasanalized by Hodak et al.5 and did not succeed in explainingtheir experimental results since (i) the lifetime of excited elec-trons in metals is too short and (ii) it would lead to higheramplitudes as the field intensity increases, while the oppositebehavior is observed. The reason given for (ii) is that for highfields/small particles the heat transfer to the lattice is slowerthan the QBM period, leading to non-coherent excitation whichprecludes its experimental observation in ensemble experi-ments. However, for our modeled scenario these arguments areless important since (i) we consider shorter times than the elec-tron–phonon coupling times, so most electrons remain excited(only redistribution of energy in the electronic subsystemoccurs†) and (ii) we look at a single particle and hence coherentexcitation in a polydisperse sample is not considered.

With this picture in mind, simple scaling laws relating tothe parameters of the oscillation with the external pertur-bation can be obtained. The radius vs. time plots as thoseshown in Fig. 2 were fitted to a sine function R = A sin(ωt + ϕ)+ R0 using standard non-linear least-squares techniques. Fivesimulations for each particle size, with laser field intensitiesranging from 0.01 V Å−1 to 0.25 V Å−1 were considered here(their radius vs. time profile and fittings can be found in theESI†). The frequency ω and phase ϕ show little dependence onthe field intensity, since the QBM is excited in all cases exceptfor small fields.

Some general scaling laws can be obtained from the picturedescribed in the preceding paragraph. Both the new equili-brium radius R0 and amplitude A increase proportionally withthe external field E0, but they remain almost independent ofthe particle size.† Better insight can be drawn if the absorbedelectronic energy, ΔEel, is considered instead of E0. SinceΔEel is proportional to the absorbed power, it is directlyproportional to the number of atoms n and quadraticallyproportional to E0 (since the cross section increases withvolume32 (∝n) and the instantaneous power absorbed fromthe external field is proportional to its square41).Also ΔEel �

PA

PBΔEAB

pb , therefore, ΔEel/n is a good approxi-

mation to the increase of bond energy per unit volume. InFig. 5(a) the amplitude is plotted vs. ΔEel/n for all particle sizes. Alinear trend is found where slopes are all similar regardless ofthe particle size. Hence, this relationship can help predict theoscillation amplitude as a function of field intensity.

The change in the equilibrium radius ΔR0 also follows asimilar trend, as plotted in Fig. 5(b), but considering ΔEel/n2/3

as the abscissa. This is explained considering that the averageincrease of the particle radius due to the excited electrons hasto be proportional to the radius itself. Hence, to the volume-

normalized weakening of bonds ΔEel/n should be multipliedby n1/3 (since R ∝ n1/3) to account for the previous conditions.This leads to ΔEel/n × n1/3 = ΔEel/n2/3.

These simple scaling laws nicely illustrate the proposedpicture of how the plasmon excitation drives the displaciveQBM excitation in our simulations.

4. Conclusions

We have proved for the first time that mechanical oscillationscompatible with the quasi-breathing mode can be launched insmall silver nanoparticles even “turning off” electron–phononinteractions. Hence, the mechanism in this case is based on aquasi-adiabatic process in which excited electrons populatethe excited states of antibonding character, weakening thebonds uniformly and causing an expansion of the nano-particle, which then vibrates around a new equilibrium radius.New scaling laws can be derived from this explanation, inwhich the amplitude is proportional to the absorbed electronicenergy normalized by the number of atoms n and the changein the equilibrium radius follows the same trend using n2/3 asthe normalization factor. These conclusions have been vali-dated by highly accurate TDDFT calculations, despite differ-ences observed due to the hardness of the oscillator thatemerges from the different approaches.

Our results imply that a different, quasi-adiabatic mechan-ism, might be playing a role in the short-time small-sizeregime where electron–phonon scattering is not relevant dueto its longer time scale.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

The authors acknowledge financial support from ConsejoNacional de Investigaciones Científicas y Técnicas (CONICET)

Fig. 5 Scaling of the (a) amplitude and (b) change in the equilibriumradius of the particles as a function of absorbed energy normalized bythe (a) volume and (b) volume/radius.

Paper Nanoscale

12396 | Nanoscale, 2017, 9, 12391–12397 This journal is © The Royal Society of Chemistry 2017

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online

Page 7: Plasmon-driven sub-picosecond breathing of metal nanoparticleseverest.iphy.ac.cn/papers/nanoscale9.12391.pdf · 2.2. Time-dependent density functional tight-binding The electronic

through grant PIP 112-201101-0092 and wish to thank SeCyTUNC for the funding received. FPB is grateful for a studentshipfrom CONICET. The authors thank Huziel Sauceda for sharingdata from a previous published work21 to be compared withour results, and the anonymous reviewers for their valuablesuggestions. This work used computational resources from theUniversität Bremen computing clusters and cloud computingprovided by Quantum Dynamics.

References

1 G. V. Hartland, Chem. Rev., 2011, 111, 3858–3887.2 M. Nisoli, S. De Silvestri, A. Cavalleri, A. M. Malvezzi,

A. Stella, G. Lanzani, P. Cheyssac and R. Kofman, Phys. Rev.B: Solid State, 1997, 55, R13424–R13427.

3 N. Del Fatti, S. Tzortzakis, C. Voisin, C. Flytzanis andF. Vallee, Phys. B, 1999, 263–264, 54–56.

4 J. H. Hodak, I. Martini and G. V. Hartland, J. Chem. Phys.,1998, 108, 9210–9213.

5 J. H. Hodak, A. Henglein and G. V. Hartland, J. Chem.Phys., 1999, 111, 8613–8621.

6 N. Del Fatti, C. Voisin, F. Chevy, F. Vallee, C. Flytzanis andF. Vallée, J. Chem. Phys., 1999, 110, 11484.

7 N. Del Fatti, C. Flytzanis and F. Vallee, Appl. Phys. B: LasersOpt., 1999, 68, 433–437.

8 J. H. Hodak, A. Henglein and G. V. Hartland, J. Phys. Chem.B, 2000, 104, 9954–9965.

9 C. Voisin, N. Del Fatti, D. Christofilos and F. Vallee, J. Phys.Chem. B, 2001, 105, 2264–2280.

10 G. V. Hartland, J. Chem. Phys., 2002, 116, 8048–8055.

11 A. Nelet, A. Crut, A. Arbouet, N. Del Fatti, F. Vallee,H. Portales, L. Saviot and E. Duval, Appl. Surf. Sci., 2004,226, 209–215.

12 O. Varnavski, G. Ramakrishna, J. Kim, D. Lee andT. Goodson, ACS Nano, 2010, 4, 3406–3412.

13 V. Juve, A. Crut, P. Maioli, M. Pellarin, M. Broyer,N. Del Fatti and F. Vallee, Nano Lett., 2010, 10, 1853–1858.

14 S. Hashimoto, D. Werner and T. Uwada, J. Photochem.Photobiol., C, 2012, 13, 28–54.

15 S. A. Miller, J. M. Womick, J. F. Parker, R. W. Murray andA. M. Moran, J. Phys. Chem. C, 2009, 113, 9440–9444.

16 H. Qian, M. Y. Sfeir and R. Jin, J. Phys. Chem. C, 2010, 114,19935–19940.

17 M. Perner, S. Gresillon, J. März, G. von Plessen,J. Feldmann, J. Porstendorfer, K.-J. Berg and G. Berg, Phys.Rev. Lett., 2000, 85, 792–795.

18 L. Bonacina, A. Callegari, C. Bonati, F. V. Mourik andM. Chergui, Nano Lett., 2006, 6, 7–10.

19 S. Link and M. A. El-Sayed, Annu. Rev. Phys. Chem., 2003,54, 331–366.

20 C. Voisin, N. Del Fatti, D. Christofilos and F. Vallée, Appl.Surf. Sci., 2000, 164, 131–139.

21 H. E. Sauceda, D. Mongin, P. Maioli, A. Crut, M. Pellarin,N. D. Fatti, F. Vallee and I. L. Garzón, J. Phys. Chem. C,2012, 116, 25147–25156.

22 A. J. Neukirch, Z. Guo and O. V. Prezhdo, J. Phys. Chem. C,2012, 116, 15034–15040.

23 C. M. Lethiec, L. R. Madison and G. C. Schatz, J. Phys.Chem. C, 2016, 120, 20572–20578.

24 S. Meng and E. Kaxiras, J. Chem. Phys., 2008, 129,054110.

25 L. Yan, F. Wang and S. Meng, ACS Nano, 2016, 10, 5452–5458.

26 L. Yan, Z. Ding, P. Song, F. Wang and S. Meng, Appl. Phys.Lett., 2015, 107, 083102.

27 C. Lian, S. B. Zhang and S. Meng, Phys. Rev. B: Condens.Matter Mater. Phys., 2016, 94, 1–7.

28 B. Aradi, B. Hourahine and T. Frauenheim, J. Phys. Chem.A, 2007, 111, 5678–5684.

29 B. Szucs, Z. Hajnal, T. Frauenheim, C. Gonzalez, J. Ortega,R. Perez, F. Flores and S. Sanna, Appl. Surf. Sci., 2003,212–213, 861–865.

30 B. Szucs, Z. Hajnal, R. Scholz, S. Sanna and T. Frauenheim,Appl. Surf. Sci., 2004, 234, 173–177.

31 C. F. a. Negre, V. C. Fuertes, M. B. Oviedo, F. Y. Oliva andC. G. Sanchez, J. Phys. Chem. C, 2012, 116, 14748–14753.

32 C. F. A. Negre and C. G. Sanchez, J. Chem. Phys., 2008, 129,034710.

33 M. B. Oviedo, C. F. A. Negre and C. G. Sanchez, Phys. Chem.Chem. Phys., 2010, 12, 6706.

34 O. A. Douglas-Gallardo, M. Berdakin and C. G. Sanchez,J. Phys. Chem. C, 2016, 120, 24389–24399.

35 A. P. Horsfield, D. R. Bowler, A. J. Fisher, T. N. Todorov andM. J. Montgomery, J. Phys.: Condens. Matter, 2004, 16,3609–3622.

36 A. P. Horsfield, D. R. Bowler, A. J. Fisher, T. N. Todorov andC. G. Sánchez, J. Phys.: Condens. Matter, 2004, 16, 8251–8266.

37 A. P. Horsfield, D. R. Bowler, A. J. Fisher, T. N. Todorov andC. G. Sánchez, J. Phys.: Condens. Matter, 2005, 17, 4793–4812.

38 T. N. Todorov, J. Phys.: Condens. Matter, 2001, 13, 10125–10148.

39 T. A. Niehaus, D. Heringer, B. Torralva and T. Frauenheim,Eur. Phys. J. D, 2005, 35, 467–477.

40 C. Henning, K. Fujioka, P. Ludwig, A. Piel, A. Melzer andM. Bonitz, Phys. Rev. Lett., 2008, 101, 1–4.

41 J. Posthumus, Molecules and clusters in intense laser fields,Cambridge University Press, 2001, p. 261.

42 H. Lamb, Proc. London Math. Soc., 1881, s1–13, 189–212.43 H. J. Zeiger, J. Vidal, T. K. Cheng, E. P. Ippen,

G. Dresselhaus and M. S. Dresselhaus, Phys. Rev. B:Condens. Matter, 1992, 45, 768–778.

44 M. M. Mariscal, J. a. Olmos-Asar, C. Gutierrez-Wing,A. Mayoral and M. J. Yacaman, Phys. Chem. Chem. Phys.,2010, 12, 11785.

45 J. a. Olmos-Asar, M. Ludueña and M. M. Mariscal, Phys.Chem. Chem. Phys., 2014, 16, 15979.

Nanoscale Paper

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 12391–12397 | 12397

Publ

ishe

d on

03

Aug

ust 2

017.

Dow

nloa

ded

by I

nstit

ute

of P

hysi

cs, C

AS

on 1

8/11

/201

7 04

:03:

03.

View Article Online