Sum frequency generation spectroscopy

Sum frequency generation spectroscopy (SFG) is a technique used to analyze surfaces and interfaces. This nonlinear laser spectroscopy method was developed in 1987 and rapidly applied to deduce the composition, orientation distributions, and some structural information of molecules at gas–solid, gas–liquid and liquid–solid interfaces. In a typical SFG setup, two laser beams mix at a surface and generate an output beam with a frequency equal to the sum of the two input frequencies. SFG has advantages in its ability to be monolayer surface sensitive, ability to be performed in situ (for example aqueous surfaces and in gases), and not causing much damage to the sample surface. SFG gives complementary information Infrared and Raman spectroscopy.[1]

Theory

IR-visible sum frequency generation spectroscopy uses two laser beams that spatially and temporally overlap at a surface of a material or the interface between two media. An output beam is generated at a frequency of the sum of the two input beams. The two input beams have to be able to access the surface, and the output beam needs to be able to leave the surface to be picked up by a detector.[2] There are two main types of sum frequency spectrometer, scanning systems and broadband systems. For the former type of spectrometer, one of the beams is a visible wavelength laser held at a constant frequency and the other is a tunable infrared laser - by tuning the IR laser, the system can scan over resonances and obtain the vibrational spectrum of the interfacial region in a piecewise fashion.[1]. In the case of broadband spectrometers, one laser (typically the visible laser) is kept at a fixed narrow wavelength, and the other laser produces a spectrally broad beam. These laser beams again overlap at the interface being studied, but can cover a wider range of resonances at once than a spectrometer operating in scanning mode and hence spectra can be acquired much faster, allowing the ability to perform time-resolved measurements with interfacial sensitivity [3].

Nonlinear susceptibility

For a given nonlinear optical process, the polarization which generates the output is given by

where is the th order nonlinear suspectibility, for .

It is worth noting that all the even order susceptibilities become zero in centrosymmetric media. A proof of this is as follows.

Let be the inversion operator, defined by for some arbitrary vector . Then applying to the left and right hand side of the polarization equation above gives

Adding together this equation with the original polarization equation then gives

which implies for in centrosymmetric media. Q.E.D.

[Note 1: The final equality can be proven by mathematical induction, by considering two cases in the inductive step; where is odd and is even.]

[Note 2: This proof holds for the case where is even. Setting gives the odd case and the remainder of the proof is the same.]

As a second-order nonlinear process, SFG is dependent on the 2nd order susceptibility , which is a third rank tensor. This limits what samples are accessible for SFG. Centrosymmetric media include the bulk of gases, liquids, and most solids under the assumption of the electric-dipole approximation, which neglects the signal generated by multipoles and magnetic moments.[2] At an interface between two different materials or two centrosymmetric media, the inversion symmetry is broken and an SFG signal can be generated. This suggests that the resulting spectra represent a thin layer of molecules. A signal is found when there is a net polar orientation.[2][4]

SFG intensity

The output beam is collected by a detector and its intensity is calculated using[2][5]

where is the visible frequency, is the IR frequency and is the SFG frequency. The constant of proportionality varies across literature, many of them including the product of the square of the output frequency, and the squared secant of the reflection angle, . Other factors include index of refractions for the three beams.[1]

The second order susceptibility has two contributions

where is the non-resonating contribution and is the resonating contribution. The non-resonating contribution is assumed to be from electronic responses. Although this contribution has often been considered to be constant over the spectrum, because it is generated simultaneously with the resonant response, the two responses must compete for intensity. This competition shapes the nonresonant contribution in the presence of resonant features by resonant attenuation.[6] Because it is not currently known how to adequately correct for nonresonant interferences, it is very important to experimentally isolate the resonant contributions from any nonresonant interference, often done using the technique of nonresonant suppression.[7]

The resonating contribution is from the vibrational modes and shows changes in resonance. It can be expressed as a sum of a series of Lorentz oscillators

where is the strength or amplitude, is the resonant frequency, is the damping or linewidth coefficient (FWHM), and each indexes the normal (resonant vibrational) mode. The amplitude is a product of , the induced dipole moment, and , the polarizability.[2][4] Together, this indicates that the transition must be both IR and Raman active.[1]

The above equations can be combined to form

which is used to model the SFG output over a range of wavenumbers. When the SFG system scans over a vibrational mode of the surface molecule, the output intensity is resonantly enhanced.[1][4] In a graphical analysis of the output intensity versus wavenumber, this is represented by Lorentzian peaks. Depending on the system, inhomogeneous broadening and interference between peaks may occur. The Lorentz profile can be convoluted with a Gaussian intensity distribution to better fit the intensity distribution.[8]

Orientation information

From the second order susceptibility, it is possible to ascertain information about the orientation of molecules at the surface. describes how the molecules at the interface respond to the input beam. A change in the net orientation of the polar molecules results in a change of sign of . As a rank 3 tensor, the individual elements provide information about the orientation. For a surface that has azimuthal symmetry, i.e. assuming rod symmetry, only seven of the twenty seven tensor elements are nonzero (with four being linearly independent), which are

and

The tensor elements can be determined by using two different polarizers, one for the electric field vector perpendicular to the plane of incidence, labeled S, and one for the electric field vector parallel to the plane of incidence, labeled P. Four combinations are sufficient: PPP, SSP, SPS, PSS, with the letters listed in decreasing frequency, so the first is for the sum frequency, the second is for the visible beam, and the last is for the infrared beam. The four combinations give rise to four different intensities given by

and

where index is of the interfacial -plane, and and are the linear and nonlinear Fresnel factors.

By taking the tensor elements and applying the correct transformations, the orientation of the molecules on the surface can be found.[1][4][8]

Experimental setup

Since SFG is a higher order function, one of the main concerns in the experimental setup is being able to generate a signal strong enough to detect, with discernible peaks and narrow bandwidths. Pico-second and femto-second pulse width lasers are used due to being pulsed lasers with high peak fields. Nd:YAG lasers are commonly used. However, the bandwidth is increased with shorter pulses, forming a tradeoff for desired properties.

Another limitation is the tunable range of the IR laser. This has been augmented by optical parametric generation (OPG), optical parametric oscillation (OPO), and optical parametric amplification (OPA) systems.[8]

Signal strength can be improved by using special geometries, such as a total internal reflection setup which uses a prism to change the angles so they are close to the critical angles, allowing the SFG signal to be generated at its critical angle, enhancing the signal.[8]

Common detector setups utilize a monochromator and a photomultiplier for filtering and detecting.[2]

References

  1. 1 2 3 4 5 6 Shen, Y.R.;"Surface properties probed by 2nd harmonic and sum frequency generation". Nature, v 337, 1989, p 519-525.doi:10.1038/337519a0
  2. 1 2 3 4 5 6 Rangwalla, H.; Dhinojwala, A; (2004) "Probing Hidden Polymeric Interfaces Using IR-Visible Sum-Frequency Generation Spectroscopy". The Journal of Adhesion, v80, Issue 1 & 2, p 37 - 59, doi:10.1080/00218460490276768
  3. "Princess Hashtag Presents: Science". Retrieved 2017-10-06.
  4. 1 2 3 4 Schultz, D.S.; (2005), Interrogating the Electrochemical Interface Using Sum Frequency Generation Spectroscopy".
  5. Chen, Z.; Shen, Y.R.; Samorjai, G.A.; (2002) "Studies of polymer surfaces by sum frequency generation vibrational spectroscopy". Annual Review of Physical Chemistry, v 53, 2002, p 437-465.
  6. Curtis, Alexander D.; Burt, Scott R.; Calchera, Angela R.; Patterson, James E. (19 May 2011). "Limitations in the Analysis of Vibrational Sum-Frequency Spectra Arising from the Nonresonant Contribution". The Journal of Physical Chemistry C: 110519094237033. doi:10.1021/jp200915z.
  7. Lagutchev, A.; Hambir, S.A.; Dlott, D.D. (20 September 2007). "Nonresonant Background Suppression in Broadband Vibrational Sum-Frequency Generation Spectroscopy". Journal of Physical Chemistry C. 111 (37): 13645–13647. doi:10.1021/jp075391j.
  8. 1 2 3 4 Richmond, G.L.; (2002) "Molecular Bonding and Interactions at Aqueous Surfaces as Probed by Vibrational Sum Frequency Spectroscopy", Chemical Reviews, v102, n8, August, 2002, p 2693-2724.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.