Fact-checked by Grok 2 weeks ago

Sum-frequency generation

Sum-frequency generation (SFG) is a second-order nonlinear optical process in which two input optical fields at frequencies \omega_1 and \omega_2 interact within a nonlinear medium to produce an output field at the sum frequency \omega_3 = \omega_1 + \omega_2. This interaction arises from the quadratic term in the medium's polarization response, governed by the second-order nonlinear susceptibility tensor \chi^{(2)}, which is nonzero only in materials lacking inversion symmetry, such as certain crystals like beta barium borate (BBO) or . For efficient SFG, phase matching must be achieved, ensuring the wave vectors of the input and output fields satisfy \mathbf{k_3} = \mathbf{k_1} + \mathbf{k_2} to prevent destructive over the interaction length. This condition is typically met through in anisotropic crystals or quasi-phase matching using periodic poling, allowing conversion efficiencies up to several percent in optimized setups. The first experimental observations of nonlinear optical effects, including sum-frequency generation, were reported in 1961 by Franken et al. using a on . The process was first theoretically described in 1962 by Armstrong, Bloembergen, Ducuing, and Pershan, building on foundational work in following the invention of the . SFG finds broad applications in frequency conversion to access wavelengths not directly available from lasers, such as generating mid-infrared for gas by mixing near-infrared and visible beams, or radiation for photochemical studies. In bulk media, it enables tunable sources. At interfaces, SFG serves as a surface-selective probe in vibrational , providing insights into molecular orientation, , and dynamics at buried interfaces like liquid-solid or air-water boundaries, with sub-monolayer sensitivity due to the symmetry-breaking at surfaces. These capabilities have made SFG indispensable in fields ranging from and to biological interface analysis.

Fundamentals

Definition and basic principles

Sum-frequency generation (SFG) is a second-order nonlinear optical process wherein two input , possessing frequencies \omega_1 and \omega_2, interact within a nonlinear medium to generate a single output at \omega_3 = \omega_1 + \omega_2. This interaction conserves , as the of the emergent equals the combined of the incident , \hbar \omega_3 = \hbar \omega_1 + \hbar \omega_2. SFG belongs to the class of nonlinear processes, in which the medium acts as a catalyst without undergoing net or permanent change, relying instead on excitations of the material's electrons to facilitate the mixing. The fundamental mechanism of SFG stems from the second-order nonlinear induced in the medium by the oscillating of the input beams. In the electric dipole approximation, this arises from the quadratic response of the material's electrons to the applied fields, proportional to the product of the two input field strengths. However, such a second-order response vanishes in centrosymmetric media due to inversion , which forbids a permanent ; thus, SFG requires non-centrosymmetric materials, such as certain crystals lacking a center of , to exhibit a nonzero second-order nonlinear . Qualitatively, the interaction can be visualized as the two input "colliding" via the medium's nonlinear response, effectively annihilating to produce the higher-energy output , with the medium providing the necessary transfer without real energy exchange. In addition to energy conservation, SFG obeys momentum conservation, requiring the wave vectors of the photons to satisfy \mathbf{k_3} = \mathbf{k_1} + \mathbf{k_2} vectorially, where \mathbf{k_i} = n_i (\omega_i / c) \hat{\mathbf{d_i}} and n_i is the at each frequency. This condition often necessitates careful alignment to achieve efficient generation, though detailed phase-matching strategies are essential for practical implementation. The related process of (SHG), a degenerate case of SFG, was first observed in 1961 by Franken et al., who irradiated a crystal with a beam. Non-degenerate SFG was first experimentally observed in 1962 by Maker et al., establishing these parametric processes as pioneering demonstrations of shortly after the invention of the .

Relation to other nonlinear processes

Sum-frequency generation (SFG) is closely related to (SHG), which represents a specific instance of the SFG process. In SHG, the two input possess identical frequencies, denoted as \omega_1 = \omega_2, leading to an output at \omega_3 = 2\omega_1. This equivalence arises because SHG can be viewed as the degenerate case of SFG, where the nonlinear interaction simplifies due to frequency degeneracy. However, SFG extends this capability by allowing inputs at distinct frequencies \omega_1 \neq \omega_2, enabling the production of tunable output wavelengths across a broader spectral range, which is particularly useful for applications requiring flexibility in frequency selection. Another key related process is difference-frequency generation (DFG), which serves as the inverse of SFG in the context of second-order . While SFG produces an output frequency \omega_3 = \omega_1 + \omega_2 through up-conversion, combining lower-frequency inputs to yield a higher-frequency output, DFG generates \omega_3 = |\omega_1 - \omega_2|, facilitating down-conversion from higher to lower frequencies. This duality allows DFG and SFG to complement each other in frequency manipulation schemes, such as in tunable systems, where one can generate signals for the other. Both rely on the same second-order nonlinear \chi^{(2)} in non-centrosymmetric . SFG differs fundamentally from third-order nonlinear processes, such as third-harmonic generation (THG), in terms of the governing material response. SFG depends on the second-order nonlinear \chi^{(2)}, which is absent in centrosymmetric materials and drives three-wave mixing interactions. In contrast, THG involves \chi^{(3)}, enabling and occurring even in materials lacking \chi^{(2)} symmetry, such as glasses or isotropic media. This distinction underscores SFG's requirement for non-centrosymmetric crystals, while THG offers broader material compatibility but typically lower efficiency for harmonic generation without additional cascading. As a member of the parametric family of nonlinear optical processes, SFG functions as a frequency mixer within a nonlinear medium, converting input frequencies without the need for population inversion, unlike laser gain media. In parametric amplification contexts, SFG can contribute to the generation of idler waves or signal amplification through coupled wave interactions, maintaining energy conservation via Manley-Rowe relations. This parametric nature highlights SFG's role in efficient, coherent frequency translation for photonics applications.

Theoretical framework

Mathematical description

Sum-frequency generation (SFG) arises from the second-order nonlinear polarization induced in a non-centrosymmetric medium by interacting electric fields. The nonlinear polarization \mathbf{P}^{(2)}(t) is expressed as \mathbf{P}^{(2)}(t) = \epsilon_0 \boldsymbol{\chi}^{(2)} : \mathbf{E}(t) \mathbf{E}(t), where \boldsymbol{\chi}^{(2)} is the second-order nonlinear susceptibility tensor and \epsilon_0 is the vacuum permittivity. This quadratic response generates new frequencies when the input field \mathbf{E}(t) comprises multiple frequency components. For input fields at frequencies \omega_1 and \omega_2, the is expanded in the , yielding a term at the sum frequency \omega_3 = \omega_1 + \omega_2:
P_i^{(2)}(\omega_3) = 2 \epsilon_0 \sum_{j,k} \chi_{ijk}^{(2)}(\omega_3; \omega_1, \omega_2) E_j(\omega_1) E_k(\omega_2),
where the indices i, j, k denote Cartesian components, and the factor of 2 accounts for time-dependent permutations in the expansion. This acts as a source for the generated field at \omega_3.
The propagation of the fields is described by the nonlinear wave equation:
\nabla^2 \mathbf{E} - \frac{n^2}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = \frac{1}{\epsilon_0 c^2} \frac{\partial^2 \mathbf{P}^{(2)}}{\partial t^2},
where n is the and c is the in vacuum. For collinear plane waves propagating along z, assuming slowly varying envelopes A_j(z) for each field E_j(z, t) = A_j(z) e^{i(k_j z - \omega_j t)} + \mathrm{c.c.}, the equation yields coupled differential equations. Neglecting pump depletion, the equation for the SFG amplitude is
\frac{d A_3}{dz} = i \frac{\omega_3 \chi^{(2)}}{n_3 c} A_1 A_2 e^{i \Delta k z},
with \Delta k = k_3 - k_1 - k_2 the phase mismatch and n_3 the at \omega_3. Integrating over interaction length L gives the SFG field amplitude proportional to A_1 A_2 L \mathrm{sinc}(\Delta k L / 2).
The generated SFG intensity I_3 follows from I_3 \propto |A_3|^2 and scales as
I_3 \propto I_1 I_2 |\chi^{(2)}|^2 L^2 \mathrm{sinc}^2(\Delta k L / 2),
where I_1 and I_2 are the input intensities; efficiency is maximized when \Delta k = 0. This relation highlights the quadratic dependence on input powers and the role of phase matching in enhancing conversion.
The susceptibility \boldsymbol{\chi}^{(2)} is a third-rank tensor with 27 components, constrained by the medium's point group symmetry to fewer independent elements. In contracted notation, it is often expressed via the second-order nonlinear coefficient tensor d_{il} = \frac{1}{2} \chi_{ijk}^{(2)}, where l runs over 1 to 6 corresponding to strain-like indices (e.g., l=1 for xx, l=4 for yz). The effective nonlinearity d_{\mathrm{eff}} in the coupled equations depends on the polarization directions of the fields and crystal orientation, determining the SFG response for specific input configurations. For many crystals, Kleinman symmetry further simplifies the tensor by assuming invariance under index permutation (\chi_{ijk}^{(2)}(\omega_3; \omega_1, \omega_2) = \chi_{ikj}^{(2)}(\omega_3; \omega_2, \omega_1), etc.), valid in the absence of strong dispersion; this reduces the number of independent components, as seen in classes like 3m (e.g., LiNbO_3) where only a few d_{ij} elements dominate.

Phase-matching requirements

In sum-frequency generation (SFG), efficient conversion requires phase matching, where the wave vectors of the interacting fields satisfy the condition that the generated sum-frequency wave propagates in phase with the input waves over the interaction length. The phase mismatch parameter is defined as \Delta k = k_3 - k_1 - k_2, where k_i = \frac{\omega_i n(\omega_i)}{c} for i = 1, 2, 3, with \omega_3 = \omega_1 + \omega_2, n(\omega_i) the refractive index at frequency \omega_i, and c the speed of light in vacuum. When \Delta k \neq 0, the fields accumulate a phase slip, limiting the effective interaction distance to the coherence length L_c = \frac{\pi}{|\Delta k|}. The conversion efficiency in SFG is governed by the intensity of the generated field, which scales as \eta \propto \left[ \frac{\sin(\Delta k L / 2)}{\Delta k L / 2} \right]^2, where L is the length; this sinc-squared function peaks at \Delta k = 0 and decays rapidly for |\Delta k| L > \pi, reducing efficiency when the phase slip exceeds one over the . To achieve \Delta k \approx 0, birefringent matching exploits the wavelength- and polarization-dependent refractive indices in anisotropic crystals, aligning the effective indices such that n_e(\omega_3) = n_o(\omega_1) = n_o(\omega_2) for Type I configurations (both input waves as rays, output as ) or n_e(\omega_3) = n_o(\omega_1) = n_e(\omega_2) for Type II (one and one input). These configurations allow tuning via crystal angle or to compensate for dispersion-induced mismatches. For materials lacking sufficient or to access non-critical matching, quasi-phase matching (QPM) periodically reverses the sign of the second-order nonlinear \chi^{(2)} along the propagation direction, introducing a vector k_g = \frac{2\pi}{\Lambda} (with poling period \Lambda) that compensates the mismatch via \Delta k = k_3 - k_1 - k_2 = m k_g, typically for (m=1) interactions. This technique, implemented through electric-field poling in ferroelectric crystals like periodically poled (PPLN), enables efficient SFG over broader wavelength ranges and higher nonlinear coefficients without walk-off losses inherent in birefringent methods. Material , arising from the frequency dependence of the , inherently limits the phase-matching in SFG, as small deviations in \omega_1 or \omega_2 alter \Delta k via mismatches. Sellmeier equations empirically model this as n^2(\omega) = 1 + \sum_j \frac{B_j \lambda^2}{\lambda^2 - C_j}, where \lambda = 2\pi c / \omega and coefficients B_j, C_j are material-specific, allowing prediction of phase-matching angles or periods but highlighting trade-offs in broadband operation.

Experimental aspects

Setup and configurations

Sum-frequency generation (SFG) experiments typically employ a basic optical arrangement where two input beams at frequencies \omega_1 and \omega_2 are spatially and temporally overlapped within a nonlinear medium to produce the output at \omega_3 = \omega_1 + \omega_2. In vibrational SFG (VSFG), a common uses a fixed visible beam (e.g., 532 nm from the second harmonic of a Nd:YAG ) and a tunable beam (e.g., 2–8 \mum), focused onto the sample or using lenses to achieve tight overlap and maximize nonlinear interaction efficiency. The nonlinear medium, such as a or , is placed at the , often in a to probe surfaces while minimizing bulk contributions. Beam geometries can be collinear or noncollinear, each offering distinct practical benefits. In collinear setups, the input beams propagate along the same path after combination, simplifying and providing high stability, which is advantageous for phase-resolved measurements; however, timing control is limited without additional elements like birefringent crystals. Noncollinear geometries, where beams approach at different angles (e.g., visible at 53° and at 62° relative to the surface ), allow independent manipulation for precise temporal overlap via delay stages but require careful spatial and may suffer from reduced reproducibility. Detection of the SFG signal focuses on isolating the weak output from the strong inputs. Photomultiplier tubes (PMTs) with gated electronics are commonly used for in scanning setups, while (CCD) cameras enable rapid spectral acquisition in configurations, often achieving spectra in under 10 ms with 1 kHz repetition rates. Spectral and spatial filtering, via interference filters, dielectric mirrors, or monochromators, separates \omega_3 from \omega_1 and \omega_2, with apertures and pinholes rejecting ; is enhanced by chopper modulation of one input beam (e.g., the visible pump) at a few hundred Hz, coupled with lock-in amplification to suppress background and noise. SFG setups predominantly use pulsed lasers due to the quadratic dependence of nonlinear efficiency on peak intensity, though continuous-wave (CW) sources are occasionally employed for stability in specific applications. Femtosecond pulses (e.g., ~100 fs) provide high peak powers (up to several GW) for efficient generation with low average powers, enabling broadband spectroscopy but requiring precise temporal overlap on the order of the pulse duration; picosecond pulses (~10–80 ps) offer better spectral resolution with moderate peak powers. Recent advancements include high-repetition-rate femtosecond systems operating at kHz rates, allowing efficient SFG with reduced pulse energies while maintaining high average powers. CW lasers, by contrast, ensure long-term stability and simpler operation without timing jitter but demand high average powers (mW to W) to achieve comparable efficiency, limiting their use to high-power systems like frequency combs. Alignment and safety protocols emphasize precise beam combination and power management. Dichroic mirrors combine the inputs (e.g., reflecting visible while transmitting IR) to maintain spatial overlap without dispersion, with half-wave plates and polarizers adjusting polarization (often p-polarized for surface studies); delay stages fine-tune temporal synchronization. Typical power levels range from 50–250 \muJ per pulse for pulsed systems (corresponding to mW average power at typical repetition rates of 1–20 Hz and GW peak powers for fs durations), kept below damage thresholds (e.g., <200 \muJ for metals) to ensure safe operation.

Nonlinear media and materials

Sum-frequency generation (SFG) relies on second-order nonlinear optical media characterized by a non-zero second-order susceptibility tensor \chi^{(2)}, which enables the interaction of input fields at frequencies \omega_1 and \omega_2 to produce an output at \omega_3 = \omega_1 + \omega_2. Inorganic crystals such as and are widely used for UV-visible SFG due to their high nonlinear coefficients, broad phase-matching ranges, and high damage thresholds. BBO exhibits an effective nonlinear coefficient d_\mathrm{eff} \approx 2 pm/V and a transparency window from approximately 190 nm to 3.5 \mum, making it suitable for generating deep-UV radiation via SFG of visible and near-IR beams. LBO, with d_\mathrm{eff} \approx 0.85 pm/V and transparency from 160 nm to 2.6 \mum, offers superior thermal stability and is preferred for high-power applications where temperature control is feasible. Ferroelectric materials like periodically poled lithium niobate (PPLN) enable quasi-phase matching (QPM) for SFG, circumventing birefringence limitations in bulk crystals. PPLN achieves high d_\mathrm{eff} values up to 16 pm/V through periodic poling of the \chi^{(2)} tensor, with a transparency range of 350 nm to 5.2 \mum, facilitating efficient SFG in the near-IR to visible regime. Organic crystals, such as those based on urea derivatives (e.g., urea L-malic acid), provide broad transparency extending into the UV and potentially higher \chi^{(2)} due to intramolecular charge transfer, though they often suffer from lower mechanical stability compared to inorganics. Key selection criteria for SFG media include a high \chi^{(2)} (typically >1 pm/V for practical ), high laser-induced damage (e.g., >10 /cm^2 for BBO at 1064 ), and minimal absorption at \omega_3 to avoid thermal lensing. BBO and LBO, grown via high-temperature flux methods, exhibit phase-matchable ranges of 400-2000 , but absorption edges in the UV can limit output power. Temperature sensitivity is pronounced in birefringent crystals like LBO, necessitating thermo-optic tuning with stability better than 0.1°C for optimal phase matching. Emerging nanostructured media, including metamaterials and waveguides, enhance SFG efficiency through local field confinement and engineered \chi^{(2)} . nanocylinders, such as AlGaAs structures, have demonstrated SFG at telecommunication wavelengths with enhanced conversion via Mie resonances, overcoming bulk limitations in centrosymmetric materials. Thin-film integrated based on PPLN waveguides further boost interaction lengths while minimizing walk-off, though fabrication challenges like poling uniformity persist. Growth techniques for bulk crystals, such as Czochralski pulling for LiNbO_3 or flux methods for borates, influence defect density and thus damage thresholds and optical quality.

Applications and techniques

Sum-frequency generation spectroscopy

Vibrational sum-frequency generation (VSFG) employs an beam at \omega_1 and a visible beam at \omega_2 to produce a sum- signal at \omega_3 = \omega_1 + \omega_2, enabling the probing of vibrational resonances specifically at non-centrosymmetric interfaces, such as air-water or solid-liquid boundaries. This technique selects interfaces lacking inversion symmetry, where the second-order nonlinear susceptibility \chi^{(2)} is nonzero, allowing selective detection of molecular vibrations without bulk interference. VSFG has been instrumental in studying molecular ordering at these interfaces since its experimental establishment in the late 1980s. Spectral analysis in VSFG reveals resonance enhancement when \omega_1 aligns with molecular vibrational transitions, amplifying the signal for IR-active modes while the visible beam provides Raman-like . The resulting spectra display peaks corresponding to vibrational modes, with and information extracted via phase-sensitive detection to determine molecular orientation; for instance, the tilt angle of methyl groups on chains can be quantified from the relative intensities and phases of symmetric and asymmetric stretches. This orientation analysis is crucial for understanding interfacial structure, as the shift indicates the directionality of polar groups relative to the surface normal. Broadband VSFG advancements since 2010 utilize mid-IR supercontinua generated from lasers to cover wide ranges (e.g., 950–1750 cm⁻¹), facilitating simultaneous detection of multiple vibrational modes in a single acquisition. These developments enhance throughput and for complex spectra, often combined with schemes—such as SS (both input beams and output S-polarized), SP, PS, and PP (P-polarized variants)—to isolate tensor elements of \chi^{(2)} and probe specific molecular orientations or symmetries. For example, the SSP configuration is particularly sensitive to methyl symmetric stretches at hydrophobic interfaces. As of 2025, sub-1 cm⁻¹ high- VSFG has further improved sensitivity and signal-to-noise ratios for air/ interfaces. The surface specificity of VSFG arises from quantum mechanical selection rules requiring vibrational modes to be both IR- and Raman-active, ensuring signals only from asymmetrically distributed molecules at the interface. This has enabled applications to biomembranes, where VSFG elucidates and , orientation, and interactions with layers, revealing how amyloidogenic peptides disrupt structure. In catalysis, early 1990s studies applied VSFG to monitor CO adsorption on Pt electrodes, providing insights into coverage-dependent bonding geometries and reaction intermediates during electrochemical oxidation. Recent in-operando applications as of 2025 use VSFG alongside to study electrochemical interfaces in .

Microscopy and advanced imaging

Sum-frequency generation (SFG) microscopy extends the interface-specific capabilities of SFG to spatially resolved , enabling the visualization of molecular structures and dynamics at buried or complex interfaces with sub-micron resolution. Early implementations combined SFG with confocal optics in scanning configurations, where a focused beam raster-scans the sample to generate maps of vibrational signals from non-centrosymmetric regions, such as lipid bilayers in cell membranes. Wide-field setups, introduced in the , illuminate larger areas simultaneously for faster acquisition, achieving 3D imaging through z-sectioning and revealing orientational order in self-assembled monolayers or biomimetic materials. Hyperspectral vibrational SFG (VSFG) microscopy represents a post-2020 advancement, integrating spectrometers with laser-scanning or line-scanning geometries to produce wavelength-resolved maps that correlate with . These systems use mid-infrared and visible/near-infrared pulses, dispersed via monochromators and detected with arrays, yielding lateral resolutions of approximately 1 μm via high-numerical-aperture objectives. For instance, multimodal hyperspectral VSFG has imaged fibers and self-assemblies, distinguishing molecular ordering through spectral signatures like I bands. As of 2024, azimuthal-scanning hyperspectral SFG has enabled extraction of heterogeneous 3D structures in molecular films, while tip-enhanced SFG in 2025 achieves nanoscale resolution using plasmonic nanocavities for in-operando control. Electronic SFG (ESFG) probes electronic transitions at optoelectronic interfaces by employing UV-visible input beams resonant with excitonic states, providing insights into and alignment without bulk . In solar cells, ESFG reveals cation orientations and surface defects influencing photovoltaic performance, with time-resolved variants using pulses to track ultrafast dynamics like hot carrier relaxation on timescales. In biological applications, SFG microscopy has imaged protein adsorption on surfaces since the 2010s, capturing conformational changes in fibrinogen or at hydrophobic interfaces with sub-monolayer sensitivity. For , it visualizes ligand distributions on surfaces, such as or silica colloids, to assess stability and reactivity. Multimodal integration with (SHG) enhances contrast, combining electronic and vibrational signals for comprehensive 3D mapping of heterogeneous samples like polymer blends or tissue scaffolds.

References

  1. [1]
    [PDF] Nonlinear Optics
    Feb 1, 2022 · Page 1. Nonlinear Optics. Fourth Edition. Robert W. Boyd. The Institute of Optics ... Sum- and Difference-Frequency Generation ...
  2. [2]
    Sum and Difference Frequency Generation – SFG, DFG
    Sum and difference frequency generation are nonlinear processes generating beams with the sum or difference of the frequencies of the input beams.
  3. [3]
    [PDF] Sum-Frequency Generation at Surfaces - Raschke Nano-Optics Group
    Under sufficiently intense illumination the optical properties of a material system depend nonlinearly on the strength of the electromagnetic field. Related to.
  4. [4]
    Sum-Frequency Generation - an overview | ScienceDirect Topics
    Sum frequency generation (SFG) is a second-order nonlinear optical process that converts two intense input lights into a new light the frequency of which is the ...
  5. [5]
    Historical perspective (Chapter 1) - Fundamentals of Sum ...
    Sum-frequency spectroscopy is among the most powerful and versatile nonlinear optical techniques that have been developed for material studies, and has been ...
  6. [6]
    Difference‐frequency generation and sum ... - AIP Publishing
    We present recent experimental results of difference‐frequency generation (DFG) and sum‐frequency generation (SFG) from zincblende materials.
  7. [7]
    [PDF] Overview of second and third order optical nonlinear processes - arXiv
    These harmonic generations are described in all aspects and how their efficiency will be affected by other nonlinear processes, those cannot be removed by phase.
  8. [8]
  9. [9]
    Nonlinear Optics - ScienceDirect.com
    One of the sections treats the process of sum-frequency generation in the simple limit in which the two input fields are undepleted by the nonlinear interaction ...
  10. [10]
    Interactions between Light Waves in a Nonlinear Dielectric
    Interactions between Light Waves in a Nonlinear Dielectric. J. A. Armstrong, N. Bloembergen, J. Ducuing*, and P. S. Pershan. Division of Engineering and ...
  11. [11]
    Quasi-phase-matched second harmonic generation: tuning and ...
    Abstract: The theory of quasi-phase-matched second-harmonic generation is presented in both the space domain and the wave vector mismatch domain.Missing: LiNbO3 | Show results with:LiNbO3
  12. [12]
    [PDF] Sum Frequency Generation vibrational studies of catalytic surfaces
    The experimental setup of SFG and the UHV-high pressure system are described in chapter 3. An in depth characterization of the laser and optical parametric ...
  13. [13]
    A General Approach To Combine the Advantages of Collinear and ...
    Common heterodyned SFG spectroscopy setups: (A) noncollinear geometry; (B) collinear geometry. ... Sum Frequency Generation Vibrational Spectroscopy: A Tutorial ...
  14. [14]
    [PDF] A Practical Guide to Sum Frequency Generation - OSTI.GOV
    In this tutorial review, we discuss how the choice of upconversion pulse shape in broadband vibrational sum frequency generation (SFG) spectrometer design ...
  15. [15]
    [PDF] Maximizing Sum Frequency Generation Efficiency
    They achieved a 41-dB increase in the signal-to-noise ratio as compared to direct detection. ... The pump is modulated by a chopper, assume the frequency is ω.
  16. [16]
    Three-Wave Mixing between Continuous-Wave and Ultrafast Lasers
    May 7, 2024 · This experiment investigated sum-frequency generation between a CW laser and an ultrafast frequency comb laser. The resulting SFG beam has the ...<|control11|><|separator|>
  17. [17]
    Nonlinear Crystal Materials - RP Photonics
    Nonlinear crystal materials exhibit optical nonlinearity, often of χ(2) type, used for frequency conversion and other purposes.Missing: susceptibility | Show results with:susceptibility
  18. [18]
    Characteristics of an organic nonlinear optical material urea l-malic ...
    Urea l-malic acid is one of such OIC molecular crystals. It is a derivative of urea, which keeps the nonlinear optical properties and broad transparency range ...
  19. [19]
    [PDF] Crystals - EKSMA Optics
    Nonlinear coeff. d36, pm/V (at 1.06 µm). 0.43. 0.40. Effective nonlinear coefficient type 1 type 2 dooe = d36 × sinθ × sin2φ deoe = d36 × sinθ × cos2φ laser ...
  20. [20]
    Nonlinear Frequency Conversion - RP Photonics
    Nonlinear frequency conversion uses optical nonlinearities to change light frequencies, like frequency doubling and optical parametric oscillation.
  21. [21]
    Frequency Tripling via Sum-Frequency Generation at the Nanoscale
    Multiple electromagnetic waves can interact in a material through nonlinear optical processes to generate coherent light of a different frequency. The ...<|control11|><|separator|>
  22. [22]
    Surface vibrational spectroscopy by infrared-visible sum frequency ...
    Feb 15, 1987 · We report the first observation of a vibrational spectrum of a monolayer of molecular adsorbates by infrared-visible sum frequency generation.Missing: seminal paper
  23. [23]
    Computational Analysis of Vibrational Sum Frequency Generation ...
    May 5, 2017 · Annual Review of Physical Chemistry, Volume 68, 2017, Computational Analysis of Vibrational Sum Frequency Generation Spectroscopy.
  24. [24]
    Structure and Dynamics of Interfacial Peptides and Proteins from ...
    Jan 15, 2020 · This review is organized as follows. Section 2 provides a ... Vibrational Sum-Frequency Generation Spectroscopy of Aqueous Protein Films.
  25. [25]
    [PDF] Sum Frequency Generation Spectroscopy of Fluorinated Organic ...
    Apr 24, 2023 · SFG spectra for APFN, F17C9 acid, and F16C9 acid solutions under the ssp (c) and ppp (d) polarization combinations. Reproduced from ref. 80 with ...
  26. [26]
    Tutorials in vibrational sum frequency generation spectroscopy. I ...
    Jan 20, 2022 · ... selection rule that the molecular vibration being studied must be both IR and Raman active for it to produce a nonzero VSFG signal. It is ...
  27. [27]
    Quantitative vibrational sum-frequency generation spectroscopy of ...
    Aug 5, 2025 · The chemisorption and anodic oxidation of CO has been studied by SFG on Pt, [48][49][50] [51] [52] but not on Cu. ... Electrochemical ...
  28. [28]
    Vibrational Sum-Frequency Generation Hyperspectral Microscopy ...
    Apr 20, 2021 · In this review, we discuss the recent ... Vibrational sum frequency generation spectroscopy of fullerene at dielectric interfaces.
  29. [29]
    Rapid vibrational imaging with sum frequency generation microscopy
    We demonstrate rapid vibrational imaging based on sum frequency generation (SFG) microscopy with a collinear excitation geometry.
  30. [30]
    Multimodal Nonlinear Hyperspectral Chemical Imaging Using Line ...
    A multimodal, rapid hyperspectral imaging framework was developed to obtain broadband vibrational sum-frequency generation (VSFG) images.
  31. [31]
  32. [32]
    Time-resolved electronic sum-frequency generation spectroscopy ...
    Oct 27, 2021 · Sum-frequency generation (SFG) is a non-linear optical process that occurs under intense electric fields and relies on the second-order electric ...INTRODUCTION · II. EXPERIMENTAL · III. RESULTS · IV. DISCUSSION AND...
  33. [33]
    Sum frequency generation for the investigation of interfaces
    Sum frequency generation (SFG), for example, involves two incoming electric fields of frequency ω1 and ω2 and generates an outgoing field of frequency ωSFG= ω1 ...
  34. [34]
    Preferred orientations of organic cations at lead-halide perovskite ...
    In this study, we access molecular dynamics at perovskite interfaces by monitoring the alignment of cations in methylammonium (MA) lead bromide perovskite.Missing: electronic | Show results with:electronic
  35. [35]
    Polarization‐Dependent Sum‐Frequency‐Generation Spectroscopy ...
    Apr 4, 2023 · It is demonstrated that polarization‐dependent sum‐frequency‐generation (SFG) spectroscopy reveals the average surface structure and shape of the nanoparticles.
  36. [36]
    Revealing molecular conformation–induced stress at embedded ...
    Apr 14, 2021 · Revealing molecular conformation–induced stress at embedded interfaces of organic optoelectronic devices by sum frequency generation ...