Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Aug 7.
Published in final edited form as: Phys Chem Chem Phys. 2013 Jun 19;15(29):12348–12359. doi: 10.1039/c3cp51117k

Broadband infrared and Raman probes of excited-state vibrational molecular dynamics; Simulation protocols based on loop diagrams

Konstantin E Dorfman *, Benjamin P Fingerhut , Shaul Mukamel
PMCID: PMC3744248  NIHMSID: NIHMS496992  PMID: 23783120

Abstract

Vibrational motions in electronically excited states can be observed by either time and frequency resolved infrared absorption or by off resonant stimulated Raman techniques. Multipoint correlation function expressions are derived for both signals. Three representations for the signal which suggest different simulation protocols are developed. These are based on the forward and the backward propagation of the wavefunction, sum over state expansion using an effective vibration Hamiltonian and a semiclassical treatment of a bath. We show that the effective temporal (Δt) and spectral (Δω) resolution of the techniques is not controlled solely by experimental knobs but also depends on the system dynamics being probed. The Fourier uncertainty ΔωΔt > 1 is never violated.

1 Introduction

The excited state vibrational dynamics of molecules plays a key role in many photophysical and photochemical processes and has attracted considerable experimental and theoretical attention14. Real time structural information about rearrangement of atoms in complex reactions can be inferred directly from time resolved vibrational spectroscopy57. Typically an ultrashort laser pulse in the visible or the UV excites the molecule to a bright valence excited state, launching a photoreaction or non-adiabatic relaxation process. The vibrational dynamics can then be probed either by the absorption of a delayed IR probe pulse813 or by a spontaneous or stimulated Raman process14.

Unique marker bands in UV pump/IR probe signals serve as fingerprint of the excited state evolution allowing to resolve transient reaction 2,15, intermediates structural details16 and reveal the reaction mechanism. Such investigations helped identify the real time reaction mechanism leading to the formation of photolesions in DNA nucleobases2,6,7,17, to monitor isomerization reactions in protein environments18, to resolve the consecutive steps in proton transfer reactions 15, to identify the participating ion pairs upon photoinduced bimolecular electron transfer19 and to follow light-induced electrocyclic reactions20,21. Frequency shifts of IR marker bands have also been used to monitor the response of the local environment2225 and molecular energy redistribution26. More elaborate pulse sequences allow to spread the IR signal in two dimensions, resolving the couplings between localized vibrations2729.

Spontaneous Raman 30 has long been used as an alternative probe. Recent stimulated Raman measurements that employ a femtosecond and a picosecond pulse had generated considerable excitement4,5,3141. A rich pattern of narrow (10 cm−1) vibrational lines has been reported in 25 fs intervals. Applications were made to pNA, pDNA42, rhodopsin1, carbon dioxide43, bacterial endospores 44, and other systems. Frequency domain stimulated Raman has proven validity in cell imaging45.

In this paper we focus on two techniques, both starting with an optical pump pulse but followed by a different detection: frequency - dispersed broadband infrared probe (FDIR) or off resonant stimulated Raman spectroscopy (SRS) 5. We show how both techniques can be described and interpreted with minor modifications using very similar vibrational correlation functions. The signals are intuitively described by loop diagrams which connect them to forward and backward propagation of the wavefunction.

We present a general analysis and derive closed expressions that can be used for microscopic quantum simulations of both infrared and Raman signals. Three representations for these correlation functions are presented each suggesting a different simulation strategy. The first is based on the numerical propagation of the wavefunction which includes all relevant electronic and nuclear (including bath) degrees of freedom explicitly. This is the most general, expensive and accurate method46,47. A second protocol uses a Sum Over States (SOS) expansion of the signals. Here we must diagonalize an effective vibrational hamiltonian. This offers a numerically more tractable algorithm when it is possible to truncate the relevant phase space. The third approach is semiclassical. A small vibrational system is treated quantum mechanically and a classical bath which causes a time dependent modulation of the system Hamiltonian is added. This is the simplest theory to implement by e.g. assuming that the vibrational frequencies change with time. This change can be either introduced phenomenologically or by using atomistic molecular dynamics simulations.

The time and frequency in these experiments are controlled by independent knobs. We can formally define uncertainties Δt and Δω associated with the pulse duration and the frequency resolution of a spectrometer. This suggests that there is no lower bound to the product ΔωΔt; the measurement can apparently be interpreted in terms of instantaneous snapshots with high spectral resolution. For example recent experiments1,5 use pulses (< 50 fs) and reported spectral features (< 10 cm−1) such that ΔωΔt ~ 0.5 ps cm−1 which is an order of magnitude smaller than the Fourier uncertainty for Gaussian pulses. An additional goal of this paper is to provide a proper definition of Δω and Δt and show that they are not purely instrumental but depend on the system as well. We find that the simple snapshot interpretation is false. We discuss the limitations of the spectral and temporal resolutions of these techniques and how they can be manipulated.

2 Loop diagram representation of frequency-dispersed stimulated signals

Stimulated optical signals are defined as the energy change of the electromagnetic field

S=-ddtE(t)E(t)dt. (1)

The radiation-matter interaction Hamiltonian in the rotating wave approximation (RWA) is

H(t)=V(t)E(t)+H.c., (2)

where V(t) + V(t) is a Heisenberg dipole operator and the electric field operator E(t) = ℰ(t) + ℰ(t). Both are separated into positive (non dagger) and negative (dagger) frequency components (lowering and raising photon operators, respectively). The dipole operator is given by the sum of the electronic and nuclear dipole moments V(t) = Ve(t) +Vn(t).

The Heisenberg equation of motion for the field operator E(t) then gives for the above integrated signal

S=2-dtIP(t)E(t)=2-dω2πIP(ω)E(ω), (3)

where where ℐ denotes the imaginary part,

P(ω)=-dtP(t)eiωt (4)

with P(t) = 〈V(t) representing the nonlinear polarization that arise from the interaction with the pump and the probe pulses. The angular brackets denote 〈…〉 = Tr[ρ (t) ] with the density operator ρ(t) defined in the joint field-matter space of the entire system. In practice, the temporal or spectral range of the integrations in Eq. (3) is restricted by the response function of the detector. If the detector contains a narrow time gate with nearly δ function response δ(t′t), Eq. (3) yields

STG(t;Γ)=2IP(t)E(t), (5)

where Γ denotes a set of parameters that characterize the various laser pulses. Similarly if the detector consists of a spectrometer with narrow frequency response δ (ω′ω), we obtain the frequency-gated signal

SFG(ω;Γ)=2IP(ω)E(ω), (6)

Note that the two signals Eqs. (5) and (6) carry different information and are not related by a simple Fourier transform. A Wigner spectrogram representation4850 was used in51 for the integrated pump probe signals Eq. (3). Here we use loop diagrams to describe the more detailed frequency- or time-gated signals (6) and (5), respectively. For clarity in the following we focus on the frequency-gated expressions, the corresponding time-gated signals are given in the Appendix A of the Supplementary Information.

3 First protocol; Numerical propagation of the wave function

We start with the visible-pump/IR-probe signal as sketched in Fig. 1a. FDIR is somewhat simpler than SRS since it only involves four rather than six radiation-matter interactions. The pump pulse centered at time τ3 = 0 promotes the system from its ground electronic state g to the vibrational state a of an excited electronic state and launches the vibrational dynamics. The IR probe pulse centered around τ3 = T can then either stimulate emission that couples the vibrational state a and lower vibrational state d or an absorption to higher vibrational state c. The signal is defined as the change in probe intensity and is either time-gated (Eq. (5) or frequency-gated (Eq. (6)). Both can be represented by the loop diagrams shown in Fig. 1b which contain the four field-matter interactions -two with each pulse. Diagram rules are given in Ref.52. Note that the signals (5) and (6) are expressed in terms of ℰ which makes the arrow corresponding the the last interaction pointing to the left. We further choose the last interaction to occur on the left branch. This choice removes any ambiguity in the diagrams rules without loss of generality.

Fig. 1.

Fig. 1

(Color online) FDIR: Level scheme - (a), closed-time path-loop diagrams - (b).

The electric field operator consists of the pump field 1 and a probe field 2

E(t)=E1(t)+E2(t-T), (7)

where T represents the delay of the probe pulse relative to the pump. The signal is given by the two loop diagrams shown in Fig. 1b, plus their complex conjugates. These give for the frequency-gated signal (6)

SIR(ω,T)=I-dΔ2πE2(ω)E2(ω+Δ)SIR(ω,T;Δ), (8)

where the Δ-dispersed signal is given by the two diagrams SIR(ω,T;Δ)=SIR(i)(ω,T;Δ)+SIR(ii)(ω,T;Δ) and

SIR(i)(ω,T;Δ)=2-dt-tdτ1-tdτ3-τ3dτ5E1(τ5)E1(τ1)×VeG(τ3,τ5)VnG(t,τ3)VnG(t,τ1)Veeiω(t-τ3)-iΔ(τ3-T), (9)
SIR(ii)(ω,T;Δ)=2-dt-tdτ1-tdτ3-τ3dτ5E1(τ5)E1(τ1)×VeG(t,τ1)VnG(t,τ3)VnG(τ3,τ5)Veeiω(t-τ3)-iΔ(τ3-T). (10)

(ω, T; Δ) represents the contribution of the ω and ω + Δ frequency components of ℰ2 to the signal, where ω is the detected frequency. The signal is obtained by integration over Δ. Here G(t1,t2) = (−i/ℏ)θ (t1 − t2)e−iH(t1−t2) is the retarded Green’s function. Even though this Green’s function only depends on the difference of its two time arguments, we retain both arguments and write G(t1,t2) rather than G(t1 − t2). This is done since in the reduced (semiclassical) description to be developed later when the system is coupled to some stochastic bath degrees of freedom, time translational invariance is lost and G then depends on both arguments. The corresponding time-gated signal (5) is given by Eqs. (S2) – (S3). Eqs. (9)(10) may be simplified further when the first pulse is impulsive. We can then set ℰ1(τ) = ℰ1δ (τ) and the τ1 and τ5 integrations can be eliminated.

Diagram (i) (Eq. (9)) can be understood using a forward and backward time evolving vibronic wave packet. First, the pulse ℰ1 electronically excites the molecule via Ve. The wave-function then propagates forward in time from τ1 to t. Then the IR probe pulse ℰ2 deexcites the vibrational transition to the lower vibrational level via Vn which then propagates back-ward in time from t to τ3. Pulse ℰ2 excites the vibration via Vn and the wavefunction propagates backward in time from τ3 to τ5. The final deexcitation by pulse ℰ1 returns the system to its initial state by acting with Ve. Diagram ii Eq. (10) can be interpreted similarly. Following initial electronic excitation the wavefunction propagates forward in time from τ5 to τ3. At this point a vibrational excitation promotes it to the higher vibrational state and the wavefunction propagates forward in time from τ3 to t. After vibrational deexcitation it then propagates backward from t to τ1 where an electronic excitation brings the system back in its initial ground state.

In Eqs. (9)(10) the matter correlation function is given in the time domain. Alternatively one can read the signal (8) from the diagrams when both field and matter correlation functions are given in the frequency domain

SIR(i)(ω,T)=I4π-dω2πdω12πdω12π×E2(ω)E2(ω)E1(ω1)E1(ω1)δ(ω-ω+ω1-ω1)×VeG(ωg+ω1)VnG(ωg+ω1-ω)VnG(ωg+ω1)Ve, (11)
SIR(ii)(ω,T)=I4π-dω2πdω12πdω12π×E2(ω)E2(ω)E1(ω1)E1(ω1)δ(ω-ω+ω1-ω1)×VeG(ωg+ω1)VnG(ωg+ω+ω1)VnG(ωg+ω1)Ve. (12)

Here G(ω) = h−1/[ω − H/ℏ + ], δ(ω-ω+ω1-ω1) represents the energy conservation that follows from time translation symmetry of all four field-matter interactions. One can separate the preparation pulse ℰ1 and break the δ-function as follows

δ(ω-ω+ω1-ω1)=-dΔδ(ω-ω+Δ)δ(ω1-ω1-Δ), (13)

where Δ defines the spectral bandwidth of the incoming pulse which translates into the spectral bandwidth of the relevant matter degrees of freedom. Eqs. (11)(12) then yield

SIR(i)(ω,T)=I2-dΔdω1E2(ω)E2(ω+Δ)E1(ω1)E1(ω1-Δ)VeG(ωg+ω1)VnG(ωg+ω1-ω-Δ)VnG(ωg+ω1-Δ)Ve, (14)
SIR(ii)(ω,T)=I2-dΔdω1E2(ω)E2(ω+Δ)E1(ω1)E1(ω1-Δ)VeG(ωg+ω1)VnG(ωg+ω1+ω1)VnG(ωg+ω1-Δ)Ve. (15)

We now turn to the electronically off-resonant SRS signal shown in Fig. 2a,b, which is completely analogous to the FDIR signal. Even though these signals represent different physical processes and even involve different number of field-matter interactions they can be described using very similar diagrams where we simply replace Vn → αn and ω by ω − ω3. In SRS the pump pulse initiates the vibrational dynamics in the excited electronic state. Pulse 3 and the probe then induce the Raman process (see. Fig. 2a). The relevant diagrams are shown in Fig. 2b (plus their complex conjugates). An electronically off-resonant Raman process induced by pulses 2 and 3 is instantaneous since by Heisenberg uncertainty the system can only spend a very short time in the intermediate state. The Raman process is thus described by an effective field/matter interaction Hamiltonian

Fig. 2.

Fig. 2

(Color online) SRS: Level scheme - (a), closed-time path-loop diagrams - (b).

H(t)=αnE2(t)E3(t)+E1(t)Ve(t)+H.c., (16)

where αn=αn+αn is the excited state polarizability that couples fields 2 and 3 parametrically via a Raman process. It is symmetric (real) operator. The time-domain signal (5) can be read directly from diagrams (i) and (ii). Assuming that pulse 3 is narrow band (picosecond) and set ℰ3(t − T) = ℰ3e−iω3(t−T). We obtain the frequency gated Raman analogues of Eqs. (9)(10)

SSRS(i)(ω-ω3,T;Δ)=2-dt-tdτ1-tdτ3-τ3dτ5×E32E1(τ5)E1(τ1)ei(ω-ω3)(t-τ3)-iΔ(τ3-T)×VeG(τ3,τ5)αnG(t,τ3)αnG(t,τ1)Ve, (17)
SSRS(ii)(ω-ω3,T;Δ)=2-dt-tdτ1-tdτ3-τ3dτ5E32E1(τ5)E1(τ1)ei(ω-ω3)(t-τ3)-iΔ(τ3-T)×VeG(t,τ1)αnG(t,τ3)αnG(τ3,τ5)Ve. (18)

The corresponding time-gated signals are given by Eqs. (S4) – (S5). Similarly to Eqs. (11)(12) we can recast (17) – (18) using frequency domain matter correlation functions

SSRS(i)(ω-ω3,T)=I4π-dω2πdω12πdω12π×E32E2(ω)E2(ω)E1(ω1)E1(ω1)δ(ω-ω+ω1-ω1)×VeG(ωg+ω1)αnG(ωg+ω1-ω+ω3)αnG(ωg+ω1)Ve, (19)
SSRS(ii)(ω-ω3,T)=I4π-dω2πdω12πdω12π×E32E2(ω)E2(ω)E1(ω1)E1(ω1)δ(ω-ω+ω1-ω1)×VeG(ωg+ω1)αnG(ωg+ω+ω1-ω3)αnG(ωg+ω1)Ve. (20)

Breaking up the δ-function according to Eq. (13) we get

SSRS(i)(ω-ω3,T)=I2-dΔdω1×E32E2(ω)E2(ω+Δ)E1(ω1)E1(ω1-Δ)×VeG(ωg+ω1)αnG(ωg+ω1-ω+ω3-Δ)αnG(ωg+ω1-Δ)Ve, (21)
SSRS(ii)(ω-ω3,T)=I2-dΔdω1×E32E2(ω)E2(ω+Δ)E1(ω1)E1(ω1-Δ)×VeG(ωg+ω1)αnG(ωg+ω1-ω3+ω1)αnG(ωg+ω1-Δ)Ve. (22)

The simulation protocol based on these equations requires the full forward (G) and backward (G) propagation of the wavefunction retaining all electronic and nuclear degrees of freedom. This task can be accomplished by numerically exact propagation techniques, based on the split-operator Fourier-transform, the short iterative Lanczos method or a Chebyshev expansion46,53, where the wavefunction is commonly expanded in the set of orthogonal eigenstates of H. Non-adiabatic effects can be conviniently be accounted for either in a diabtic or adiabatic basis of the participating electronic states54. The major drawback of this numerically exact treatment is that the computational effort and storage requirements grow exponentially with the number of degrees of freedom considered which limits their application to molecular systems with less than six degrees of freedom (4 atoms). The change to a nonorthogonal representations of the time-dependent wave-funtion allows to evaluate the Trotter expansion analytically and thus to avoid the unfavourable scaling behaviour which is accordingly not an intrinsic property of the powerful propagators47,55. The approximate multiconguration time-dependent Hartree (MCTDH) method 56 formally still scales exponentially but superior scaling and less memory requirements compared to the exact propagation methods can be achieved if the number of degrees of freedom and contraction coefficients is large. A major drawback of all propagation methods is that the global multi-dimensional potential energy surface has to be known a priori. Approximate direct quantum dynamical approaches like e.g. the variational multi-configuration Gaussian wavepacket method57 or ab initio multiple spawning 58,59 which rely on Gaussian functions as basis set circumvent this shortcoming as the potential energy surface is only sampled in the actual fraction of space where it is actually required. In some applications it may be desirable to only consider a few vibrational modes explicitly and treat the rest classically. Even in this case we may use the Green’s functions expressions (9) - (10) propagated forward and backward along the loop under an effective time dependent Hamiltonian60.

4 Second protocol; Sum Over States expansion

One can evaluate the matter correlation functions in Eqs. (9) -(10) by expanding them in the eigenstates of the total system. Again in this approach all bath degrees of freedom cannot be separated and must be included explicitly. The resulting SOS expansion provides useful insights and a convenient computational algorithm. Starting with Eqs. (9) - (10). A frequency-gated signal can be expressed

SIR(i)(ω,T)=-I2i4a,a,dμgaμagμadμade-(iωaa+γaa)Tω-ωad+iγad×E2(ω)E2(ω-ωaa+iγaa)E1(ωa+iγa)E1(ωa-iγa) (23)
SIR(ii)(ω,T)=-I2i4a,a,cμgaμagμacμace-(iωaa+γaa)Tω-ωac+iγac×E2(ω)E2(ω-ωaa+iγaa)E1(ωa+iγa)E1(ωa-iγa). (24)

The corresponding time-gated signals are given by Eqs. (S6) – (S7).

So far we had expanded the density operator starting with the ground state and including the preparation pulse ℰ1. Alternatively we can avoid the explicit treatment of preparation and simply assume that the system has been initially prepared in non stationary state represented by the density operator ρaa′. Elaborate pulse sequences can be used in this preparation. We can then evaluate the matter correlation function that corresponds to the last two interactions with the probe pulse for the i contribution:

V^nL(t)V^nR(τ3)=Tr[V^n(τ3)ρV^n(t)]=a,a,dρa,adV^n(t)aaV^n(τ3)d=a,a,dρaaμadμade-[iωad+γad]te[iωad+γd-γa]τ3, (25)

The Δ - dispersed signal (9) – (10) then yield

SIR(i)(ω,T;Δ)=-4πi2a,a,dμadμadρaaδ(Δ-ωaa+iγaa)eiΔTω+Δ-ωad+i(γd-γa), (26)
SIR(ii)(ω,T;Δ)=-4πi2a,a,cμacμacρaaδ(Δ-ωaa-iγaa)eiΔTω+Δ-ωac+i(γc-γa). (27)

The delta function arises from the time translation invariance of correlation functions: Δ = ωaa′ + aa′ that involves two frequencies of the probe field and the frequency band of the nonequilibrium preparation state aa′. Time translational invariance is maintained provided we treat the preparation explicitly via interaction with pulse ℰ1(ω1) and E1(ω1) as in Eqs. (23)(24). This implies that ω1-ω1+ω+Δ-ω=0. We can then write

δ(ω1-ω1-Δ)=-dω0δ(ω1-ω1-ω0)δ(ω0-Δ). (28)

The probe pulse by itself does not obey this symmetry as Δ ≠ 0. Thus, a description that excludes the preparation (actinic) pulse 1 does not have this symmetry. In this case, for a narrowband preparation pulse ω1ω1 results in Δ = 0 which means that the signal has low frequency resolution limited only by state lifetimes. The preparation pulse launches the vibrational dynamics, which results in high frequency resolution due to joint field plus matter bandwidth as shown in Eqs. (26)(27).

When both the pump and the probe pulses are ultrashort, i.e. ℰ1(τ) = ℰ1δ(τ) is centered at τ = 0 and ℰ2(τ) = ℰ2δ(τ − T) is centered at τ = T, we can neglect the frequency dispersion of the pulse envelopes. Eqs. (23)(24) then give

SIR(ω,T)=-I2i4a,aμgaμagE12E22×[dμadμade(iωaa-γaa)Tω-ωad+iγad+cμacμcae(iωaa-γaa)Tω-ωac+iγac]. (29)

One can derive similar SOS expressions for the frequency-gated SRS signals. Eqs. (23)(24) are then recast as

SSRS(i)(ω-ω3,T)=-I2i4a,a,dμgaμagαadαade-(iωaa+γaa)Tω-ω3-ωad+iγadE32E2(ω)×E2(ω-ωaa+iγaa)E1(ωa+iγa)E1(ωa-iγa), (30)
SSRS(ii)(ω-ω3,T)=-I2i4a,a,cμgaμagαacαace-(iωaa+γaa)Tω-ω3-ωac+iγacE32E2(ω)×E2(ω-ωaa+iγaa)E1(ωa+iγa)E1(ωa-iγa), (31)

As we did for FDIR, we shall express the signals (30) - (31) in a form that reveals the broken time translational symmetry. For the general pulse envelope of the pump field ℰ3(ω3) that enters twice in the signal, e.g. E3(ω3) and E3(ω3), the overall translational symmetry for all six interactions yields

δ(ω1-ω1+ω3-ω3-Δ)=-dω0δ(ω1-ω1-ω0)δ(ω0+ω3-ω3-Δ), (32)

where the product of two delta functions reveals the broken symmetry for the pump/probe fields when the preparation pulse is excluded. Assuming a narrowband pump ℰ3(t) = ℰ3e−iω3(t−T) Eqs. (26)(27)

SSRS(i)(ω-ω3,T;Δ)=-4πi2E32a,a,dαadαadρaaδ(Δ-ωaa+iγaa)eiΔTω-ω3+Δ-ωad+i(γd-γa), (33)
SSRS(ii)(ω-ω3,T;Δ)=-4πi2E32a,a,cαacαacρaaδ(Δ-ωaa+iγaa)eiΔTω-ω3+Δ-ωac+i(γc-γa). (34)

Finally for a broadband probe ℰ2 the SRS signal (30) – (31) reduce to

SSRS(ω-ω3,T)=-I2i4a,aμgaμagE12E22E32×[dαadαade(iωaa-γaa)Tω-ω3-ωad+iγad+cαacαcae(iωaa-γaa)Tω-ω3-ωac+iγac]. (35)

In the SOS protocol the basis set expansion has to cover the complete vibrational dynamics under investigation, which can be tedious for complex reactive systems and the diagonalization of the resulting Hamiltonian is non-trivial. Model Hamiltonians may be used to truncate the system size and provide an affordable simulation. For example exciton hamiltonians are commonly used to describe multiple excitations in chromophore aggregates61. Once the exact eigenstates are obtained, this protocol allows for the straightforward interpretation of the signals.

5 Third protocol; Coupling to a classical bath

A simpler and often more intuitive description can be developed by treating some (bath) degrees of freedom as classical. We start with the ultrafast visible pump and IR probe of the excited vibrational states. We assume that probe pulse is impulsive and set ℰ1(t) = ℰ1δ (t), and further evaluate the remaining time integrals using Eqs. (S10) and (S11). The resulting semi-classical Δ - dispersed signal (9) – (10) reads

SIR(ω,T;Δ)=-2i4-dτ3τ3dt×E12eiω(t-T)e-i(ω+Δ)(τ3-T)aμag2e-2γat×[cμac2e-iτ3tωac(t)dt+dμad2eiτ3tωad(t)dt], (36)

where ωα βα − ωβ|. Ensemble averaging 〈…〉e over the classical set of trajectories is performed on the signal level SIR(ω, T). Similarly one can derive the corresponding SRS result when the extra pump pulse is narrow band and can be approximated as ℰ3(t) = ℰ3e−iω3(t−T). The signal (17) – (18) then reads

SSRS(ω-ω3,T;Δ)=-2i4-dτ3τ3dt×E12E32ei(ω-ω3)(t-T)e-i(ω+Δ)(τ3-T)aμag2e-2γat×[cαac2e-iτ3tωac(t)dt+dαad2eiτ3tωad(t)dt]. (37)

Eqs. (36) and (37) involve a path integral over the stochastic vibrational frequency ωac(t) and ωad (t). The signal depends not only on the initial and final value of the vibrational frequency ωνν′, but rather on the entire pathway from time T to the time when the polarization decays to zero. The time dependent frequency ων,ν′(t) can be calculated by running classical MD trajectories.

In the semiclassical protocol of Eqs. (36)(37) the system is partitioned into a classical bath retaining only the quantum character of a few vibrational modes. For non-reactive systems (i.e. no chemical bonds are broken or formed) evolving on a single adiabatic potential energy surface (i.e. the Born-Oppenheimer approximation remains valid) common molecular dynamics simulations can be used which scale by N2 if all pair-wise electrostatic and van der Waals interactions are explicitly accounted for. The computational cost can be further reduced to linear scaling by suitable cutoffs. The quantum character of the vibrations under investigation can be retained by collective solvent coordinates which allow to map the classical dynamics onto ab initio derived electrostatic maps62,63.

If the process under investigation is characterized by ultra-fast relaxation in the vicinity of conical intersection as commonly observed in photoreactions the breakdown of the Born-Oppenheimer approximation requires to treat the system by non-adiabatic on-the-fly molecular dynamics64. Based on the independent trajectory approximation the nuclear wavepacket is represented by a swarm of independently evolving trajectories where, within the framework of Tullys fewest switches trajectory surface hopping65,66, relaxation between different electronic states is induced by the non-adiabatic coupling (NAC). Here the numerical effort of the dynamics is shifted to the calculation of excited state gradients and NACs between electronic states on an appropriate quantum chemical level but the construction of global potential energy surfaces is avoided as only the relevant configuration space is explored during the dynamics. The quantum character of vibrations is reconstructed by evaluating the excited state Hessian. The restriction on a few vibrational degrees of freedom allows for an efficient algorithm for the calculation of the semiclassical signal which is based on a mode tracking procedure67, only the desired frequencies and normal mode vectors are obtained. As the construction of the complete Hessian matrix is avoided linear scaling with the number of considered vibrational modes can be achieved68.

Rather than calculating the path integral numerically we can expand the integral in the exponent into the cumulant series and extract the mean ensemble averaged time dependent frequency ω̄νν′ (t) and approximate the remaining nuclear motion by harmonic Gaussian fluctuations. The signal calculated in Appendix B of the Supplementary Information may be then expressed in terms of the spectral density of the harmonic bath

SIR(ω,T;Δ)=-2i4-dτ3τ3dt×E12eiω(t-T)e-i(ω+Δ)(τ3-T)aμag2e-2γat×[cμac2e-iτ3tω¯ac(t)dt-gac(T,t)+dμad2eiτ3tω¯ad(t)dt-gad(T,t)], (38)
SSRS(ω-ω3,T;Δ)=-2i4-dτ3τ3dt×E12E32ei(ω-ω3)(t-T)e-i(ω+Δ)(τ3-T)aμag2e-2γat×[cαac2e-iτ3tω¯ac(t)dt-gac(T,t)+dαad2eiτ3tω¯ad(t)dt-gad(T,t)]. (39)

where the Gaussian fluctuations are manifested via the two point linewidth function gaj(T,t), j = c, d given by

gac(T,t)=4λacTβΛ+(2λacβΛ2-iλacΛ)×[e-Λt+(Λ(t-T)-1)e-ΛT], (40)

where λac represents the reorganization energy and Λ corresponds to the fluctuation time scale. Note that, the linewidth function depends on both initial and final times, not only the difference. This is a consequence of the non stationary vibrational dynamics.

6 The genuine temporal and spectral resolution of frequency-gated signals

The picture emerging from our theory is that the effective temporal and spectral resolution of FDIR and SRS signals is affected by, but not solely controlled by, experimental knobs. Achieving ultrafast resolution requires the active involvement of the entire probe bandwidth. This is eroded when a narrower slice of the pulse is selected by the system. Below we discuss how the Fourier uncertainty relation between spectral and temporal resolution ΔωΔt > 1 is always satisfied once Δω and Δt are properly defined.

The combined spectral and the temporal resolution of these IR and Raman techniques stem from two interactions with a single device: the probe pulse. Nonlinear multidimensional spectroscopy signals depend on several time intervals and there is no conceptual problem in having simultaneous high temporal and spectral resolutions in different independent dimensions69,70. This is not the case when both dimensions are associated with the same probe pulse. The issue was addressed for Raman detection in 71 using a semiclassical treatment of bath coordinates.

Below we present a more general analysis and elaborate on this point for the three protocols and identify the factors that determine the genuine resolution. In the first protocol, the signals (9) – (10) and (17) – (18) are given by a sum over paths spanning both branches of the loop. Naively one can argue that a short pulse must interact impulsively with the system at a precisely defined time. This is not necessarily the case for the following reason: a pulse is a superposition of modes with well defined phases. The broader the bandwidth, the shorter can the pulse be. Eqs. (9)(10) and (17)(18) show that the relevant range of frequencies that actually contribute to a given signal is spanned by the variable Δ. Thus, only some of the probe modes contribute to a given signal, and the full bandwidth of the pulse may become immaterial in some cases. A superposition of the relevant modes has a narrow bandwidth and is necessarily less impulsive than the original pulse, thus reducing the temporal resolution. The number of contributing modes is governed by the width of the relevant spectral features of the system and can be easily understood by the selection of the relevant pathways in the joint field plus matter space. Therefore, the resolution is controlled by the pulse, the measuring device as well as the system in Eqs. (9)(10) and (17)(18). The relevant range of the τ3 integration is controlled by the effective bandwidth of Δ, Δ = 0 implies a CW probe. In both diagrams (i) and (ii) the probe is frequency -dispersed in the detection. If only a single mode is selected for detection one can ask why does the probe duration matter at all? This is apparent from the diagrams which show that the signal involves two interactions with the probe. Frequency -dispersed detection only selects the frequency of the last interaction E2(ω) whereas the other interaction ℰ2(ω + Δ) can still involve many modes, making the signal depend on the probe bandwidth. The time resolution is diminished only if the second interaction also selects a single mode so that Δ = 0.

Turning now to the second protocol, we first note that the time gated measurements (S6) – (S7) and (S8) – (S9) are given by E2(t)E2(τ) which are peaked around t = τ = T. This means that in a time-gated measurement the signal represents a snapshot of the dynamics taken at fixed time t coming from the flat frequency distribution of the probe pulse. Thus, in the joint field plus matter space, a time-domain measurement selects quantum pathways corresponding to the fixed time measurement that is infinitely broad in frequency. However, the frequency - dispersed signals (23) – (24) and (30) – (31) depend on the product E2(ω)E2(ω-ωaa+iγaa). This creates an uncertainty in the interaction time with the probe which is governed by the vibrational dynamics time scale (spectral width of ωa′a) and bath dephasing γa′a. Therefore, the quantum pathways selected by the dynamics of the system yield the effective bandwidth of the probe pulse that interacts with the system. This introduces uncertainty to the interaction time τ3 in Figs. 1b and 2b stemming the finite bath dynamics time scale. The corresponding measurement cannot be viewed as a snapshot of the system, but is determined by the vibrational dynamics that is represented by the coherence between a and a′. The bandwidth of the pump pulse which prepares the system in the density matrix ρaa′ is crucial. The energy spread of ωaa′ is controlled by the pump bandwidth and is also a measure of the inverse time scale of the matter dynamics initiated by the pump. If a single state is selected (ρaa) then there is no dynamics and the same signal can be generated by a CW pump tuned generally to level a. The pump duration then becomes immaterial. A broad distribution of vibrational states will result in fast dynamics that is affected by the pump duration. The broadband technique amounts to multiple two-mode experiments in parallel, which is experimentally convenient since it does not require to scan the frequency, but reveals no additional information beyond the two mode experiment. With initiation, which prepares a wave packet with different a, a′ pairs, the technique may be viewed as many four-wave-mixing (FWM) experiments done in parallel. This is essentially a broad band FWM which only has three modes. For comparison CARS is a four-mode process.

SRS that combines a long picosecond pump with a femtosecond probe has low temporal resolution if treated as a 4-field interaction starting with state a = a′ similar to Eqs. (26)(27) and replacing ω → ω − ω3. In this case, in order to have a highly resolved frequency gated signal, the energy conservation law which follows from the time translational invariance enforces ωω3. In the more general case of a broader pump pulse, we have two interactions with ℰ2(ω3) and E2(ω3). The symmetry breaking then involves four modes of the field and bandwidth of the non stationary preparation state: Δ-ω3+ω3=ωaa+iγaa.

Finally we turn to the third protocol. It is clear that in a frequency gated measurement the probe pulse bandwidth must be broader than the inverse timescale of the vibrational dynamics. The latter is given by the spread of ωa′a and the dephasing rate γa′a. Even if the probe pulse is impulsive and delayed by T, the fact that it is an infrared pulse requires to explicitly take into account the pulse shape of ℰ2(τ) and the pulse may not be simply replaced by δ function, since it may not be shorter than the infrared period. The optical pulses used in the Raman process in contrast can be shorter than the vibrational period and can be truly impulsive. An infrared pulse can be at most “semi impulsive” (i.e. short compared to vibrational relaxation process but not compared to high frequency vibrations >300 cm−1). In the case of the Raman signal this large bandwidth can be easily realized for visible frequencies, and the δ -function approximation is well justified. However, this is not as obvious in the case of an IR probe since the bandwidth of the IR pulses are naturally smaller than in the visible range. Therefore, one must keep the probe pulse envelope and the δ-function approximation is not justified.

To better illustrate the resolution, we examine the Δ - dispersed time-domain signal (S1) dressed by the probe pulse

S¯(t,T;Δ)=-dτE2(τ-T)S(t,T;τ)ei(ω+Δ)τ. (41)

and its variation with Δ. For simplicity in the following we omit the subscript for τ3. Note that in contrast with the τ - dispersed signal (t, T; τ), (t, T; Δ) depends on the probe pulse envelope. As discussed above Δ may be broadened due to finite timescale of the bath dynamics. The relevant frequency domain (8) signal can be calculated using Eq. (41)

SIR(ω,T)=I-dΔ2π-dτ0dteiω(t-T)-iΔτ×E2(ω)S¯(t+τ,T;Δ). (42)

We first consider a simple example for the bath and calculate the effective bandwidth Δ within the semiclassical approximation. We assume linear time variation of the matter transition frequency (linear “matter” chirp): ωac(t)=ωac(0)+αt, where α is a chirp rate. Taking into account Eq. (38) and assuming a harmonic potential with single states a, c and d such that ωac = ωad, setting μad = μac we obtain for the Eq. (41)

S¯(t,T;Δ)=θ(t)0tE2(τ-T)eiΔτ-γa(t+τ)×μag2μac2E12[eiωac(0)(t-τ)+i2α(t2-τ2)+c.c.]. (43)

Assuming a gaussian probe pulse centered around τ = T

E2(τ-T)=E2e-(τ-T)22σpr2-iω0τ, (44)

where ω0 is the central frequency and σpr is the duration of the pulse we obtain

S¯(t,T;Δ)~e-(Δ-Δ0)22σeff2 (45)

where Δ0=ω0-ωac(0)+α(T-σpr2γa) and σeff2=σpr-2+α2σpr2. Note, that effective range for Δ given by σeff contains two contributions. One is the inverse duration of the pulse, and the second is governed by α - a characteristic timescale of the matter dynamics. This effect is similar to the broadening of a chirped pulse compared to the transform-limited pulse with the chirp added by the matter, instead.

We would like to capture the matter dynamics at a given time scale α−1. For a long pulse the dominant contribution to σeff comes from the matter which ensures high frequency resolution. In the limit of resonant CW excitation ( ω0=ωac(0)) Eqs. (45) gives δ (Δ). The latter implies that the original Δ - dispersed signal (38) has no time resolution with respect to τ. This result is independent of time delay of the probe pulse T. In the opposite limit when the pulse duration is small the leading contribution comes from the pulse and σeffσpr-1. Therefore, high temporal resolution is accompanied by poor spectral resolution and vice versa. In both limits, the time and frequency resolution are not independent or solely controlled by external manipulation of pulse parameters. Rather they are governed by a combination of pulse and matter parameters. This simple example provides some basic intuition. However in this linear matter chirp model the transition frequency is changing in an unbounded fashion. We next consider a more realistic model where the transition frequency switches between two values during a finite time interval. For instance

ωac(t)=ωac(0)+12α[F(t0σm)-F(t0-tσm)], (46)

where F(t)=2π0tdxe-x2 is the error function. The transition matter frequency switches from its initial ωac(0) to its final value ωac(0)+α during time interval σm in the vicinity of t0. Fig. 3a depicts Eq. (46) with ωac(0)=200cm-1, α = 200 cm−1, σm=20 fs, t0=500 fs. Fig. 3c shows the Fourier transform of the Δ - dispersed signal

Fig. 3.

Fig. 3

(Color online) Frequency profile [Eq. (46)] - (a) and corresponding 2D the Fourier transform of the Δ - dispersed signal (ω, T = 500 fs, τ) in Eq. (47) - (c) for ωac(0)=2000cm-1, α = 200 cm−1, σm = 20 fs. (b) and (d) - same as (a) and (c), respectively with σm = 200 fs.

S(ω,T;τ)=-dΔ2πS(ω,T;Δ)e-i(ω+Δ)τ (47)

vs ω and τ. For T = 500 fs one can see how the pattern evolves with dominating emission peak at initial frequency ωac(0)=2000cm-1 for short times τ < 500 fs turning into the final frequency ωac(0)+α=2200cm-1 for longer times τ > 500 fs. The oscillatory region of the plot for times shorter than γa-1=1ps shows the frequency beating and matter chirp. At longer times the signal decays exponentially ~ e−γaτ. For slower dynamics, σm = 200 fs (see Fig. 3b) the Δ - dispersed signal (47) plotted in Fig. 3d is similar to Fig. 3c but is stretched according to the longer time scale σm.

In order to determine the time and frequency resolution for the system dynamics given by Eq. (46) with σm = 20 fs and σm = 200 fs, we plotted the Δ - dispersed time-domain signal (41) for different values of the probe pulse duration σpr in Fig. 4a–d. For long probe σpr = 400 fs Fig. 4a shows that the slow matter dynamics results in a single peak at the final frequency ω+Δ=ωac(0)+α, whereas fast dynamics gives two peaks which correspond to the initial and final frequencies. Further increase of the pulse duration [not shown] does not change the fast dynamics case while for slow dynamics the two peaks become narrower. It means that the high frequency resolution is accompanied by poor time resolution in this case. For a shorter pulse σpr = 200 fs both fast and slow dynamics give a single emission peak centered at final frequency ωac(0)+α. However fast dynamics yields a larger bandwidth due to combined pulse and matter bandwidths (see Fig. 4b). Further decrease of the pulse duration for σpr = 50 (Fig. 4c) and 20 fs (Fig. 4d) shows that the fast dynamics converges and becomes indistinguishable from the slow dynamics at σprσm = 20 fs (Fig. 4d). In this case the spectrum does not carry any matter information and looses its frequency resolution. It simply gives the Fourier transform of the probe pulse. Therefore, for long pulse, the spectrum has perfect frequency and poor time resolution. This corresponds to the CW experiment, when the initiation pulse prepares the system in equilibrium population state described by ρaa. In this case time translation invariance via (13) yields ω1=ω1 and consequently ω = ω′. In the opposite limit of the short pulse, the perfect time resolution is accompanied by poor frequency resolution. The resulting spectrum will not contain any relevant matter information and will be given by a Fourier transform of the probe pulse. In both limits the time and the frequency resolution are not independent and are governed by a combination of matter and field parameters.

Fig. 4.

Fig. 4

(Color online) The Δ - dispersed signal Eq. (41) vs Δ for fast (σm = 20 fs, blue) and slow ( σm = 200 fs, red) switchover of the vibrational frequency as depicted in Fig. 3 (a) and (b), respectively. We assume resonant excitation t. Various panels represent different values of the pulse duration σpr = 400 fs - (a), 200 fs - (b), 50 fs - (c), and 20 fs - (d).

7 Discussion

We have employed a superoperator diagrammatic techniques to derive similar expressions for stimulated signals detected by frequency - dispersed transmission of a broadband IR probe and stimulated Raman signals following the broadband visible pump pulse. The resolution is determined by both field and matter degrees of freedom and cannot be solely controlled by the experimental apparatus. The time and frequency resolution was analyzed using three representations and the corresponding computational protocols for the signal. Loop diagrams provide a convenient compact tool for computing and interpreting these signals in terms of the evolving vibronic wavepacket. The vibrational resonances are generated during a single time interval in this diagram where the wave function propagates backward from the observation time corresponding to the ket interaction with the probe field and further to the previous bra interaction with the same probe which is close to the delay time T relative to the preparation field. Note that t − T is a time interval between two successive interactions along the loop but not in real time. A completely time ordered description based on ladder diagrams will separate the loop into several terms52.

The SRS signals may be obtained from the FDIR expressions by the substitution Vn → αn and E2(t)E2(t)E3(t). Note that due to the additional narrow band field 3 in frequency domain SRS the infrared frequency ω must be replaced by the Raman shift ω − ω3. Both SRS and FDIR are given by two diagrams which represent different physical processes. Diagrams (i) and (ii) in Fig. 1b correspond to emission and absorption, respectively of the IR probe pulse, whereas in the case of Fig. 2b each diagram contains both emission and absorption (Stokes and anti-Stokes process). In FDIR we can make the rotating wave approximation (RWA) for the radiation matter coupling and obtain Eq. (2) where V, V are non Hermitian operators. The RWA does not apply for off resonant Raman where we have for the Raman part of Eq. (16)

H(t)=α[E2(t)E3(t)+E2(t)E3(t)], (48)

where α is a real (Hermitian) operator α = α. Both E2(t)E3(t) and E2(t)E3(t) can excite or de-excite the vibrations (Stokes and anti-Stokes process) as permitted by their bandwidths whereas in the FDIR case (Eq. (2)) ℰ2 excites and E2 de-excite the vibrations. This is an important distinction, especially in the case of CARS signals (diagram ii in Fig. 2b), which involve four field modes with different wave vectors where the spatial phase matching becomes crucial72.

Supplementary Material

ESI

Acknowledgments

We gratefully acknowledge the support of the National Institute of Health Grant No. GM-59230, National Science Foundation through Grant No. CHE-1058791 and computations are supported by CHE-0840513, the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science and US Department of Energy. B. P. F. gratefully acknowledges support from the Alexandervon-Humboldt Foundation through the Feodor-Lynen program.

References

  • 1.Kukura P, McCamant DW, Yoon S, Wandschneider DB, Mathies RA. Science. 2005;310:1006–1009. doi: 10.1126/science.1118379. [DOI] [PubMed] [Google Scholar]
  • 2.Schreier WJ, Schrader TE, Koller FO, Gilch P, Crespo-Hernández CE, Swaminathan VN, Carell T, Zinth W, Kohler B. Science (New York, NY) 2007;315:625–9. doi: 10.1126/science.1135428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Adamczyk K, Prémont-Schwarz M, Pines D, Pines E, Nibbering ETJ. Science (New York, NY) 2009;326:1690–4. doi: 10.1126/science.1180060. [DOI] [PubMed] [Google Scholar]
  • 4.Kuramochi H, Takeuchi S, Tahara T. The Journal of Physical Chemistry Letters. 2012;3:2025–2029. [Google Scholar]
  • 5.Kukura P, McCamant DW, Mathies RA. Annual Review of Physical Chemistry. 2007;58:461–488. doi: 10.1146/annurev.physchem.58.032806.104456. [DOI] [PubMed] [Google Scholar]
  • 6.Haiser K, Fingerhut BP, Heil K, Glas A, Herzog TT, Pilles BM, Schreier WJ, Zinth W, de Vivie-Riedle R, Carell T. Angewandte Chemie (International ed in English) 2012;51:408–411. doi: 10.1002/anie.201106231. [DOI] [PubMed] [Google Scholar]
  • 7.Fingerhut BP, Herzog TT, Ryseck G, Haiser K, Graupner FF, Heil K, Gilch P, Schreier WJ, Carell T, de Vivie-Riedle R, Zinth W. New Journal of Physics. 2012;14:065006. [Google Scholar]
  • 8.Fayer MD. Ultrafast Infrared Vibrational Spectroscopy. CRC Press, Tylor and Francis Group; USA: 2013. [Google Scholar]
  • 9.Anfinrud PA, Han C, Hochstrasser RM. Proceedings of the National Academy of Sciences. 1989;86:8387–8391. doi: 10.1073/pnas.86.21.8387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Sawyer KR, Steele RP, Glascoe EA, Cahoon JF, Schlegel JP, Head-Gordon M, Harris CB. The Journal of Physical Chemistry A. 2008;112:8505–8514. doi: 10.1021/jp802705w. [DOI] [PubMed] [Google Scholar]
  • 11.Keane PM, Wojdyla M, Doorley GW, Watson GW, Clark IP, Greetham GM, Parker AW, Towrie M, Kelly JM, Quinn SJ. Journal of the American Chemical Society. 2011;133:4212–4215. doi: 10.1021/ja1106089. [DOI] [PubMed] [Google Scholar]
  • 12.Lynch MS, Slenkamp KM, Cheng M, Khalil M. The Journal of Physical Chemistry A. 2012;116:7023–7032. doi: 10.1021/jp303701b. [DOI] [PubMed] [Google Scholar]
  • 13.Bingaman JL, Kohnhorst CL, Van Meter GA, McElroy BA, Rakowski EA, Caplins BW, Gutowski TA, Stromberg CJ, Webster CE, Heilweil EJ. The Journal of Physical Chemistry A. 2012;116:7261–7271. doi: 10.1021/jp2121774. [DOI] [PubMed] [Google Scholar]
  • 14.Schrader T, Sieg a, Koller F, Schreier W, An Q, Zinth W, Gilch P. Chemical Physics Letters. 2004;392:358–364. [Google Scholar]
  • 15.Mohammed OF, Pines D, Dreyer J, Pines E, Nibbering ETJ. Science (New York, NY) 2005;310:83–86. doi: 10.1126/science.1117756. [DOI] [PubMed] [Google Scholar]
  • 16.Lim M, Jackson T, Anfinrud P. Science. 1995;269:962–966. doi: 10.1126/science.7638619. [DOI] [PubMed] [Google Scholar]
  • 17.Schreier WJ, Kubon J, Regner N, Haiser K, Schrader TE, Zinth W, Clivio P, Gilch P. Journal of the American Chemical Society. 2009;131:5038–9. doi: 10.1021/ja900436t. [DOI] [PubMed] [Google Scholar]
  • 18.Heyne K, Mohammed O, Usman A, Dreyer J, Nibbering E, Cusanovich M. J Am Chem Soc. 2005;127:18100–18106. doi: 10.1021/ja051210k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Mohammed OF, Adamczyk K, Banerji N, Dreyer J, Lang B, Nibbering ETJ, Vauthey E. Angewandte Chemie (International ed in English) 2008;47:9044–8. doi: 10.1002/anie.200803164. [DOI] [PubMed] [Google Scholar]
  • 20.Mohammed OF, Ahmed SA, Vauthey E, Nibbering ETJ. J Phys Chem A. 2009;113:5061–5065. doi: 10.1021/jp811248u. [DOI] [PubMed] [Google Scholar]
  • 21.Nibbering ETJ, Fidder H, Pines E. Annual Review of Physical Chemistry. 2005;56:337–67. doi: 10.1146/annurev.physchem.56.092503.141314. [DOI] [PubMed] [Google Scholar]
  • 22.Szyc L, Yang M, Nibbering ETJ, Elsaesser T. Angewandte Chemie (International ed in English) 2010;49:3598–610. doi: 10.1002/anie.200905693. [DOI] [PubMed] [Google Scholar]
  • 23.Mohammed OF, Luber S, Batista VS, Nibbering ETJ. The journal of physical chemistry A. 2011;115:7550–8. doi: 10.1021/jp202277t. [DOI] [PubMed] [Google Scholar]
  • 24.Prémont-Schwarz M, Xiao D, Batista VS, Nibbering ETJ. The Journal of Physical Chemistry A. 2011;115:10511–10516. doi: 10.1021/jp207642k. [DOI] [PubMed] [Google Scholar]
  • 25.Xiao D, Prémont-Schwarz M, Nibbering ETJ, Batista VS. The Journal of Physical Chemistry A. 2012;116:2775–2790. doi: 10.1021/jp208426v. [DOI] [PubMed] [Google Scholar]
  • 26.Kozich V, Werncke W, Dreyer J, Brzezinka K-W, Rini M, Kummrow a, Elsaesser T. The Journal of Chemical Physics. 2002;117:719. [Google Scholar]
  • 27.Hamm P, Lim M, DeGrado WF, Hochstrasser RM. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:2036–41. doi: 10.1073/pnas.96.5.2036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Zanni MT, Gnanakaran S, Stenger J, Hochstrasser RM. Proceedings of the National Academy of Sciences of the United States of America. 2001:6520–6535. doi: 10.1073/pnas.201412998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Zhuang W, Hayashi T, Mukamel S. Angewandte Chemie International Edition. 2009;48:3750–3781. doi: 10.1002/anie.200802644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Mizutani Y, Kitagawa T. Science. 1997;278:443–446. doi: 10.1126/science.278.5337.443. [DOI] [PubMed] [Google Scholar]
  • 31.McCamant DW, Kukura P, Mathies RA. The Journal of Physical Chemistry A. 2003;107:8208–8214. doi: 10.1021/jp030147n. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Lee SY, Zhang D, McCamant DW, Kukura P, Mathies RA. The Journal of Chemical Physics. 2004;121:3632–3642. doi: 10.1063/1.1777214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Laimgruber S, Schreier WJ, Schrader T, Koller F, Zinth W, Gilch P. Angewandte Chemie International Edition. 2005;44:7901–7904. doi: 10.1002/anie.200501642. [DOI] [PubMed] [Google Scholar]
  • 34.Umapathy S, Lakshmanna A, Mallick B. Journal of Raman Spectroscopy. 2009;40:235–237. [Google Scholar]
  • 35.Mehlenbacher RD, Lyons B, Wilson KC, Du Y, McCamant DW. The Journal of chemical physics. 2009;131:244512. doi: 10.1063/1.3276684. [DOI] [PubMed] [Google Scholar]
  • 36.Takeuchi S, Ruhman S, Tsuneda T, Chiba M, Taketsugu T, Tahara T. Science. 2008;322:1073–1077. doi: 10.1126/science.1160902. [DOI] [PubMed] [Google Scholar]
  • 37.Cheng J-x, Volkmer A, Book LD, Xie XS. The Journal of Physical Chemistry B. 2002;106:8493–8498. [Google Scholar]
  • 38.Mller M, Schins JM. The Journal of Physical Chemistry B. 2002;106:3715–3723. [Google Scholar]
  • 39.Kee TW, Cicerone MT. Opt Lett. 2004;29:2701–2703. doi: 10.1364/ol.29.002701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Volkmer A. Journal of Physics D: Applied Physics. 2005;38:R59. [Google Scholar]
  • 41.von Vacano B, Meyer L, Motzkus M. Journal of Raman Spectroscopy. 2007;38:916–926. [Google Scholar]
  • 42.An Q, Zinth W, Gilch P. Optics Communications. 2002;202:209–216. [Google Scholar]
  • 43.Roy S, Wrzesinski PJ, Pestov D, Dantus M, Gord JR. Journal of Raman Spectroscopy. 2010;41:1194–1199. [Google Scholar]
  • 44.Pestov D, Murawski RK, Ariunbold GO, Wang X, Zhi M, Sokolov AV, Sautenkov VA, Rostovtsev YV, Dogariu A, Huang Y, Scully MO. Science. 2007;316:265–268. doi: 10.1126/science.1139055. [DOI] [PubMed] [Google Scholar]
  • 45.Zumbusch A, Holtom GR, Xie XS. Phys Rev Lett. 1999;82:4142–4145. [Google Scholar]
  • 46.Kosloff R. Annual Review of Physical Chemistry. 1994;45:145–178. [Google Scholar]
  • 47.Chen X, Batista VS. The Journal of Chemical Physics. 2006;125:124313. doi: 10.1063/1.2356477. [DOI] [PubMed] [Google Scholar]
  • 48.Stolz H. Time-Resolved Light Scattering from Excitons. 1994 [Google Scholar]
  • 49.Dorfman KE, Mukamel S. Phys Rev A. 2012;86:013810. [Google Scholar]
  • 50.Dorfman KE, Mukamel S. Phys Rev A. 2012;86:023805. [Google Scholar]
  • 51.Polli D, Brida D, Mukamel S, Lanzani G, Cerullo G. Phys Rev A. 2010;82:053809. [Google Scholar]
  • 52.Rahav S, Mukamel S. Adv At Mol, Opt Phys. 2010;59:223. [Google Scholar]
  • 53.Leforestier C, Bisslinger RH, Cerjan C, Feit MD, Friesner R, Guldberg A, Hammerich A, Jolicard G, Karrlein W, Meyer HD, Lipkin N, Roncero O, Kosloff R. J Comput Phys. 1991;94:59–80. [Google Scholar]
  • 54.Tannor DJ. Introduction to Quantum Mechanics: A Time-Dependent Perspective. University Science Books; 2006. [Google Scholar]
  • 55.Wu Y, Batista VS. The Journal of Chemical Physics. 2004;121:1676–1680. doi: 10.1063/1.1766298. [DOI] [PubMed] [Google Scholar]
  • 56.Beck M. Physics Reports. 2000;324:1–105. [Google Scholar]
  • 57.Lasorne B, Bearpark MJ, Robb MA, Worth GA. Chemical Physics Letters. 2006;432:604–609. [Google Scholar]
  • 58.Martinez TJ, Ben-Nun M, Levine RD. The Journal of Physical Chemistry. 1996;100:7884–7895. [Google Scholar]
  • 59.Ben-Nun M, Martnez TJ. The Journal of Chemical Physics. 1998;108:7244. [Google Scholar]
  • 60.Falvo C, Palmieri B, Mukamel S. The Journal of chemical physics. 2009;130:184501. doi: 10.1063/1.3120771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Abramavicius D, Palmieri B, Voronine DV, Sanda F, Mukamel S. Chemical Reviews. 2009;109:2350–2408. doi: 10.1021/cr800268n. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kwac K, Cho M. The Journal of Chemical Physics. 2003;119:2247–2255. [Google Scholar]
  • 63.Hayashi T, la Cour Jansen T, Zhuang W, Mukamel S. The Journal of Physical Chemistry A. 2005;109:64–82. doi: 10.1021/jp046685x. [DOI] [PubMed] [Google Scholar]
  • 64.Barbatti M, Shepard R, Lischka H. In: Computational and methodological elements for nonadiabatic trajectory dynamics simulations of molecules. Domcke W, Yarkony DR, Koppel H, editors. World Scientific; Singapore: 2011. p. 415. [Google Scholar]
  • 65.Tully JC. J Chem Phys. 1990;93:1061. [Google Scholar]
  • 66.Hammes-Schiffer S, Tully JC. J Chem Phys. 1994;101:4657–4667. [Google Scholar]
  • 67.Reiher M, Neugebauer J. The Journal of Chemical Physics. 2003;118:1634–1641. [Google Scholar]
  • 68.Fingerhut BP, Dorfman KE, Mukamel S. The Journal of Physical Chemistry Letters. doi: 10.1021/jz400776r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Mukamel S. Principles of Nonlinear Optics and Spectroscopy. Oxford University Press; Oxford, UK: 1995. [Google Scholar]
  • 70.Mukamel S. Phys Rev A. 2008;77:023801. doi: 10.1103/PhysRevA.77.022110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Mukamel S, Biggs JD. The Journal of Chemical Physics. 2011;134:161101. doi: 10.1063/1.3581889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Yuan L, Lanin A, Jha P, Traverso A, Voronine D, Dorfman K, Fedotov A, Welch G, Sokolov A, Zheltikov A, Scully M. Laser Physics Letters. 2011;8:736–741. [Google Scholar]
  • 73.Rahav S, Mukamel S. Phys Rev A. 2010;81:063810. doi: 10.1103/PhysRevE.81.031116. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

ESI

RESOURCES