Coherent Raman scattering processes such as coherent anti-Stokes Raman scattering and stimulated Raman scattering are described in a tutorial way keeping simple physical pictures and simple derivations. The simplicity of the presentation keeps however most of the key features of these coherent and resonant processes and their intimate relation with spontaneous Raman scattering. This tutorial provides a digest of introduction to the fundamental physics at work, and it does not focus on the numerous technological implementations; rather, it provides the concepts and the physical tools to understand the extensive literature in this field. The presentation is made simple enough for under-graduate students, graduate students, and newcomers with various scientific backgrounds.

The stimulated Raman effect1 was discovered accidentally by Eckhardt in 1962 shortly after the first ruby-laser action was demonstrated.2 This is because the weak molecular stimulated Raman cross section requires laser flux density of the order of 108 W/cm2 that can only be achieved through stimulated emission of radiation. In 1965, Maker and Terhune who were studying nonlinear wave mixing reported a four wave mixing process that can be made resonant with a molecular vibration,3 and the coherent anti-Stokes Raman scattering (CARS) effect was just discovered. Since this early time and until the early 1980s, both stimulated Raman scattering (SRS) and CARS light matter interaction processes were extensively discussed and used in the context of nonlinear optical spectroscopy.4–9 The first implementation of CARS in the context of imaging was reported in 1982 by Duncan10 using two synchronized dye lasers pumped by an Ar-ion laser, whereas the real revival using solid state lasers was reported in 1999 by Zumbusch,11 a seminal study that launched a very active field nowadays known as coherent Raman scattering (CRS) imaging among which the first implementation of SRS microscopy imaging has been reported in 2007.12–14 This active field has been recently reviewed15,16 and takes advantage of the last technological developments in the fields of lasers, detectors, and fast imaging modalities to improve the ability of CRS microscopes to image a variety of compounds in the fields of biology, chemistry, and material sciences. The key asset of the CRS technology is being label-free in contrary to fluorescence. Coherent Raman processes address chemical bonds that are inherently present in matter.

The basics of coherent Raman light matter interaction processes can be found in reference text books such as Refs. 15 and 17–21. However, the information is sometimes scattered or embedded in more general considerations that require significant efforts from the reader to be brought together. The scope of this tutorial is to present the fundamental basics of CRS in a concise yet rigorous manner. The content is certainly not new but is aimed to provide to newcomers in the CRS field the necessary background to grasp the physics at work that is common to the numerous technical implementations that have been reported so far.

The light matter interaction in the near infra-red (NIR) is dominated by the absorption of molecular vibrational levels, the latter giving birth to narrow or broad absorption bands for electromagnetic radiation with wavelengths ranging from 3 μm to 1000 μm.22 Absorption is a resonant process that can be viewed in the simplest case as the interaction of a monochromatic electromagnetic wave with a molecule being modelled as a mass-spring system. We will go through this simple picture to present the NIR molecular absorption, the spontaneous Raman process, and later on, the coherent Raman processes such as CARS and SRS.

The harmonic oscillator is an ideal physical object whose temporal oscillation is a sine wave with constant amplitude and with a frequency that is solely dependent on the system parameters. It is found in many fields of physics, and it is a good approximation of physical systems that are close to a stable position. In the mechanical framework, the simplest harmonic oscillator is a mass m attached to a spring with a stiffness k (Fig. 1). Being set vertical, the mass feels the gravity attraction g and its center of mass is described by
(1)
FIG. 1.

A mass-spring system. x0 is the equilibrium position, and x is the relative displacement.

FIG. 1.

A mass-spring system. x0 is the equilibrium position, and x is the relative displacement.

Close modal
It is possible to study the mass displacement x from its equilibrium position x0. Using the fundamental principles of dynamics (energy conservation), it can be shown23 that the mass displacement x follows
(2)
where ω0=k/m is the system resonant frequency.
Consider now that the mass experiences a friction force F = −2γm(dx/dt), where γ is the damping coefficient. This force is zero when the mass is at rest but increases with the mass speed. The equation describing the mass movement is now23 
(3)
We are now interested in the mass movement when it is driven by a periodic excitation force F(t) = F0 cos(ωt), where F0 is the force amplitude and ω is its angular frequency. The mass movement equation becomes
(4)

It is convenient to use the complex notation F(t) = F0eiωt, with this notation a phase lag is noted positive.

We are now looking for a displacement solution of the form x(t) = x(ω)eiωt. Inserting this generic solution in Eq. (4), it becomes
(5)
where x(ω) can be expressed as
(6)
Close to the resonance (ωω0), and for small damping coefficient (γω0), the harmonic oscillator solution (6) can be approximated by a complex Lorentzian function
(7)

The complex plane [Re(x), Im(x)] is a nice way to display this Lorentzian function [Fig. 2(a)] as the displacement x(ω) describes a circle when the angular frequency ω varies.24 

FIG. 2.

System at resonance. (a) x(ω) describes a circle in the complex plane. (b) Amplitude spectrum. (c) Phase spectrum. At resonance, x(ω) experiences a π/2 phase shift as compared to the phase of the field (red). (d) Time domain description of the 3 regimes: below, at, and above resonance.

FIG. 2.

System at resonance. (a) x(ω) describes a circle in the complex plane. (b) Amplitude spectrum. (c) Phase spectrum. At resonance, x(ω) experiences a π/2 phase shift as compared to the phase of the field (red). (d) Time domain description of the 3 regimes: below, at, and above resonance.

Close modal

Denoting ρ and φ as the amplitude and phase of the displacement, x(ω) = ρ(ω)e(ω). Figure 2(b) displays the displacement amplitude squared with the excitation frequency ω. Figure 2(c) shows the phase response. We note that the system response can be split into three regimes depending on the value of ω [Fig. 2(d)]:

  1. ωω0: The displacement amplitude is weak and its phase is close to zero. The oscillation is in phase with the excitation drive.

  2. ω = ω0: The displacement amplitude is strong at its phase which is π/2. The oscillation lags by 90° with respect to the excitation drive (phase quadrature).

  3. ωω0: The displacement amplitude is weak and its phase is close to π. The oscillation is lagging by 180° with respect to the excitation drive (phase opposition).

Using a simple oscillator model, we have described the phenomenon of resonance. We will now move to vibrational molecular resonances that can be probed by an electromagnetic wave.

A molecular structure sustains many vibrational and rotational intra-molecular vibrations that can be described in terms of normal modes. Each of these “roto-vibrational” normal modes is independent of the other normal modes, and it has a center of mass that is preserved and a specific energy that is quantized as described by the quantum mechanical theory. The atomic mass, the number of involved chemical bonds, the atomic species, the molecule geometry and symmetry, and the possible hydrogen bond interactions affect the stiffness of the vibrational forces at work, which at the end dictate the possible vibrational energies. All these vibrational energies belong to the infrared domain and can be probed by absorption spectroscopy. Rotational modes are only observable when molecules are in the gas phase and correspond to energies ranging from 10 to 400 cm−1, which correspond to the far infrared electromagnetic domain (wavelengths between 25 and 10 000 μm). Roto-vibrational modes correspond to energies ranging from 1 to 4000 cm−1 corresponding to the mid-infrared domain (wavelengths between 2.5 and 25 μm). Near infrared radiations ranging from 4000 to 14 000 cm−1 (wavelengths from 0.7 to 2.5 μm) can also excite overtones of these vibrations.

A molecule having a geometry which is not linear and constituted of n atoms will have 3n − 6 normal vibration modes. For instance, water molecules (H2O) have 3 modes of vibrations that are shown in Fig. 3, and other molecules may exhibit other modes of vibration such as scissoring modes, twisting modes, rocking modes, torsion modes, and wagging modes.25 A molecule having a linear geometry will have only 3n − 5 normal vibration modes because any rotation along its molecular axis keeps the molecule unchanged. Therefore diatomic molecules (n = 2) will have only one normal vibration mode.

FIG. 3.

Vibrational modes of the H2O molecule.

FIG. 3.

Vibrational modes of the H2O molecule.

Close modal

For the sake of clarity, we will consider now a single and isolated diatomic molecule. We suppose that this molecule is made with two nuclei, separated by x0, that are considered as two points with masses m1 and m2. We consider now that the elongation vibrational mode can be described by a simple harmonic oscillator with a resonant frequency ΩR. Within this simple picture, the two nuclei are connected with a spring, as displayed in Fig. 4(a).

FIG. 4.

Infrared (IR) absorption. (a) Classical view of a diatomic molecule absorbing an IR radiation. (b) Energy diagram showing the absorption of a photon exciting a molecule from the ground state f to the first excited vibrational state v.

FIG. 4.

Infrared (IR) absorption. (a) Classical view of a diatomic molecule absorbing an IR radiation. (b) Energy diagram showing the absorption of a photon exciting a molecule from the ground state f to the first excited vibrational state v.

Close modal

We consider now a polar diatomic molecule where an asymmetric distribution of positive and negative charges induces a dipole moment p. We suppose that the first atom holds the charge q, whereas the second atom has the charge −q. The dipole moment of this two punctual charge system can then be expressed in vector form p=qd, where d is the displacement vector oriented from the negative to the positive charge.

We consider now the molecule embedded in an electromagnetic field E(t) with angular frequency ω that is linearly polarized along the molecular axis. This electric field generates a force on the molecule Florentz(t) = qE0eiωt, and the elongation response of the molecule x follows a driven harmonic oscillation [Eq. (4)]
(8)
with μ as the reduced mass, μ = m1m2(m1 + m2). The damping term γ expresses the radiation loss of this oscillating dipole.
Let us now consider a macroscopic medium constituted with N of these molecules. Under the electric field influence, the electronic cloud displacement of each molecule induces a polarization
(9)
The linear electronic susceptibility χ(1) is defined as
(10)
Comparing Eq. (10) with (9), and introducing the solution x(ω) of Eq. (8), we get
(11)
which gives the vibrational contribution to susceptibility χ(1) of the molecular assembly.26 For diluted media, the real part of χ(1) is related to the refractive index dispersion [Fig. 5(b)]
(12)
where n0 is the mean refractive index of the medium, whereas the imaginary part of χ(1) is related to the absorption coefficient of the medium [Fig. 5(c)]
(13)
where α is often expressed in cm−1 and is the key parameter involved in the Beer-Lambert law that describes the intensity loss of a light beam that propagates in a medium over the distance L,
(14)
FIG. 5.

Spectral evolution of the linear susceptibility χ(1) across a molecular resonance. (a) Complex plane representation. (b) The real part of χ(1) is related to the medium dispersion. (c) The imaginary part of χ(1) is related to the medium absorption. A: real part maximum, B: resonance, C: real part minimum.

FIG. 5.

Spectral evolution of the linear susceptibility χ(1) across a molecular resonance. (a) Complex plane representation. (b) The real part of χ(1) is related to the medium dispersion. (c) The imaginary part of χ(1) is related to the medium absorption. A: real part maximum, B: resonance, C: real part minimum.

Close modal

IR absorption is a powerful spectroscopic method to identify and quantify the absorption bands corresponding to molecular species in a sample.22 

At thermodynamic equilibrium, the interaction of light with matter is governed by absorption (the molecule retains the light energy for a certain time) and scattering (the molecule instantaneously scatters the incoming light in a different direction). When a molecule scatters light, most of the scattered photons keep their original frequencies, a phenomenon known as elastic scattering or “Rayleigh scattering” (Fig. 6). However, a small fraction of the incident light is scattered in an inelastic way, i.e., scattered with a frequency that is different from the original incoming light, a phenomenon known as the “Raman effect.” Inelastic Raman scattering was first observed in 1928 by Raman and Krishnan27 in India and independently by Landsherg and Mandelstam28 in USSR. By focusing the spectrally filtered sun light, they could observe new frequencies in the scattered light.

FIG. 6.

Spontaneous Raman scattering mechanisms.

FIG. 6.

Spontaneous Raman scattering mechanisms.

Close modal
To understand the Raman effect, let us again consider the diatomic molecule described earlier. The molecule is not necessary polar, but its polarizability (mostly from electronic origin) depends on the intra-molecular distance x. This distance fluctuates at the resonant molecular bond frequency x(t) = xf cos(ΩRt), where xf is the amplitude fluctuation. For tiny displacements, it is possible to perform a Taylor expansion of the polarizability α(t) near its initial value α0
(15)
The exciting field E(t) = E0 cos(ωpt) induces a dipole moment into the molecule,
(16)
Introducing the polarizability expression (15) into (16), we get
(17)
(18)

The first term with frequency ωP describes the Rayleigh scattering (Fig. 6). The second term with frequency ωS = ωP − ΩR describes a red shifted scattering known as the Raman Stokes scattering. In this case, the molecule moves from its ground state f to an excited vibrational state v (phonon creation), whereas the incoming field loses energy. The third term with frequency ωAS = ωP + ΩR describes a blue shifted scattering known as the anti-Stoke Raman scattering where the molecule goes from the excited vibrational level v toward its ground state f giving energy to the incoming photon (and absorbing a phonon). Experimentally, the anti-Stokes scattered intensity is less than the Stokes scattered intensity (Fig. 7). This is because at thermal equilibrium atomic level populations are described by the Boltzmann statistic indicating that the excited level v is less populated than the ground state. Stokes and anti-Stokes scattered intensity become equally intense only for infinite temperature.

FIG. 7.

Scattered Raman light for a molecule having a single vibrational frequency mode ΩR. The anti-Stokes scattered line is less intense than the Stokes one.

FIG. 7.

Scattered Raman light for a molecule having a single vibrational frequency mode ΩR. The anti-Stokes scattered line is less intense than the Stokes one.

Close modal

The selection rule for a vibration to be Raman active is that the vibration must affect the polarizability, αx00. Similar to IR absorption spectroscopy, Raman spectroscopy enables to identify and quantify the chemical composition of a medium by looking at the inelastic scattered Raman spectrum. However, a normal mode of a molecule with inversion symmetry is Raman active and not IR active or vice versa. Vibrations symmetric with respect to the inversion symmetry are Raman active and those anti-symmetric are IR active.

The Raman scattered field can be depolarized as compared to the incident electromagnetic field. This is described by the Raman depolarization ratio ρR (Fig. 8),
(19)
FIG. 8.

The Raman scattered light can be depolarized as compared to the incident pump light Ip.

FIG. 8.

The Raman scattered light can be depolarized as compared to the incident pump light Ip.

Close modal

This coefficient ρR differs for various roto-vibrational modes and varies between 0 and 34.

Contrary to IR absorption, confocal Raman microscopy can perform imaging with a sub-micron resolution; however, the Raman scattering efficiency process is extremely weak. Raman scattering cross sections are of the order of 10−30 cm2 (as compared to a 1 photon absorption fluorescence cross section that reaches 10−16 cm2, Ref. 29).

1. Nonlinear optics

Nonlinear optics encompasses optical processes that result from the nonlinear response of a medium to the incoming electric field.30 These processes appear when the amplitude of the electric field becomes large as compared to the intra-atomic field (typically, ECoulomb=14πε0ea023.1011 V/m, where a0 ≈ 5 × 10−11 m is the Bohr radius). Nonlinear optical phenomena were observed when the laser was invented in 1960. One year after this discovery, the group from Frenkel could observe the first second harmonic generation (SHG).31 In the same year, Kaiser and Garrett generated two photon excited fluorescence (TPEF).32 In 1965, most of the nonlinear processes were already discovered; among them were the coherent Raman processes.

2. Coherent Raman scattering processes

Coherent Raman scattering (CRS) is a resonant and coherent process that allows gaining a factor of 107 in efficiency as compared to spontaneous Raman scattering.30 Let us consider two incoming plane waves, which we denote pump and Stokes, with frequencies ωP and ωS, respectively. These waves interact with a medium having roto-vibrational modes ΩR. The total field can be written as
(20)
where “c.c.” is the complex conjugate.
The interference between these two fields generates a beating (Fig. 9) with the frequency Ω = ωPωS,
(21)
where is the time average over one optical period and K = kPkS. If the frequency difference Ω = ωPωS is set to Ω = ΩR, the roto-vibrational mode enters in resonance with the wave beating.
FIG. 9.

(a) Classical description of coherent Raman scattering. (b) Two incoming fields, pump and Stokes, beat at frequency Ω = ωPωS that can be equal to the molecular vibrational resonance ΩR. Spectral domain representation. (c) Time domain representation.

FIG. 9.

(a) Classical description of coherent Raman scattering. (b) Two incoming fields, pump and Stokes, beat at frequency Ω = ωPωS that can be equal to the molecular vibrational resonance ΩR. Spectral domain representation. (c) Time domain representation.

Close modal

3. Excitation force calculation

Let us consider again our driven harmonic oscillator describing a diatomic molecule
(22)
The energy necessary to create a dipolar moment p(t)=ε0α(t)E(t) is given by30 
(23)
Similar to spontaneous Raman scattering, we assume that the polarizability is related to the intra-molecular distance following Eq. (15). Inserting (21) in (23) gives the excitation force produced by the two incoming fields on the molecule,
(24)

4. Harmonic oscillator solution

To solve Eq. (22), considering the force (24), we look for a solution given as
(25)
Near the resonance [Eq. (7)], the molecular vibration amplitude is given by
(26)

If the beating frequency is such that Ω = ΩR, the molecular vibration amplitude becomes large and the excitation fields will now also induce nonlinear polarizations that will be specific to the molecular resonance.

5. Induced nonlinear polarization

If N denotes the molecular density, the induced polarization in the medium is given by
(27)
which is the sum of a linear polarization PL(z, t) = 0α0E(z, t) and a nonlinear polarization
(28)
Expanding this expression, it appears that the nonlinear polarization radiates at 4 different frequencies: ωAS = 2ωPωS, ωCS = 2ωSωP, ωP, and ωS,
(29)
where the complex amplitudes P(ωAS), P(ωCS), P(ωP), and P(ωS) are
(30)
(31)
(32)
(33)

Figure 10 shows the energy diagrams associated with these 4 different processes (following the polarization order above):

  • Coherent anti-Stokes Raman Scattering (CARS),

  • Coherent Stokes Raman Scattering (CSRS),

  • Stimulated Raman Loss (SRL),

  • Stimulated Raman Gain (SRG).

FIG. 10.

Coherent Raman processes. (a) Coherent anti-Stokes Raman scattering. (b) Coherent Stokes Raman scattering. (c) Stimulated Raman loss. (d) Stimulated Raman gain.

FIG. 10.

Coherent Raman processes. (a) Coherent anti-Stokes Raman scattering. (b) Coherent Stokes Raman scattering. (c) Stimulated Raman loss. (d) Stimulated Raman gain.

Close modal

In the following, we will not consider the CSRS process whose frequency ωCS is in the IR domain where measurements are noisy and hampered by non-ideal detectors. We will concentrate first on the CARS process and later on the SRL and SRG processes.

The CARS process was first observed by Maker and Terhune in 1965,3 while Levenson and Bloembergen presented a detailed investigation in 1972.33 One of the discoveries was to find that the CARS signal at frequency ωAS was always present, even if the frequency difference between the pump and the Stokes beams did not match a molecular vibrational resonance (Ω ≠ ΩR).

To understand this, let us consider the Jablonski diagram in Fig. 11(a). As described earlier, the resonant CARS process involves the beating between the pump and the Stokes beam frequency Ω that is resonant with the molecular vibration at frequency ΩR. This molecular vibration is further probed by the pump pulse to generate the anti-Stokes signal. However, there is another route, using the frequencies ωP and ωS, to generate a scattered light with frequency ωAS [Fig. 11(b)]. This other four wave mixing (FWM) process is known as the non-resonant CARS as the exciting beams do not interact with the molecular vibrational state; rather, it originates from the instantaneous electronic response of the medium.

FIG. 11.

CARS processes. (a) Resonant CARS. (b) Non-resonant CARS. Dashed lines are virtual states, and continuous lines are real molecular states.

FIG. 11.

CARS processes. (a) Resonant CARS. (b) Non-resonant CARS. Dashed lines are virtual states, and continuous lines are real molecular states.

Close modal
Formally, the nonlinear CARS polarization at location r reads
(34)
where χ(3) is the third order nonlinear susceptibility, a third rank tensor that describes the possible interaction between the pump and Stokes exciting fields with the medium.
Avoiding for a short time the pump and Stokes field spectral dependence, the ith Cartesian component (i = x, y, z) of the nonlinear polarization P(3) generated in point r reads34 
(35)
where indices j, k, and l permute over spatial coordinates x, y, and z.
It can be shown that in the case of isotropic media, where one photon transition at ωP and ωS does not occur, the nonlinear polarization can be simply expressed in terms of the tensor element χxxyy(3) and the Raman depolarization ratio ρR,35,36
(36)
from which we note the following:
  1. When the pump and Stokes beams are linearly polarized and with the same polarization state, the induced nonlinear polarization is collinear with the excitation fields.

  2. When the pump and Stokes beams are linearly polarized with orthogonal polarizations (EpEs=0), the induced nonlinear polarization is always collinear with the Stokes beam. Totally polarized Raman bands cannot be observed.

    • This relation gives also an insight into the Raman depolarization ratio for the generation of the anti-Stokes field.

  3. When the Raman line is totally polarized (ρR = 0), the induced nonlinear polarization is collinear with the pump field and largest for parallel polarization of pump and Stokes.

  4. The more the Raman line is depolarized (ρR), the larger the second contribution of Eq. (36) is, i.e., the anti-Stokes field along the ES direction. For the maximum depolarization of 0.75, the second contribution is at least 3 times larger than the first term.

Some more information on polarization resolved CARS can be found in Ref. 37.

For the sake of simplicity, we will consider from now that the Raman line is totally polarized (ρR = 0). Furthermore, we suppose that the pump and Stokes beams propagate along the z direction, with the same and linear polarization state. In these conditions, Eq. (36) reads
(37)
where χ(3)(ωAS)=2χxxyy(3).

We point out that tackling the problem using 1-D plane wave propagation is useful to capture the basic properties of nonlinear optical microscopy. However, a full consideration of focused fields in 3D is necessary to provide a more accurate picture.

We concentrate in this section on the generation and propagation of the anti-Stokes field. Under the slowly varying envelope approximation, the anti-Stokes EAS(z,t)=AASei(kASzωASt)+c.c. field propagation writes38 
(38)
With the expression of the polarization P(ωAS) defined in (37), it becomes
(39)
where Δk=Δkez=(2kPkSkAS)ez.
Suppose that the anti-Stokes signal is generated in a medium with length L. The anti-Stokes field in z = L reads
(40)
with sinc(x) = sin(x)x.
From which, we get the expression of the anti-Stokes intensity
(41)
It is interesting to note that the anti-Stokes intensity in z = L is proportional to sinc2kL/2) [Fig. 12(a)]. The anti-Stokes signal is efficiently generated only if ΔkL2π ≪ 1 because the total anti-Stokes field is the result of the coherent summation of the anti-Stokes fields emitted from each points along z, and this interference is constructive only if all fields are in phase. For this, the anti-Stokes field must be in phase with the induced nonlinear polarization for every point along the z axis. This is what is known as the “phase matching condition” that requires
(42)
the wave vectors being
(43)
FIG. 12.

(a) CARS signal generation efficiency depending on the phase mismatch and the interaction length L. (b) Phase matching condition for collinear pump and Stokes beams. (c) Anti-Stokes generation efficiency with the interaction length L (the phase mismatch being fixed).

FIG. 12.

(a) CARS signal generation efficiency depending on the phase mismatch and the interaction length L. (b) Phase matching condition for collinear pump and Stokes beams. (c) Anti-Stokes generation efficiency with the interaction length L (the phase mismatch being fixed).

Close modal
In the considered collinear situation (eP=eS=eAS=ez), the phase matching condition is rarely fulfilled (Δk ≠ 0) because of the medium dispersion (n  Pn  Sn  AS). It is useful to define the “nonlinear coherence length” Lc as the length over which the maximum of the anti-Stokes signal develops,
(44)

When the interaction length of the exciting fields L is larger than the nonlinear coherence length, the anti-Stokes signal generation becomes less efficient [Fig. 12(c)] because of destructive interferences. It is therefore important to minimize Δk to get the longest possible coherence length.

In the context of microscopy, the CARS signal is efficiently generated for two reasons. First, the phase matching condition is relaxed when the pump and Stokes beams are strongly focused into the nonlinear medium. The broad angular spectrum of the wave vectors k allows many combinations for the pump, Stokes, and anti-Stokes wave vectors to satisfy the relation (42). Second, a tight focusing of the exciting fields leads to a small nonlinear excitation volume, of the order of 10 μm, such that the interaction length L is too short for destructive interference.

We focus in this section on the anti-Stokes intensity spectrum. We have seen previously that the CARS signal has two contributions.
  • A resonant contribution described by the susceptibility χR(3). Comparing (30) with (37), and inserting x(Ω), previously derived in Eq. (26), we get the CARS resonant nonlinear susceptibility

(45)
with a=2Nε06μΩRαx02, a negative number that represents the oscillator strength of the molecular vibration.
  • A non-resonant contribution described by the susceptibility χNR(3). Far from electronic resonances, this non-resonant contribution is real and constant.9,39

The total CARS susceptibility writes
(46)
Introducing this expression in (41), we get
(47)
(48)
(49)
where it appears that the anti-Stokes intensity is the sum of three contributions that we describe now by following their order.
  • A resonant contribution that contains all vibrational mode spectral information. This is the one that is most wanted in spectroscopy applications.

  • A non-resonant contribution that is constant in the spectral domain.

  • A resonant and non-resonant mixing contribution known as the heterodyne contribution, proportional to the real part of χR(3).

Therefore, one can see the anti-Stokes spectral signal as the output of a two-wave interferometer, the interference being the heterodyne term. Figure 13 shows the spectral behavior of the three contributions, together with the total resulting anti-Stokes signal. Because of the heterodyne term, the CARS spectrum is distorted as compared to the Raman spectrum. The anti-Stokes maximum signal is blue shifted from ΩR, and an intensity minimum appears on the red side of the intensity maximum. Furthermore the CARS spectrum is not symmetric on both sides of ΩR.

FIG. 13.

The anti-Stokes signal in the spectral domain. Contributions of the resonant, heterodyne, and non-resonant intensities. We consider here a Raman line with a half width at half maximum γ = 5 cm−1.

FIG. 13.

The anti-Stokes signal in the spectral domain. Contributions of the resonant, heterodyne, and non-resonant intensities. We consider here a Raman line with a half width at half maximum γ = 5 cm−1.

Close modal

We will now focus on the spectral evolution of the nonlinear susceptibility χ(3) in the complex plane. As every resonance, it can be described as a circle, as already presented in the case of the spring resonance. Figure 14(a) shows the resonant contribution χR(3) as a dashed black circle. As χNR(3) is real, the total CARS susceptibility χ(3)=χR(3)+χNR(3) also describes a circle (in blue) but is shifted on the real axis by χNR(3) (in red). This shift has strong consequences on the χ(3) phase and on the CARS intensity40 [Fig. 14(b)].

  • As we have seen previously, the CARS maximum intensity, or equivalently χ(3) (point 2), is shifted with respect to the Raman resonance (point 3).

  • The χ(3) phase is lower than π2 at the Raman resonance (point 3). Its spectrum shows a maximum (point 4) that is red shifted with respect to the Raman resonance.

  • The CARS minimum intensity is in point 5 and red shifted with respect to the Raman peak.

FIG. 14.

CARS spectrum. (a) CARS susceptibility in the complex plane. The resonant contribution χR(3) associated with the Raman resonance describes a circle (dashed black line). The total CARS susceptibility χ(3) (solid blue circle) is the sum of the resonant χR(3) and the non-resonant χNR(3), a real quantity (red line). (b) Spectral CARS intensity (green) and CARS phase (orange).

FIG. 14.

CARS spectrum. (a) CARS susceptibility in the complex plane. The resonant contribution χR(3) associated with the Raman resonance describes a circle (dashed black line). The total CARS susceptibility χ(3) (solid blue circle) is the sum of the resonant χR(3) and the non-resonant χNR(3), a real quantity (red line). (b) Spectral CARS intensity (green) and CARS phase (orange).

Close modal

In practice, the coherent nature of the CARS process makes the resonant and non-resonant contributions difficult to separate. These two contributions manifest themselves as follows:

  • At resonance, the resonant anti-Stokes field is dephased with respect to the anti-Stokes non-resonant field.41,42

  • If the pump and Stokes field polarization states are different, the resulting resonant and non-resonant anti-Stokes fields will not have the same polarization state. This effect is related to the Raman depolarization ratio and the tensor properties of χ(3).35 

  • In the case of the resonant CARS process, the vibrational level is populated in a coherent way. This coherence is conserved during a time T2 (known as the “coherence time”), of the order of picoseconds. The CARS signal generation is efficient only if the second photon with frequency ωP probes the vibration within a time shorter than T2, as demonstrated in Ref. 43. In the case of the non-resonant CARS, T2 is much shorter (few hundreds of femtoseconds).

Even with this non-resonant background, the CARS process is a powerful spectroscopic tool. CARS can bring similar information than spontaneous Raman, but its resonant and coherent character makes it much more efficient and, therefore, compatible with imaging. Contrary to spontaneous Raman scattering, the CARS signal is blue shifted as compared to the pump and Stokes beams and not polluted by possible molecular unwanted single photon fluorescence. Regnier and Taran were the first to use CARS spectroscopy to measure specific molecular concentration in gas in 1973.44 Begley et al. measured one year later specific molecular concentrations in toluene-benzene mixturessolutions8 using CARS.

In this section, we will describe in detail the SRL and SRG processes. Those two processes are intimately linked and are usually described under the same stimulated Raman scattering (SRS) process. They were discovered in 1962 by Woodbury and Ng45 as they were studying Q-switching processes in a Ruby laser containing a nitrobenzene Kerr cell. They could detect a strong IR radiation coming from the Kerr cell which remained mysterious for some time. Few months later, Woodbury and Eckhardt hypothesized that this came from a stimulated Raman process that could be confirmed experimentally.1 A detailed SRS process description can be found in review articles from Hellwarth,4 Bloembergen,6 and Shen and Bloembergen.5 

As in the CARS process, we define the third order nonlinear susceptibilities χR(3)(ωP) and χR(3)(ωS) associated with the SRL and SRG processes, respectively, through their respective induced polarizations
(50)
(51)
Contrary to the CARS process, the SRS polarizations are induced at the same frequencies than the incoming pump and Stokes beams. The pump and Stokes fields generate P(ωP) that, itself, generates a weak “nonlinear” field EP(3)=AP(3)ei(kPzωPt)+c.c.. Having the same frequency as the incoming pump beam, and being coherent, this weak “nonlinear” field interferes with the incoming excitation pump field EP. As a result, the SRS signal can be viewed as the result of an interference between a weak “nonlinear” field, with the strong incoming field. To better understand this phenomenon, let us consider the nonlinear medium as a single dipole located in z = 0. The resulting intensity of this interference can be written as
(52)
(53)
In a similar way, the weak “nonlinear” field ES(3)=AS(3)ei(kSzωSt)+c.c. interferes with the strong incoming Stokes field ES,
(54)
(55)

The second term A(3)2 is negligible in comparison to A2. However, the quantity AA(3) can be significant and depends on the phase difference between the generated weak field and the incoming strong field. Let us have a closer look at this phase difference term.

The generated fields are related to the nonlinear polarizations by the following expressions:46 
(56)
Note that the abovementioned expressions between the generated fields and the induced polarizations are strictly true in the molecule far field, but a more complex approach leads to the same final result. From the induced polarizations (50) and (51), we get
(57)
Comparing (32) and (33) with (50) and (51), and inserting the expression of x(Ω), previously computed in Eq. (26), we get the nonlinear susceptibility involved into the SRL and SRG processes,
(58)
where a is a negative constant that describes the oscillator strength of the molecular vibration (a=2Nε06μΩRαx02). It appears that the SRL susceptibility is intimately linked to the resonant CARS susceptibility [Eq. (45)] as χR(ωP)=12χR(ωAS). Furthermore, the SRL and SRG processes are symmetrical, χR(ωS)=χR(ωP)*.
At resonance, we note that φχR(3)(ωP)=π2 and φχR(3)(ωS)=π2. Inserting expression (57) into Eqs. (53) and (55), we get the SRL and SRG intensities at the Raman resonance,
(59)
where it clearly appears that, at resonance (Ω = ΩR), the pump intensity experiences a depletion, whereas the Stokes field experiences a gain (see Fig. 15); this is the origin of the terminologies “Stimulated Raman Loss” and “Stimulated Raman Gain.”
FIG. 15.

Spectral domain view of coherent Raman scattering. Different from CARS and CSRS processes that generate new frequencies, SRL and SRG processes appear as intensity loss ΔIP on the pump beam and intensity gain ΔIS on the Stokes beam. These intensity variations ΔIP and ΔIS are amplified by the excitation fields in a heterodyne way.

FIG. 15.

Spectral domain view of coherent Raman scattering. Different from CARS and CSRS processes that generate new frequencies, SRL and SRG processes appear as intensity loss ΔIP on the pump beam and intensity gain ΔIS on the Stokes beam. These intensity variations ΔIP and ΔIS are amplified by the excitation fields in a heterodyne way.

Close modal
We focus now on the interaction of the pump and Stokes fields with a nonlinear medium of length L (Fig. 16). Within the slowly varying envelope approximation, the propagation equations for the pump and Stokes fields write
(60)
FIG. 16.

SRS process description when the pump and Stokes beams propagate into a medium with length L.

FIG. 16.

SRS process description when the pump and Stokes beams propagate into a medium with length L.

Close modal
Inserting the nonlinear polarization expressions (50) and (51), it becomes
(61)
(62)
One of the features of the SRS process is that phase matching is always fulfilled as Δk=(kPkP+kSkS)=0. Therefore the nonlinear field generated at distance z is always in phase with the nonlinear fields generated in any other distances z′ in the medium. Let us compute now the total field resulting from the interference between the nonlinear fields and the incoming excitation field. For simplicity, we consider that the incoming Stokes beam (respectively the incoming pump beam) experiences no intensity variation when it generates the weak nonlinear pump field (respectively the weak nonlinear Stokes field) in Eq. (61) [respectively (62)]. This often valid hypothesis is usually refereed as “non-depleted pump” approximation, where the “pump” terminology refers to the excitation field that induces the nonlinear parametric process.20 Within this hypothesis it comes
(63)
(64)
where it is useful to separate χR(3) into its real and imaginary parts. Furthermore, we have seen previously that, following Eq. (58), χR(3)(ωS)=χR(3)(ωP)*. In the following, we set χR(3)=χR(3)(ωP)=χR(3)(ωS)*. With this, we can write
(65)
(66)
where it appears clearly that the pump amplitude loss and the Stokes amplitude gain along propagation depend on the imaginary part of the susceptibility Im(χR(3)). The real part of the susceptibility induces a variation of the refractive index for the pump (respectively, Stokes), depending on the Stokes (respectively, pump) intensity. Using the pump intensity IP(z)=2n  Pε0cAP(z)2 and the Stokes intensity IS(z)=2n  Sε0cAS(z)2, this refractive index variation writes
(67)
(68)

This is known as “crossed phase modulation” (XPM) and can be viewed as a crossed Kerr effect, the refractive index at one frequency is changed by a wave at another frequency. Therefore XPM can induce some spurious effects of focusing or defocusing, by one beam on the other. This is the origin of small non-resonant background signals in SRS described in Ref. 47.

Let us now focus on the pump and Stokes intensity after the interaction with a nonlinear medium of length L,
(69)
(70)
It appears that the pump gain and the Stokes loss have an exponential dependence with the length L of the medium, a property interesting for applications that use long propagation distances such as in fiber optics. In the microscopy context, the length L where the SRS process is efficient, and that corresponds to the volume where pump and Stokes beams are in focus, is sufficiently small for the quantity within the exponential to be much smaller than 1. Therefore it is possible to perform a first order development,
(71)
(72)
where the second terms of expressions (71) and (72) are the pump intensity loss ΔIP and the Stokes intensity gain ΔIS, respectively. We note that these quantities are linked by ΔIP = (ωSωPIS. And the SRS process can be viewed as an energy transfer from the pump beam toward the Stokes beam. Another feature of expressions (71) and (72) is that both loss and gain are proportional to Im(χR(3)) which is the key property that links directly SRS to spontaneous Raman as we shall see now.

In quantum optics, spontaneous emission is described as a stimulated emission from the vacuum field fluctuations. The vacuum fluctuations are present in any allowed electromagnetic modes, in homogeneous media all k vectors are allowed, but in an electromagnetic cavity, the local density of state can be different and the vacuum spectrum follows the electromagnetic modes that can exist within the cavity. As a consequence, the spontaneous emission of a resonant atom located into such a cavity concentrates in the cavity mode only. Conversely if the cavity has no available mode, the spontaneous emission is hampered. The possible enhancement or inhibition of the spontaneous emission rate is known as the Purcell effect;48 it is a direct consequence of the Fermi golden rule that links the spontaneous emission rate and the local density of electromagnetic states.49 

In the spontaneous Raman scattering process, a beam with frequency ωP interacts with a medium having molecular species. Because the vacuum fluctuations are present in the Stokes modes with frequency ωS such that ωPωS = ΩR, the vacuum fluctuation will be amplified through SRG, as described in Eq. (72),
(73)
where the first term can be neglected as compared to the second one that takes advantage of the heterodyne amplification through IP(0). Because they come from fluctuations, the amplified Stokes fields with intensities δIS have a random phase and propagate in every direction (Fig. 17). The SRG process can operate between the pump field and any Stokes field directions because in SRS the phase matching is always fulfilled. As a result, SRG processes between the incoming pump beam and the random vacuum fluctuations generate incoherent Stokes fields, in every direction, satisfying ωPωS = ΩR; it is nothing but the spontaneous Raman scattering process50 
(74)
FIG. 17.

Spontaneous Raman scattering viewed as stimulated Raman processes. The Stokes fields IS(0, φi) are present as fluctuations (thermal, vacuum field, etc.) and have random phases and random incoming directions. Each of these Stokes fields is amplified as experiencing SRG due to the presence of the pump beam. The final process is a spontaneous Raman scattering where the scattered Stokes photons have random phases and random directions.

FIG. 17.

Spontaneous Raman scattering viewed as stimulated Raman processes. The Stokes fields IS(0, φi) are present as fluctuations (thermal, vacuum field, etc.) and have random phases and random incoming directions. Each of these Stokes fields is amplified as experiencing SRG due to the presence of the pump beam. The final process is a spontaneous Raman scattering where the scattered Stokes photons have random phases and random directions.

Close modal

In 1993, Cairo et al. put a Raman medium (C6H6) in a cavity to study the spontaneous Raman process.51 They observed that the fluctuation spectrum of δIS(0) was related to the cavity modes. They demonstrated that, as spontaneous emission, it is possible to enhance or inhibit spontaneous Raman scattering for a specific Raman line, just by spectral tuning of the cavity.

We have seen that the non-resonant background distorts the CARS intensity spectrum. This is why the CARS spectrum [Eq. (74)] is fundamentally different from the Raman spectrum,
(75)

Conversely in the SRG and SRL processes, the intensity pump depletion ΔIP and intensity Stokes gain ΔIS are proportional to Im(χR(3)), as shown by (71) and (72). Therefore measuring ΔIP and ΔIS gives a direct and rigorous access to the Raman spectrum.

The authors acknowledge fruitful discussions with David Gachet and Esben Ravn Andresen. This work has been supported by the Centre National de la Recherche Scientifique (CNRS) and the Aix-Marseille University.

1.
G.
Eckhardt
,
R. W.
Hellwarth
,
F. J.
McClung
,
S. E.
Schwarz
,
D.
Weiner
, and
E. J.
Woodbury
,
Phys. Rev. Lett.
9
,
455
(
1962
).
3.
P. D.
Maker
and
R. W.
Terhune
,
Phys. Rev.
137
,
A801
(
1965
).
4.
5.
Y. R.
Shen
and
N.
Bloembergen
,
Phys. Rev.
137
,
A1787
(
1965
).
6.
N.
Bloembergen
,
Am. J. Phys.
35
,
989
(
1967
).
7.
A.
Laubereau
and
W.
Kaiser
,
Rev. Mod. Phys.
50
,
607
(
1978
).
8.
R. F.
Begley
,
A. B.
Harvey
, and
R. L.
Byer
,
Appl. Phys. Lett.
25
,
387
(
1974
).
9.
H.
Lotem
,
R. T.
Lynch
, Jr.
, and
N.
Bloembergen
,
Phys. Rev. A
14
,
1748
(
1976
).
10.
M. D.
Duncan
,
J.
Reintjes
, and
T. J.
Manuccia
,
Opt. Lett.
7
,
350
(
1982
).
11.
A.
Zumbusch
,
G. R.
Holtom
, and
X. S.
Xie
,
Phys. Rev. Lett.
82
,
4142
(
1999
).
12.
E.
Ploetz
,
S.
Laimgruber
,
S.
Berner
,
W.
Zinth
, and
P.
Gilch
,
Appl. Phys. B
87
,
389
(
2007
).
13.
C. W.
Freudiger
,
W.
Min
,
G. B.
Saar
,
S.
Lu
,
G. R.
Holtom
,
C.
He
,
J. C.
Tsai
,
J. X.
Kang
, and
X. S.
Xie
,
Science
322
,
1857
(
2008
).
14.
P.
Nandakumar
,
A.
Kovalev
, and
A.
Volkmer
,
New J. Phys.
11
,
033026
(
2009
).
15.
J. X.
Cheng
and
X. S.
Xie
,
Coherent Raman Scattering Microscopy
(
CRC Press
,
2013
).
16.
J.-X.
Cheng
and
X. S.
Xie
,
Science
350
,
aaa8870
(
2015
).
17.
Y.-R.
Shen
,
The principles of nonlinear optics
(
Wiley-Interscience
,
New York
,
1984
), p.
575
.
18.
P. N.
Butcher
and
D.
Cotter
,
The Elements of Nonlinear Optics
(
Cambridge University Press
,
1991
), Vol. 9.
19.
N.
Bloembergen
,
Nonlinear Optics
(
World Scientific
,
1996
).
20.
R. W.
Boyd
,
Nonlinear Optics
(
Academic Press
,
2003
).
21.
J.
Zyss
,
Molecular Nonlinear Optics: Materials, Physics, and Devices
(
Academic Press
,
2013
).
22.
B.
Stuart
,
Infrared Spectroscopy
(
Wiley Online Library
,
2005
).
23.
R. P.
Feynman
,
R. B.
Leighton
, and
M.
Sands
,
Feynman Lectures on Physics, Vol. 1: Mainly Mechanics, Radiation and Heat
(
Addison-Wesley
,
1963
).
24.
J. W.
Fleming
and
C. S.
Johnson
, Jr.
,
J. Raman Spectrosc.
8
,
284
(
1979
).
25.
C.
Marcott
,
J. Amer. Chem. Soc.
121
(
15
),
3809
(
1999
).
26.
B.
Saleh
and
M.
Teich
,
Fundamentals of Photonics
(
Wiley
,
1991
).
27.
C. V.
Raman
and
K. S.
Krishnan
,
Nature
121
,
501
(
1928
).
28.
G. S.
Landsherg
and
L. I.
Mandelstam
,
J. Russ. Phys.-Chem. Soc.
60
,
335
(
1928
).
29.
L.
Kastrup
and
S. W.
Hell
,
Angew. Chem., Int. Ed.
43
,
6646
(
2004
).
30.
R. W.
Boyd
,
Nonlinear Optics
, 3rd ed. (
Academic Press, Inc.
,
2008
).
31.
P. A.
Frenkel
,
A. E.
Hill
,
C. W.
Peters
, and
G.
Weinreich
,
Phys. Rev. Lett.
7
,
118
(
1961
).
32.
W.
Kaiser
and
C. G. B.
Garrett
,
Phys. Rev. Lett.
7
,
229
(
1961
).
33.
M. D.
Levenson
,
C.
Flytzanis
, and
N.
Bloembergen
,
Phys. Rev. B
6
,
3962
(
1972
).
35.
F.
Munhoz
,
S.
Brustlein
,
D.
Gachet
,
F.
Billard
,
S.
Brasselet
, and
H.
Rigneault
,
J. Raman Spectrosc.
40
,
775
(
2009
).
36.
D.
Gachet
,
N.
Sandeau
, and
H.
Rigneault
,
J. Eur. Opt. Soc.: Rapid Publ.
1
,
06013
(
2006
).
37.
S.
Brasselet
,
Adv. Opt. Photonics
3
,
205
(
2011
).
38.
P. N.
Butcher
and
D.
Cotter
,
The Elements of Nonlinear Optics
(
Cambridge University Press
,
1990
).
39.
M. D.
Levenson
and
N.
Bloembergen
,
Phys. Rev. A
10
,
4447
(
1974
).
40.
D.
Gachet
,
F.
Billard
,
N.
Sandeau
, and
H.
Rigneault
,
Opt. Express
15
,
10408
(
2007
).
41.
P.
Berto
,
D.
Gachet
,
P.
Bon
,
S.
Monneret
, and
H.
Rigneault
,
Phys. Rev. Lett.
109
,
093902
(
2012
).
42.
P.
Berto
,
A.
Jesacher
,
C.
Roider
,
S.
Monneret
,
H.
Rigneault
, and
M.
Ritsch-Marte
,
Opt. Lett.
38
,
709
(
2013
).
43.
A.
Volkmer
,
L. D.
Book
, and
X. S.
Xie
,
Appl. Phys. Lett.
80
,
1505
(
2002
).
44.
P. R.
Regnier
and
J. P. E.
Taran
,
Appl. Phys. Lett.
23
,
240
(
1973
).
45.
E. J.
Woodbury
and
W. K.
Ng
,
Proc. Inst. Radio Eng.
50
,
2367
(
1962
).
46.
C.
Rulliere
,
Femtosecond Laser Pulses
(
Springer
,
2005
).
47.
P.
Berto
,
E. R.
Andresen
, and
H.
Rigneault
,
Phys. Rev. Lett.
112
,
053905
(
2014
).
48.
E. M.
Purcell
,
Phys. Rev.
69
,
681
(
1946
).
50.
R.
Hellwarth
,
Prog. Quantum Electron.
5
,
1
(
1979
).
51.
F.
Cairo
,
F. D.
Martini
, and
D.
Murra
,
Phys. Rev. Lett.
70
,
1413
(
1993
).