Fact-checked by Grok 2 weeks ago

Second-harmonic generation

Second-harmonic generation (SHG), also known as frequency doubling, is a nonlinear optical process in which two of the same interact with a nonlinear lacking inversion to produce a single with twice the (and half the ) of the input . This χ^(2)-mediated phenomenon arises from the in the 's response to the , enabling efficient conversion under phase-matched conditions. SHG was first experimentally observed in 1961 by Peter A. Franken and colleagues, who directed a beam (at 694 nm) through a crystal and detected the generated second harmonic at 347 nm, marking the birth of shortly after the laser's invention. Subsequent advancements, including the development of phase-matching techniques by Robert C. Miller in 1962 using birefringent crystals like , dramatically improved conversion efficiencies, enabling practical applications. The process is inherently coherent and depends on the material's second-order nonlinear susceptibility tensor, which vanishes in centrosymmetric media, restricting SHG to non-centrosymmetric crystals such as KDP, BBO, or LBO. Quasi-phase matching, first proposed in 1962 and practically implemented via periodic poling in the late 1980s, further enhances efficiency in waveguides and thin films by compensating for phase mismatch. Key to SHG's utility is its role in frequency conversion for lasers, where infrared output (e.g., 1064 nm from Nd:YAG) is doubled to (532 nm) for high-power applications in , medical procedures, and displays. In , SHG provides label-free, high-resolution imaging of non-centrosymmetric structures like fibers in biological tissues, offering advantages over by avoiding and . Beyond these, SHG serves in surface-sensitive to probe interfaces and thin films, and in emerging nanophotonic devices for integrated all-optical . Conversion efficiencies can exceed 50% in optimized setups, with ongoing research focusing on nanostructured materials to push limits toward unity.

Fundamentals

Definition and Mechanism

Second-harmonic generation (SHG) is a nonlinear optical process classified as a second-order nonlinearity, characterized by the interaction of two s at \omega to produce a single photon at frequency $2\omega within a suitable medium. This frequency doubling occurs through the material's second-order nonlinear susceptibility \chi^{(2)}, which enables the coherent conversion of the fundamental optical field into its harmonic. The process requires a non-centrosymmetric medium, as inversion symmetry in centrosymmetric crystals leads to the vanishing of \chi^{(2)} due to the odd nature of the second-order polarization response under parity transformation. At the microscopic level, the mechanism arises from the anharmonic response of the medium's electrons or lattice to the applied electric field E(\omega). The induced polarization includes a second-order term, P^{(2)}(2\omega) = \epsilon_0 \chi^{(2)} E^2(\omega), which oscillates at the harmonic frequency and acts as a source for the generated $2\omega field. This polarization drives dipole radiation at $2\omega, effectively combining the energies and momenta of the input photons. Energy conservation dictates that \omega_1 + \omega_2 = \omega_3, simplifying to \omega + \omega = 2\omega in the degenerate case where \omega_1 = \omega_2 = \omega, while momentum conservation requires \mathbf{k_1} + \mathbf{k_2} = \mathbf{k_3}, often necessitating phase-matching techniques for efficient conversion over macroscopic distances.

Nonlinear Optical Susceptibility

In nonlinear optics, the polarization \mathbf{P} induced in a dielectric medium by an applied electric field \mathbf{E} can be expressed as a power series expansion: \mathbf{P} = \epsilon_0 \left( \chi^{(1)} \mathbf{E} + \chi^{(2)} \mathbf{E}\mathbf{E} + \chi^{(3)} \mathbf{E}\mathbf{E}\mathbf{E} + \cdots \right), where \epsilon_0 is the vacuum permittivity and \chi^{(n)} denotes the nth-order susceptibility tensor. The second-order term, involving \chi^{(2)}, governs second-order nonlinear processes such as second-harmonic generation (SHG), where two photons at frequency \omega combine to produce one at $2\omega. This term arises from the anharmonic response of the medium's electrons and lattice to the driving field, leading to a quadratic contribution to the induced dipole moment. The second-order susceptibility \chi^{(2)}_{ijk}(\omega_3; \omega_1, \omega_2) is a third-rank tensor with 27 components in general, where the indices i, j, k correspond to Cartesian directions and the frequency arguments satisfy \omega_3 = \omega_1 + \omega_2. Intrinsic permutation symmetry reduces the number of independent components to 18, as \chi^{(2)}_{ijk}(\omega_3; \omega_1, \omega_2) = \chi^{(2)}_{ikj}(\omega_3; \omega_2, \omega_1). Crystal symmetry further constrains the tensor; for example, in cubic class $43m (as in GaAs), only one independent component remains. In non-dispersive media or far from resonances, Kleinman symmetry applies, permitting full permutation of indices and frequencies, reducing independent components to 10 or fewer. This symmetry, derived from assuming negligible dispersion and damping, simplifies calculations but fails near electronic or vibrational resonances. In the International System of Units (SI), \chi^{(2)} has dimensions of meters per volt (m/V), reflecting the quadratic dependence of polarization on field strength. In electrostatic units (esu), values are often reported in statvolts/cm or equivalent, with conversion factors accounting for the Gaussian system's differences. Typical magnitudes for inorganic nonlinear crystals range from $10^{-12} to $10^{-9} m/V; for instance, potassium dihydrogen phosphate (KDP) exhibits d_{36} = 0.39 pm/V (where d_{ijk} = \chi^{(2)}_{ijk}/2), while beta-barium borate (BBO) reaches d_{22} = 2.2 pm/V, enabling efficient SHG. The tensor components of \chi^{(2)} are dispersive, varying with the frequencies \omega_1, \omega_2, due to the medium's electronic structure and lattice vibrations. Near electronic transitions or band edges, resonant enhancements can increase \chi^{(2)} by orders of magnitude; for example, in semiconductors like ZnTe, a strong rise occurs above the E_0 bandgap, attributed to virtual excitations of electrons to conduction bands. This \omega-dependence must be considered for applications, as it influences phase-matching bandwidth and conversion efficiency in SHG. Experimental determination of \chi^{(2)} relies on techniques that isolate the nonlinear polarization response. The Maker fringes method, introduced in early SHG studies, measures the second-harmonic as the sample is rotated relative to the incident beam, producing fringes due to varying from phase mismatch. By comparing fringe patterns to known references like , absolute values and tensor ratios are extracted, accounting for and effects. This approach has been refined for thin films and biaxial , providing precise characterization essential for device design.

Historical Development

Early Discovery

The invention of the laser by Theodore Maiman in 1960 provided the intense, coherent light sources necessary to explore nonlinear optical effects, marking the emergence of nonlinear optics as a field. Second-harmonic generation (SHG) was first experimentally observed in 1961 by Peter A. Franken, A. E. Hill, C. W. Peters, and G. Weinreich at the University of Michigan. They directed a pulsed ruby laser beam with a fundamental wavelength of 694 nm into a quartz crystal, detecting the second harmonic at 347 nm using photographic plates after long exposure times. This demonstration required the high peak powers (around 3 kW) from pulsed operation of the ruby laser, as continuous-wave sources lacked sufficient intensity for observable nonlinear effects. Early detection of SHG faced significant challenges due to its extremely low conversion efficiency, on the order of 10^{-8}, necessitating sensitive detection methods and careful control of experimental conditions to distinguish the weak signal from . The discovery was swiftly confirmed by independent groups in 1962, including experiments by R. W. Terhune, P. D. Maker, and C. M. Savage, who observed SHG in crystals using similar setups. These confirmations extended observations to other materials, solidifying SHG as a reproducible nonlinear . In the same year, J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan provided the initial theoretical framework, interpreting SHG through the nonlinear induced in the medium by the intense electric field of the .

Key Theoretical and Experimental Advances

Following the initial observation of second-harmonic generation (SHG) in , significant theoretical advancements in phase-matching concepts emerged in 1962, primarily through the work of J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan. Their seminal paper introduced the general theory of phase matching for nonlinear optical interactions, demonstrating that efficient energy transfer in SHG requires the wave vectors of the fundamental and harmonic fields to satisfy Δk = 0, where Δk = k_{2ω} - 2k_ω. This framework predicted that birefringent materials could achieve phase matching by compensating for through angular tuning, enabling conversion efficiencies up to 50% in ideal conditions without walk-off losses. Building on this theory, the first experimental demonstration of angle-tuned birefringent phase matching was achieved in 1962 by Robert C. Miller in uniaxial ferroelectric crystals such as (ADP) and potassium dihydrogen phosphate (KDP). Subsequent work in the mid-1960s by researchers like P. D. Maker and R. W. Terhune further demonstrated efficient SHG in KDP crystals by orienting the optic axis at specific angles relative to the propagation direction, achieving type I phase matching for 1.06 μm Nd:YAG laser fundamental to 532 nm second harmonic with efficiencies exceeding 10% in centimeter-long crystals. These experiments validated the theoretical predictions and established birefringent crystals as practical media for high-power frequency doubling, paving the way for applications in laser systems. In the 1970s, the concept of quasi-phase matching (QPM) was further refined, although originally proposed in 1962 by Armstrong et al. as an alternative to using periodically reversed nonlinear coefficients to periodically reset the phase mismatch. Theoretical extensions in this decade, including detailed analyses by R. H. Stolen, explored QPM in ferroelectric materials like (LiNbO₃), predicting that periodic poling with domain periods on the order of 10-30 μm could enable non-birefringent phase matching across a broad range. Experimental realization lagged until the late and , but these theoretical works laid the groundwork for achieving over 50% efficiency in QPM-SHG devices. The 1980s marked key experimental progress driven by advancements in pulsed laser sources, which dramatically increased peak intensities for SHG while mitigating thermal effects in crystals. Mode-locked Nd:YAG and dye lasers, with picosecond pulse durations and peak powers exceeding 1 GW/cm², enabled efficient SHG in KDP and LiNbO₃, achieving conversion efficiencies up to 70% for green light generation at repetition rates of 80 MHz. These developments, exemplified by systems from companies like Coherent and Spectra-Physics, shifted SHG from continuous-wave to pulsed regimes, supporting ultrafast spectroscopy and high-repetition-rate applications. Post-2000 advancements have integrated SHG with nanostructures and ultrafast , enhancing efficiency and enabling novel applications in pulse generation via frequency upconversion.

Phase Matching Techniques

Critical Phase Matching

Critical phase matching is a technique employed in uniaxial anisotropic crystals to achieve efficient second-harmonic generation (SHG) by leveraging , where the refractive index n_o differs from the index n_e. In this method, the crystal's orientation is adjusted such that the direction of the wave at \omega forms an angle \theta with the optic axis, aligning the wave vectors \mathbf{k}_\omega and \mathbf{k}_{2\omega} to satisfy the phase-matching condition \Delta \mathbf{k} = 0. This angular tuning compensates for the mismatch between the and wavelengths, enabling collinear of the and rays in type I or type II configurations. The phase mismatch parameter is defined as \Delta k = k_{2\omega} - 2k_\omega = \frac{2\omega n_{2\omega}}{c} - \frac{2 \omega n_\omega}{c}, where k = \frac{\omega n}{c} is the wave number, n is the , and c is the in . At the optimal phase-matching angle \theta_{pm}, \Delta k = 0, which occurs when the birefringence-induced variation in the effective for the extraordinary wave balances the material's chromatic . For example, in \beta- (BBO) for type I SHG of 1064 light, \theta_{pm} \approx 22.8^\circ. A significant limitation in uniaxial crystals arises from Poynting vector walk-off, where the (direction of energy flow) for the extraordinary ray deviates from its by a walk-off \rho, causing spatial separation between the ordinary fundamental beam and the extraordinary harmonic beam. This divergence reduces the effective interaction length L_{eff} below the physical crystal length L, as the beams overlap only over a distance L_{eff} \approx w / \tan \rho, where w is the beam waist , thereby degrading conversion efficiency particularly for focused beams. In BBO at 1064 nm, the walk-off is approximately 3.2° for type I phase matching. The phase-matching angle \theta_{pm} is sensitive to temperature and wavelength variations, as these alter the refractive indices through thermo-optic and dispersive effects. Temperature tuning shifts \theta_{pm} due to changes in birefringence, often requiring precise control to maintain \Delta k = 0, while wavelength detuning from the design value broadens the mismatch. Consequently, critical phase matching exhibits narrow acceptance bandwidths: the angular acceptance \Delta \theta (full angle at half-maximum efficiency) is typically on the order of 0.1°–1° for common crystals, and the spectral bandwidth \Delta \lambda is similarly limited. For BBO in type I SHG at 1064 nm, the angular acceptance is about 1.2 mrad·cm, corresponding to \Delta \theta \approx 0.07^\circ for a 1 cm crystal length.

Non-critical Phase Matching

Non-critical phase matching (NCPM) achieves the phase-matching condition Δk = 0 for second-harmonic generation by propagating the beams orthogonal to the crystal's optic at 90°, eliminating the need for angular tuning and leveraging temperature or wavelength adjustments in suitable birefringent materials such as lithium triborate (LBO). This configuration exploits differences in the temperature coefficients of the refractive indices (dn/dT) between the and polarizations, allowing the effective indices at the fundamental and harmonic wavelengths to align precisely through thermal tuning. In contrast to critical phase matching, which relies on at non-90° angles and suffers from walk-off, NCPM ensures collinear propagation of and waves with identical refractive indices, resulting in zero walk-off angle. The absence of walk-off in NCPM enables the use of longer interaction lengths in the crystal without beam separation, facilitating tighter focusing and substantially higher conversion efficiencies, particularly in high-power applications where spatial overlap is critical. Optimized extracavity setups have demonstrated efficiencies exceeding 80% for pulsed operation in KTP crystals using critical phase matching. For LBO, pulsed SHG efficiencies surpass 70%, benefiting from the technique's wide acceptance angles and damage resistance. A representative example is the frequency doubling of 1064 nm Nd:YAG laser light to 532 nm in LBO, achieved via type I NCPM at temperatures around 148 °C, where thermal tuning compensates for dispersion to satisfy Δk = 0. NCPM typically exhibits a bandwidth of 1–10 °C for 1 cm lengths, reflecting its sensitivity to thermal uniformity but providing stable operation within this range. Its wavelength acceptance is narrower compared to critical matching, limiting applications but ideal for narrow-linewidth sources like single-frequency lasers. This makes NCPM particularly advantageous for efficient, high-peak-power SHG in systems requiring minimal beam distortion and maximal nonlinear overlap.

Theoretical Frameworks

Plane Wave Derivation at Low Conversion

The plane wave derivation for second-harmonic generation (SHG) under the low-conversion regime begins with in a nonlinear medium, where the nonlinear at the second-harmonic frequency is given by \mathbf{P}^{(2\omega)} = \epsilon_0 \chi^{(2)} E_\omega^2. This acts as a source term in the wave equation for the second-harmonic field, \nabla^2 \mathbf{E}_{2\omega} + \frac{(2\omega)^2}{c^2} n_{2\omega}^2 \mathbf{E}_{2\omega} = -\frac{(2\omega)^2}{c^2} \mathbf{P}^{(2\omega)}, assuming a non-magnetic medium with refractive index n_{2\omega}. For plane waves propagating along the z-direction, the fields are expressed using slowly varying envelope approximations: E_\omega(z, t) = A_\omega(z) e^{i(k_\omega z - \omega t)} + \text{c.c.} and E_{2\omega}(z, t) = A_{2\omega}(z) e^{i(k_{2\omega} z - 2\omega t)} + \text{c.c.}, where the envelopes A_j(z) vary slowly compared to the optical wavelengths (|dA_j/dz| \ll k_j |A_j|). Substituting these into the wave equation and neglecting second derivatives of the envelopes yields the coupled amplitude equations:
\frac{dA_{2\omega}}{dz} = i \frac{\omega d_\text{eff}}{n_{2\omega} c} A_\omega^2 e^{-i \Delta k z},
\frac{dA_\omega}{dz} = i \frac{\omega d_\text{eff}}{n_\omega c} A_\omega^* A_{2\omega} e^{i \Delta k z},
with phase mismatch \Delta k = 2k_\omega - k_{2\omega} and effective nonlinear coefficient d_\text{eff} = \chi^{(2)}/2.
In the undepleted pump approximation, valid for low conversion efficiencies (<1%), the fundamental amplitude is treated as constant (dA_\omega/dz \approx 0), so A_\omega(z) \approx A_\omega(0). Integrating the equation for A_{2\omega} from z = 0 to L (crystal length), with initial condition A_{2\omega}(0) = 0, gives
A_{2\omega}(L) = i \frac{\omega d_\text{eff}}{n_{2\omega} c} A_\omega^2(0) L \cdot \text{sinc}\left( \frac{\Delta k L}{2} \right) e^{-i \Delta k L / 2}.
The resulting second-harmonic intensity is
I_{2\omega}(L) = \frac{2 \omega^2 d_\text{eff}^2 L^2}{n_\omega^2 n_{2\omega} c^3 \epsilon_0} I_\omega^2(0) \cdot \text{sinc}^2\left( \frac{\Delta k L}{2} \right),
where I_j = \frac{1}{2} n_j \epsilon_0 c |A_j|^2 relates the intensity to the envelope amplitude, assuming SI units and isotropic indices for simplicity.
The \text{sinc}^2(\Delta k L / 2) dependence arises from the coherent buildup of the second-harmonic field, with maximum efficiency at phase matching (\Delta k = 0), where \text{sinc}(0) = 1. The coherence length is defined as L_c = \pi / |\Delta k|, the distance over which the phase mismatch causes the generated fields to dephase by \pi, limiting efficiency for L > L_c. This derivation assumes monochromatic plane waves of infinite transverse extent, negligible or beyond the phase mismatch, and low conversion to justify the undepleted approximation.

Plane Wave Derivation with Depletion

In the approximation for second-harmonic generation (SHG), the low-conversion regime assumes negligible depletion, treating the wave amplitude as constant. To describe high-conversion scenarios where significant energy transfer occurs, the coupled wave equations must account for the back-action of the generated second-harmonic on the . These equations, derived from under the for collinear propagation, are given by \frac{dA_{2\omega}}{dz} = i \kappa A_{\omega}^2 e^{-i \Delta k z}, \frac{dA_{\omega}}{dz} = -i \frac{\omega d_{\mathrm{eff}}}{n_{\omega} c} A_{2\omega} A_{\omega}^* e^{i \Delta k z}, where A_{\omega} and A_{2\omega} are the complex slowly varying envelope amplitudes of the fundamental and second-harmonic waves, respectively, \Delta k = k_{2\omega} - 2k_{\omega} is the phase mismatch, n_{\omega} is the refractive index at the fundamental frequency, c is the speed of light in vacuum, and the coupling uses the effective second-order nonlinear coefficient d_{\mathrm{eff}}. The coefficients ensure proper energy scaling for the frequency-doubling process. These coupled differential equations reflect the interaction, where the second-harmonic grows at the expense of the , leading to depletion. A key consequence is the Manley-Rowe power conservation relation, which follows directly from the equations by considering the time-averaged power flow P_{\omega} \propto |A_{\omega}|^2 and P_{2\omega} \propto |A_{2\omega}|^2. For SHG, this yields P_{\omega}(z) + P_{2\omega}(z) = P_{\omega}(0), indicating that the power lost from the equals the power gained by the second harmonic, consistent with while accounting for the annihilation of two photons per second-harmonic created. For the ideal case of perfect phase matching (\Delta k = 0), the coupled equations admit an exact analytical solution assuming initial conditions A_{2\omega}(0) = 0 and arbitrary A_{\omega}(0). The solution is A_{2\omega}(z) = i A_{\omega}(0) \sin(\Gamma z), A_{\omega}(z) = A_{\omega}(0) \cos(\Gamma z), where \Gamma = \frac{\omega d_{\mathrm{eff}}}{n c} |A_{\omega}(0)|, assuming average refractive index n. This trigonometric behavior shows oscillatory energy exchange, with complete pump depletion achievable at z = \pi/(2\Gamma), where 100% conversion efficiency is theoretically possible in the absence of losses or other limitations. The conversion efficiency \eta = |A_{2\omega}(z)/A_{\omega}(0)|^2 thus reaches unity periodically, highlighting the potential for efficient frequency doubling in phase-matched media. When phase mismatch is present (\Delta k \neq 0), no closed-form analytical solution exists for the general case, and the coupled equations must be integrated numerically, such as via . Numerical solutions reveal oscillatory power transfer between the waves, modulated by a \mathrm{sinc}^2(\Delta k z / 2) envelope, with an optimal crystal length for maximum efficiency near z \approx \pi / |\Delta k|. These oscillations arise from the phase accumulation, and the peak conversion decreases with increasing \Delta k, emphasizing the need for to achieve high efficiencies. This derivation assumes collinear of infinite plane waves in a lossless, homogeneous nonlinear medium, neglecting birefringent walk-off, , or effects, which are valid only for focused beams or thin where such approximations hold.

Gaussian Beam Expressions

In the context of second-harmonic generation (SHG), the plane-wave approximation is extended to realistic sources by incorporating profiles, which account for , focusing, and spatial variations along the direction. A is characterized by its beam waist w_0 at the focus, with the spot size varying as w(z) = w_0 \sqrt{1 + (z/z_R)^2}, where z_R = \pi w_0^2 / [\lambda](/page/Lambda) is the range and \lambda is the of the . The also includes a Gouy shift \eta(z) = \arctan(z/z_R), which arises from the of the and contributes to the overall mismatch in nonlinear interactions. These parameters are essential for modeling focused beams in , as tight focusing enhances intensity but introduces spatial nonuniformity that modifies the SHG efficiency compared to uniform plane waves. Focusing effects alter the effective nonlinear coefficient d_{\text{eff}} through a dimensionless parameter h(\xi, \rho), which integrates the beam's transverse and longitudinal variations. Here, \xi = L / z_R quantifies the crystal length L relative to the Rayleigh range, while \rho = L / (2 z_{R\omega}) incorporates the walk-off due to birefringence, with z_{R\omega} denoting the Rayleigh range at the fundamental frequency \omega. In the Boyd-Kleinman theory, the SHG power is proportional to |h(\xi, \rho)|^2, capturing how focusing amplifies the nonlinear overlap while walk-off reduces it by spatially separating the interacting fields. For low conversion efficiency, this leads to an enhancement factor over plane-wave limits, with h(\xi, 0) \approx 0.69 at optimal focusing (\xi \approx 2.84) in the absence of walk-off. When phase matching is considered alongside walk-off, the efficiency \eta is evaluated via the Boyd-Kleinman integral, which effectively averages the phase mismatch \Delta k over the beam path: \eta \propto \int \text{sinc}^2(\Delta k_{\text{eff}} l) \, dl, where \Delta k_{\text{eff}} includes contributions from the Gouy phase and . This integral form accounts for the finite interaction length within the focused region, broadening the acceptance bandwidth for \Delta k compared to plane waves. In the absence of phase matching, tight focusing (\xi \ll 1) results in a short dominated by the Rayleigh range, yielding broad \Delta k acceptance and an efficiency approximating the undepleted limit \eta \approx (d_{\text{eff}} L / \lambda)^2. For phase-matched conditions, the optimal focus position shifts the beam waist inside the to maximize gain, balancing spread and walk-off losses. In non-walk-off cases (\rho = 0), the is ideally at the crystal center; with walk-off, it moves toward the input face, as quantified by empirical fits like \xi_m \approx 1.41 for moderate , enhancing efficiency by up to 1.5 times over end-focused setups. These spatial optimizations are critical for high-power applications, where Gaussian profiles ensure efficient mode preservation in the generated beam.

Configurations and Types

Bulk Crystal Second-Harmonic Generation

Bulk crystal second-harmonic generation involves the interaction of a fundamental beam with a nonlinear where the second-harmonic wave is generated within the volume of the material, typically through collinear propagation along the crystal's principal axes. In uniaxial crystals, the (o) and (e) polarizations are exploited, with input polarizers used to select the desired interaction , such as ooe for type II processes or eee in certain biaxial configurations where all waves are extraordinary. Biaxial crystals allow more flexible phase-matching directions due to their three distinct refractive indices, enabling collinear setups that align the beam with the principal planes for efficient energy transfer. For type I phase matching (o + o → e), the crystal is cut such that the optic axis lies in the perpendicular to the propagation direction, with θ between the optic axis and the determined by the n_e^{2ω}(θ) = n_o^ω to satisfy Δk = 0, often around 30–50° depending on the and material . In contrast, type II phase matching (o + e → e) requires a cut where the optic axis is oriented to allow one and one fundamental , typically involving a 45° polarization rotation of the input relative to the principal axes, resulting in θ angles shifted by 10–20° from type I for the same . These cut angles are precisely engineered during to optimize compensation over the interaction length, ensuring maximum L_c = π / Δk. Biaxial crystals extend this to non-principal directions, using vector projections to identify viable θ and φ angles for collinear propagation. Efficiency in bulk SHG is enhanced for continuous-wave operation through multi-pass designs, where the fundamental beam is reflected back through the crystal multiple times using mirrors, increasing the effective interaction length without excessive . Resonant cavities further boost by building up intracavity intensity, achieving normalization factors up to 10^4 times higher than single-pass setups for low-power inputs. These approaches are particularly vital for sources, where single-pass efficiencies are limited to below 1% without enhancement. At high pump powers, bulk SHG suffers from losses including linear of both fundamental and harmonic waves, which dissipates energy as heat, and from crystal imperfections that broadens the . Thermal lensing arises from temperature gradients inducing changes, effectively creating a dynamic that can defocus the and reduce phase-matching tolerance if power exceeds 10–100 W/cm². These effects limit scalable output, often requiring or shorter crystals to mitigate. A key output of bulk crystal SHG is the production of coherent ultraviolet or infrared radiation, exemplified by the generation of 532 nm green light from the 1064 nm output of a Nd:YAG laser, enabling high-brightness visible sources with efficiencies up to 50% in optimized type I configurations.

Surface Second-Harmonic Generation

Surface second-harmonic generation (SHG) occurs at interfaces where the inversion symmetry of the bulk material is broken, resulting in a non-zero second-order nonlinear susceptibility tensor, denoted as \chi^{(2)}, even in centrosymmetric media. This symmetry breaking permits electric dipole contributions to the nonlinear polarization that are otherwise forbidden in the bulk, making surface SHG a sensitive probe of interfacial properties. The surface \chi^{(2)} arises primarily from the altered electronic structure at the interface, where dangling bonds and modified density of states enhance the nonlinear response. In centrosymmetric materials, the absence of bulk SHG underscores the interface's role, while in non-centrosymmetric materials, the surface contribution often dominates due to enhancement factors of 10–100 times over bulk values, attributed to localized electronic states. For planar surfaces, the generated second-harmonic field is modulated by Fresnel transmission and reflection coefficients, which describe the continuity of the electromagnetic fields across the boundary and influence the effective nonlinear . In total internal reflection configurations, evanescent fields penetrate only a few hundred nanometers into the rarer medium, confining the interaction to and enabling enhanced SHG without bulk interference. For non-planar surfaces, such as nanoparticles, rough interfaces, or colloidal particles, enhancements further amplify the SHG signal. In nanoparticles, Mie resonances—stemming from the of scattered electric and magnetic multipoles—can boost the s at the , leading to quadratic enhancements in the nonlinear output. studies reveal the symmetry: isotropic surfaces typically exhibit dominant SHG in the p-in/p-out , where both input and output fields are p-polarized relative to the . Adsorbates introduce additional selection rules governed by the molecular , restricting certain \chi^{(2)} tensor elements and allowing determination of adsorbate orientation and coverage. Detection of surface SHG signals is challenging due to their weakness, typically representing about $10^{-12} of the input intensity, corresponding to cross sections where only one SH photon is generated per $10^{12}–$10^{13} incident photons. Despite this, the technique's surface specificity enables monolayer-sensitive measurements, such as probing adsorbate layers or buried interfaces, without contributions from the bulk.

Radiation Patterns

Angular Distribution in Phase-Matched Systems

In phase-matched second-harmonic generation (SHG), the far-field distribution of the generated second-harmonic (SH) is determined by the of the spatial source within the nonlinear , resulting in a sinc-like pattern centered along the phase-matched direction. This pattern arises from the phase-matching , \mathrm{sinc}\left(\frac{\Delta k L}{2}\right), where \Delta k is the wavevector mismatch that varies with output angle, and L is the crystal length, leading to a narrow angular lobe with width inversely proportional to L. The central maximum corresponds to perfect phase matching, while side lobes represent off-axis contributions, making this distribution essential for evaluating beam collimation and into optical systems. Birefringent walk-off in critically phase-matched configurations introduces asymmetry into the angular pattern, broadening the distribution and creating uneven lobes due to the divergence of the extraordinary-ray (e-ray) component. In type-I SHG, where the beam is ordinary-polarized and the SH is extraordinary, the of the e-ray deviates from its wavevector by the walk-off angle \rho, causing spatial separation that elongates the effective in one transverse direction and skews the far-field lobes toward the walk-off plane. This effect is particularly pronounced in materials like beta-barium borate (BBO), where \rho \approx 3^\circ at 1064 nm, reducing quality for high-power applications unless compensated by geometries or dual-crystal setups. In non-collinear SHG configurations, especially type-I interactions in uniaxial crystals, the SH output often manifests as a conical pattern, where the SH beam propagates at an angle to the , forming a ring in the far field. This arises from the momentum conservation allowing non-collinear matching, with the cone angle tunable by the internal non-collinearity \alpha between the beams. The angular acceptance bandwidth, \Delta \theta \sim \sqrt{\lambda / L}, governs the width of the cone, where \lambda is the , enabling broadband operation but limiting for tightly focused . Such patterns are exploited in applications requiring angularly dispersed SH light, like ultrafast pulse . Diffraction from finite-sized input beams further modulates the SH angular distribution, producing an pattern at the SH $2\omega in the far field due to the circular of the beam . The central spot size scales as $1.22 \lambda / D with D the effective diameter, but at $2\omega, the shorter halves this limit compared to the fundamental, enhancing in applications while introducing ring-like that can overlap in high-numerical-aperture setups. Experimental profiling of these angular distributions typically employs (CCD) cameras placed in the far field to capture two-dimensional intensity maps, allowing direct visualization of lobes, cones, and asymmetries. Alternatively, slit-scan techniques traverse the beam profile linearly, integrating intensity along one axis to quantify broadening or acceptance angles with sub-degree resolution, often combined with rotational stages for polar mapping. These methods confirm theoretical predictions and assess phase-matching quality in during crystal alignment.

Factors Influencing Beam Profiles

In second-harmonic generation (SHG), spatial hole burning arises from patterns formed by standing within the nonlinear medium, particularly in intracavity configurations where counterpropagating fundamental interact, leading to nonuniform depletion of the gain medium and distortions in the generated second-harmonic beam profile. This effect creates periodic intensity variations that reduce the uniformity of the SHG output, as the modulates the local intensity of the fundamental beam, causing selective conversion in high-intensity regions. In multimode lasers with intracavity doubling, spatial hole burning further couples nonlinearly with the resonator modes, exacerbating beam asymmetry and limiting power scaling. Thermal effects in SHG crystals can induce self-focusing or defocusing of the due to at the second-harmonic (2ω), which generates heat and alters the via the effect, creating a lens that distorts the near-field profile. In high-power continuous-wave operations, of the generated 2ω light leads to temperature gradients, with positive or negative lensing depending on the sign of the thermo-optic coefficient, potentially causing breakup or reduced conversion efficiency if unmitigated. For type-II phase-matched crystals, these self-induced effects are particularly pronounced in double-pass configurations, where the focal length of the lens scales inversely with pump power, impacting the overall . Birefringent walk-off in anisotropic crystals during SHG causes spatial separation between the ordinary and extraordinary polarization components of the interacting beams, leading to an elliptical distortion in the second-harmonic beam profile and degradation of the M² beam quality factor. This walk-off effect accumulates over the crystal length, reducing the effective interaction volume and increasing the M² value in the walk-off plane, often from near 1.3 for the fundamental to higher values in the harmonic output, thereby broadening the near-field intensity distribution. In critically phase-matched setups, the angular dependence of walk-off further compromises beam symmetry, with compensation schemes required to maintain M² below 1.5 for practical applications. For ultrashort pulses, group velocity mismatch (GVM) between the fundamental and second-harmonic waves causes temporal walk-off, which couples with spatial effects to broaden the near-field profile of the generated pulse through spatiotemporal distortions. In thick crystals, this mismatch limits the interaction length to L_max ≈ τ |v_g^{-1}(ω) - v_g^{-1}(2ω)|^{-1}, where τ is the pulse duration, resulting in asymmetric spatial spreading as different pulse portions convert at varying positions. For pulses in birefringent media, GVM-induced broadening can increase the SHG beam waist by up to 20-30% compared to transform-limited cases, degrading spatial coherence. To mitigate these distortions and achieve uniform beam profiles, aperiodic poling designs in quasi-phase-matched crystals compensate for phase mismatches across the beam, reducing spatial variations from walk-off and GVM by tailoring the poling for uniformity. geometries, where the poling period varies linearly across the crystal aperture, further counteract birefringent walk-off by aligning local phase matching to the beam's , preserving close to unity and enabling efficient, distortion-free SHG in high-power systems. These approaches have demonstrated near-Gaussian profiles with < 1.2 in periodically poled lithium niobate for visible output.

Materials and Selection Criteria

Common Nonlinear Materials

Inorganic crystals are among the most prevalent materials for second-harmonic generation (SHG) due to their robust optical properties. Beta-barium borate (BBO, \beta-BaB_2O_4) possesses an effective nonlinear coefficient d_{\mathrm{eff}} \approx 2 pm/V for Type I SHG, enabling efficient frequency conversion. It offers transparency from 190 nm to 3500 nm, supporting UV generation down to approximately 200 nm, and accommodates Type I phase matching for fundamental wavelengths in the 400–900 nm range. BBO crystals exhibit a high damage threshold of about 10 GW/cm² for 100 ps pulses at 1064 nm. Lithium triborate (LBO, LiB_3O_5) is another key inorganic crystal, featuring a broad transparency window from 160 nm to 2600 nm and a notably high damage threshold exceeding 25 GW/cm² for nanosecond pulses at 1064 nm. Its effective nonlinear coefficient reaches d_{\mathrm{eff}} \approx 0.85 pm/V for SHG at 1064 nm, with capabilities for both Type I and Type II phase matching across a wide spectral range. LBO demonstrates excellent optical homogeneity, with refractive index variation \delta n \approx 10^{-6}/cm. Potassium dihydrogen phosphate (KDP, KH_2PO_4) serves as a cost-effective inorganic option for SHG applications, with a nonlinear coefficient d_{36} = 0.39 pm/V and transparency extending from 180 nm to 1550 nm. It supports Type I phase matching effectively for visible and near-UV wavelengths, though its hygroscopic nature necessitates protective coatings or sealed environments to prevent moisture-induced degradation. KDP is particularly noted for its large crystal sizes achievable through growth processes.
MaterialChemical Formulad_{\mathrm{eff}} (pm/V)Transparency Range (nm)Phase-Matching Example
BBO\beta-BaB_2O_4~2 (Type I SHG)190–3500Type I, 400–900 nm fundamental
LBOLiB_3O_5~0.85 (at 1064 nm)160–2600Type I/II, broad UV-VIS-NIR
KDPKH_2PO_40.39 (d_{36})180–1550Type I, visible-UV
Organic materials offer high nonlinear susceptibilities but often face challenges in mechanical stability. Urea (CO(NH_2)_2) displays a significant second-order susceptibility \chi^{(2)}, making it a standard reference for powder SHG efficiency measurements, with transparency in the UV-visible range and phase-matching suitability for fundamental wavelengths around 1064 nm. However, urea crystals exhibit poor mechanical properties, limiting their use in high-power applications. 2-(N,N-Dimethylamino)-5-nitroacetanilide (DAN) achieves high SHG efficiency due to its large \chi^{(2)} tensor components, supporting efficient conversion in the visible spectrum, though it suffers from suboptimal mechanical durability and crystal quality issues. Semiconductors extend SHG capabilities into the mid-infrared regime. Gallium arsenide (GaAs) is utilized for mid-IR SHG, leveraging its low absorption beyond 1 \mum and support for quasi-phase-matched configurations, with transparency from about 0.9 \mum to 17 \mum. Zinc selenide (ZnSe) similarly enables mid-IR frequency doubling, offering high thermal conductivity and a transmission window from 0.6 \mum to 20 \mum, often employed in total internal reflection geometries for broadband operation. Both materials facilitate phase matching in the 1.7–5.3 \mum SHG output range when pumped by infrared sources. These nonlinear crystals are commonly fabricated through specialized techniques to ensure high optical quality. Borate-based materials like BBO and LBO are primarily grown via the flux method, involving high-temperature solutions of barium or lithium oxides in boric acid fluxes, yielding crystals up to several centimeters in size over weeks to months. Phosphate crystals such as KDP are produced by slow evaporation or temperature-controlled aqueous solution growth, allowing for rapid scaling to large boules. Hydrothermal methods, using pressurized aqueous solutions at elevated temperatures, are applied to certain inorganic and semiconductor crystals to minimize defects and enhance uniformity.

Properties and Performance Metrics

The selection of materials for second-harmonic generation (SHG) relies on key performance metrics that balance conversion efficiency, operational robustness, and practical constraints. A primary figure of merit for SHG efficiency in bulk crystals is given by \frac{d_{\text{eff}}^2}{n^3}, where d_{\text{eff}} is the effective second-order nonlinear coefficient and n represents the refractive indices at the fundamental and harmonic frequencies; this metric quantifies the nonlinear drive relative to the material's dispersive losses, with higher values enabling more compact devices for a given power conversion. However, this must be traded off against the phase-matching bandwidth, inversely proportional to the tolerance in wavevector mismatch \Delta k^{-1}, which determines the spectral and angular acceptance for broadband or focused beams—materials with sharp dispersion may offer high FOM but limit operational flexibility in pulsed applications. Damage threshold is another critical metric, particularly for high-peak-power systems using nanosecond pulses, where thresholds exceeding 1 GW/cm² at 1064 nm are typically required to avoid catastrophic failure; for instance, β-BaB₂O₄ (BBO) achieves over 10 GW/cm² for pulses around 1 ns, supporting efficient SHG in high-energy without surface ablation. Temperature sensitivity impacts stability in non-critically phase-matched configurations, quantified by the tuning rate dT/d\lambda (or equivalently, the derivative of phase-matching temperature with respect to wavelength), which for in type-I SHG at 1064 nm to 532 nm is relatively low (on the order of 1 K/nm), allowing broad temperature bandwidths (around 4°C for a 1 cm crystal) but requiring precise control to maintain alignment under thermal drifts. Practical limitations further influence material choice, including hygroscopicity, cost, and size availability. Potassium dihydrogen phosphate (KDP) is deliquescent and hygroscopic, readily absorbing moisture in humid environments that degrades its surfaces and necessitates protective coatings or dry storage, whereas exhibits high chemical stability and non-hygroscopic behavior, enabling reliable operation in ambient conditions. KDP benefits from low production costs and scalability to large apertures (>50 mm) via solution growth, making it economical for high-power, large-beam applications, while LBO's flux growth limits sizes to ~20 mm and increases costs, restricting it to moderate-scale systems despite superior optical performance. Emerging materials address integration challenges in by engineering the second-order susceptibility \chi^{(2)} in polymer waveguides through poling or hybrid structures, offering compact SHG with efficiencies comparable to bulk crystals but enabling on-chip frequency conversion; for example, poled polymer claddings on waveguides achieve phase-matched SHG at wavelengths with low propagation losses (<1 dB/cm), paving the way for scalable integrated optics. Recent advances include enhanced SHG in 2D WS2/MoS2 interfaces and photo-induced nonlinearity in Si3N4 microresonators for wavelengths.

Applications

Frequency Doubling in Lasers

Second-harmonic generation (SHG) is widely employed in lasers to produce visible wavelengths, particularly green and blue light, by doubling the frequency of infrared fundamental beams from solid-state gain media such as or . This process enables compact, efficient sources that have largely replaced gas lasers like argon-ion systems in applications requiring high-brightness visible output. In laser systems, SHG can be implemented in extracavity or intracavity configurations, each offering trade-offs in efficiency, complexity, and power handling. Extracavity frequency doubling involves directing the output from the laser oscillator through a in a single pass, typically after amplification or . This setup simplifies alignment and allows independent optimization of the fundamental laser, achieving conversion efficiencies of 10-30% for nanosecond pulses. For instance, an 8 W, 10 kHz, 30 ns laser at 946 nm extracavity doubled in a yields 3 W at 473 nm, corresponding to an efficiency of about 37.5%, though typical values for 532 nm ns pulses from remain in the 10-30% range due to phase-matching constraints. Intracavity doubling integrates the nonlinear crystal within the laser resonator, leveraging the higher circulating intensity—often tens to hundreds of times the output power—for enhanced nonlinear interaction and overall efficiency. This configuration benefits from multiple passes through the crystal, reducing the required fundamental power while suppressing competing processes like sum-frequency generation. An example is a diode-pumped laser producing 1 W at 532 nm from an effective 5 W circulating at 1064 nm, with optical-to-optical efficiencies exceeding 20% under continuous-wave operation. Materials like are commonly used for critical phase matching in such setups. Despite these advantages, intracavity SHG in green lasers faces challenges, including the "green problem," which manifests as output instability, beam pointing errors, and intensity fluctuations due to spatiotemporal hole burning and weak coupling between orthogonal modes at 1064 nm. Thermal lensing in the gain medium and nonlinear crystal further exacerbates this, inducing phase mismatching and reducing efficiency, particularly at high powers where absorbed green light causes local heating. Commercially, diode-pumped solid-state (DPSS) lasers employing SHG dominate low-to-medium power applications, such as laser pointers and displays, with outputs from milliwatts to tens of watts at 532 nm using or crystals. For higher powers, quasi-phase-matching (QPM) in materials like enables scaling to kilowatt levels; thin-disk lasers with intracavity doubling have demonstrated 1.3 kW at 515 nm with efficiencies over 20%, suitable for industrial processing. SHG performance varies across pulse regimes: in continuous-wave (CW) operation, stable, low-noise outputs are prioritized, often limited to hundreds of watts to mitigate thermal effects, whereas mode-locked lasers generate femtosecond pulses with peak powers enabling near-unity conversion efficiencies due to enhanced nonlinearities. For example, mode-locked Nd:YVO₄ systems produce picosecond green pulses at 532 nm with >50% efficiency from 1064 nm fundamentals, contrasting with CW intracavity designs focused on average power stability.

Optical Microscopy Techniques

Second-harmonic generation (SHG) microscopy is a label-free nonlinear optical imaging technique that exploits the coherent second-order nonlinear susceptibility of non-centrosymmetric structures to generate contrast without the need for exogenous labels or fluorescent markers. Unlike two-photon excited fluorescence, SHG involves a parametric process where two incident at the fundamental wavelength combine to produce a single emitted at exactly twice the , with no net energy absorption or population of excited states in the sample. This makes it particularly suited for imaging endogenous biomolecules such as collagen fibrils and , which exhibit strong SHG signals due to their ordered, non-centrosymmetric architectures. The technique was first demonstrated for biological applications in by Freund and Deutsch, enabling high-contrast visualization of structural proteins in tissues. Typical SHG microscopy setups employ a femtosecond Ti:sapphire laser operating in the near-infrared range of 800–1000 nm, with pulse durations around 100 fs and repetition rates of ~80 MHz, to excite the sample while minimizing photodamage. The laser beam is focused through a high-numerical-aperture objective (NA > 1.2) in a scanning confocal configuration, allowing raster scanning for 2D or 3D image acquisition. The backward- or forward-scattered SHG signal, collected at half the excitation wavelength (400–500 nm), is detected using sensitive photomultiplier tubes (PMTs) equipped with bandpass filters to isolate the coherent emission from any autofluorescence. Polarization control of the incident beam further enhances the setup's capability to resolve molecular orientations. Key advantages of SHG microscopy include its intrinsic optical sectioning, providing axial on the order of 1 μm due to the nonlinear confinement of the volume, which is superior to wide-field methods and comparable to confocal without pinhole artifacts. The absence of or stems from the instantaneous, non-resonant nature of the process, enabling prolonged imaging sessions of live samples. Additionally, the directional dependence of SHG—stronger backward scattering from thin structures like and forward scattering from bulk —facilitates depth-resolved imaging by combining both detection geometries. In biological applications, SHG microscopy excels in tissue imaging, such as mapping organization in skin, tendons, and corneas to assess remodeling. It has proven valuable in cancer diagnostics by quantifying fiber alignment and density alterations in tumors, correlating these changes with progression and therapeutic response in models of and ovarian cancers. Polarization-resolved SHG further reveals molecular orientations, such as the helical pitch in or polarity in networks during , providing insights into biomechanical properties without invasive labeling. Advances in the 2020s have pushed SHG microscopy toward real-time capabilities, with video-rate imaging achieved through techniques like interferometric SHG and resonant scanning endoscopes, enabling dynamic observation of physiological processes at 30 frames per second over fields of view up to 500 μm. These developments, integrated with fiber-optic probes, have expanded applications to in vivo microendoscopy for non-invasive tissue monitoring.

Material Characterization Methods

Second-harmonic generation (SHG) serves as a powerful, non-destructive for characterizing materials due to its sensitivity to non-centrosymmetric structures, enabling the probing of surfaces, interfaces, thin films, and crystalline domains without invasive labeling. In , SHG arises from the second-order nonlinear susceptibility tensor χ^(2), which vanishes in centrosymmetric bulk materials but persists at symmetry-broken interfaces or in non-centrosymmetric crystals, allowing selective detection of these features with sub-monolayer sensitivity. This makes SHG particularly valuable for assessing molecular orientation, , and structural order in materials ranging from semiconductors to biomolecules. One common method involves polarization-resolved SHG measurements, where the beam's is varied to map the χ^(2) tensor components, providing insights into the material's and orientation. For instance, in thin films, one-beam SHG techniques compare s- and p-polarized signals at fixed incidence angles, fitting data to extract χ^(2)_zzz values (e.g., 1.25 ± 0.10 pm/V for nanolaminate thin films such as In₂O₃/TiO₂/Al₂O₃), while accounting for substrate contributions via reference samples like . Rotational measurements, by rotating the sample under a focused , exploit fringes between film and signals to quantify χ^(2) in weakly nonlinear layers, revealing temporal walk-off effects (e.g., 45 fs in substrates) that influence signal phase. These approaches are advantageous for their simplicity and robustness, though they require careful calibration to mitigate errors from angle uncertainty or neglected . SHG microscopy extends these principles to spatially resolved , ideal for microstructural features in materials. In two-dimensional (2D) materials like dichalcogenides (TMDCs), polarization-dependent SHG reveals layer and stacking order, with signal intensity scaling as I_SHG ∝ (χ^(2))^2 I_ω^2, where I_ω is the fundamental intensity, enabling thickness determination via contrast analysis (e.g., monolayer vs. bilayer MoS2). For ferroelectric domains, SHG intensity variations track domain walls and strain, as demonstrated in crystals where optimized beam geometries enhance contrast for nanoscale . In organic and pharmaceutical crystals, SHG combined with coherent anti-Stokes Raman scattering (CARS) characterizes polymorphism and growth rates, with forward-scattered SHG highlighting non-centrosymmetric arrangements in microcrystals. Surface-specific applications include buried interface studies in silicon-on-insulator (SOI) wafers, where SHG detects and defects non-destructively, correlating signal enhancements to layer quality. For biomolecules and soft materials, SHG probes ordered assemblies like fibers, offering label-free imaging of fibrillar in tissues with high contrast due to coherent signal emission. Seminal work by and colleagues established SHG as a surface probe in the , demonstrating its use for adsorbate detection at liquid-solid interfaces with quadratic intensity dependence on coverage. Recent advances, such as high-throughput scanning SHG, automate of crystal in large-area samples, facilitating in nonlinear optical device fabrication. Overall, SHG's and selectivity position it as a complementary tool to linear , though quantitative analysis demands precise modeling of and absorption effects.

References

  1. [1]
    frequency-doubled laser, second-harmonic generation, SHG, design ...
    Definition: the phenomenon that an input wave in a nonlinear material can generate a wave with twice the optical frequency. Alternative term: second-harmonic ...What is Frequency Doubling? · Phase Matching · Second-harmonic Generation...
  2. [2]
    Second Harmonic Generation - an overview | ScienceDirect Topics
    Second harmonic generation (SHG) is defined as a nonlinear optical process in which photons interact with noncentrosymmetric media to produce new photons ...
  3. [3]
  4. [4]
    Optical second harmonic generation in anisotropic multilayers with ...
    Mar 29, 2024 · Optical second harmonic generation (SHG) refers to the nonlinear optical process where two photons of the same energy ( ) combine to generate a ...
  5. [5]
    High-Resolution Nonlinear Optical Imaging of Live Cells by Second ...
    In this paper we describe the use of surface second harmonic generation (SHG) in laser scanning microscopy as a new contrast mechanism for live cell imaging.
  6. [6]
    [PDF] The Nonlinear Optical Susceptibility - Elsevier
    Mar 1, 2014 · In Section 1.3 we show how to treat the vector nature of the fields; in such a case χ(1) becomes a second-rank tensor, χ(2) becomes a third-rank ...
  7. [7]
    Nonlinear Dielectric Polarization in Optical Media | Phys. Rev.
    Nonlinear Dielectric Polarization in Optical Media. D. A. Kleinman. Bell ... 126, 1977 – Published 15 June, 1962. DOI: https://doi.org/10.1103/PhysRev ...
  8. [8]
    The general failure of Kleinman symmetry in practical nonlinear ...
    Simply stated, Kleinman symmetry is defined as the interchangeability of all n indices in the rank n tensor χ(n) describing the macroscopic nonlinear ...Missing: susceptibility | Show results with:susceptibility
  9. [9]
    [PDF] Nonlinear Optical Materials
    The optical properties of some important crystals for use in second-order nonlinear optics are reviewed in Table 1. More extensive lists of second-order ...
  10. [10]
    [PDF] Dispersion of the second-order nonlinear susceptibility in ZnTe ...
    In the excitonic region of ZnTe only a weak resonance appears. Above the E0 band gap a strong increase of the SHG coefficient is observed, which is in.
  11. [11]
    Dispersion of the nonlinear susceptibility of and from second ...
    Dec 4, 2020 · Dispersion of the absolute second-order susceptibility of both M o ⁢ S 2 and W ⁢ S 2 is assessed on a wide spectral excitation range (710–1300 nm)
  12. [12]
  13. [13]
    Generation of Optical Harmonics | Phys. Rev. Lett.
    Oct 31, 2014 · In 1961, researchers showed that laser light could be converted from one color to another, the first nonlinear optical effect.Missing: observation | Show results with:observation
  14. [14]
    Optics Rotation Report for Anne Marie March - Stony Brook University
    The doubling process utilizes the nonlinear phenomenon called second harmonic generation ... SHG was first observed experimentally in 1961 by Peter A.
  15. [15]
    Optical Harmonic Generation in Calcite | Phys. Rev. Lett.
    Optical Harmonic Generation in Calcite. RW Terhune, PD Maker, and CM Savage Scientific Laboratory, Ford Motor Company, Dearborn, Michigan.Missing: SHG | Show results with:SHG
  16. [16]
    Interactions between Light Waves in a Nonlinear Dielectric
    Energy and power relationships are derived for the nonlinear dielectric which correspond to the Manley-Rowe relations in the theory of parametric amplifiers.
  17. [17]
    NONLINEAR OPTICAL FREQUENCY CONVERSION - AIP Publishing
    Birefringent phase matching. An essential step toward efficient frequency conversion, taken in 1962, was the use of birefringent phase matching in second ...
  18. [18]
    [PDF] periodically-poled lithium niobate Elliott J. Mason, III - DSpace@MIT
    Another means of phase matching was suggested by. Armstrong ... Multigrating quasi-phase-matched optical parametric oscillator in periodically poled LiNbO 3.
  19. [19]
    [PDF] pulsed lasers - Princeton University
    Pulsed lasers are versatile tools, useful for studying transient phenomena, and are easier to operate and have higher peak powers than continuous-wave lasers.Missing: advancements | Show results with:advancements
  20. [20]
    Nanostructures Enable Record High-Harmonic Generation From ...
    Jul 28, 2021 · Cornell researchers have developed nanostructures that enable record-breaking conversion of laser pulses into high-harmonic generation.Missing: ultrafast | Show results with:ultrafast
  21. [21]
    Robust Isolated Attosecond Pulse Generation with Self-Compressed ...
    Mar 18, 2024 · We theoretically demonstrate a novel scheme for the straightforward and compact generation of IAPs from multicycle infrared drivers using hollow capillary ...
  22. [22]
    Critical Phase Matching - RP Photonics
    There is only a finite range of beam angles (the acceptance angle, also called angular phase-matching bandwidth) where critical phase matching works; in the ...What is Critical Phase Matching? · Example: Critical Phase... · Spatial Walk-off
  23. [23]
    Phase Matching - RP Photonics
    Figure 1: Phase mismatch for second-harmonic generation. Due to chromatic dispersion, the wavenumber of the second harmonic is more than twice as large as that ...Missing: uniaxial | Show results with:uniaxial
  24. [24]
  25. [25]
    Spatial Walk-off - RP Photonics
    Spatial walk-off is encountered in nonlinear frequency conversion schemes based on critical phase matching in nonlinear crystals (even with collinear wave ...
  26. [26]
    Properties of BBO Crystal - United Crystals
    Acceptance Angle (mrad-cm), 1.0, 0.5, 0.3, 0.2. Walk-off Angle (degree), 3.2, 4.1, 4.9, 5.5. Type I phase matching BBO is also very efficient for the intra- ...
  27. [27]
    [PDF] bbo - EKSMA Optics
    Phase-matching angle, deg. 15. 0. 30. 45. 60. 75. 90. Type 2 (o-beam). Type 2 (e ... 1.2 mrad × cm (Type 1 SHG 1064 nm). Thermal acceptance. 70 K × cm (Type 1 SHG ...
  28. [28]
    [PDF] Contents - United Crystals
    Non-Critical Phase-Matching (NCPM) of LBO is featured by no walk-off, very ... dn/dT (10-6/°C). 1.06µm dno/dT= 167; dne/dT= 176 dno/dT= 98; dne/dT= 66.
  29. [29]
    Noncritically phase‐matched second‐harmonic generation and ...
    Apr 15, 1991 · We evaluate the potential of lithium triborate (LBO) temperature‐tuned to achieve type‐I noncritical phase matching, for picosecond ...Missing: seminal | Show results with:seminal
  30. [30]
  31. [31]
    Your source for BBO, LBO, BiBO, KTP, KTA, Nd:YAG, Nd:YVO4 ...
    By using extracavity KTP SHG, a conversion efficiency of over 80% and 700mJ ... The recently developed new application is the non-critical phase-matched (NCPM) ...
  32. [32]
    LBO Properties - United Crystals
    Non-Critical Phase-Matching Applications:​​ It helps LBO work in its optimal conditions. SHG conversion efficiencies of more than 70% for pulsed and 30% for cw ...
  33. [33]
    Nonlinear Crystals for Harmonics Generation - Optogama
    Apr 8, 2025 · ... 532nm) of these same lasers under non-critical phase matching conditions (NPCM, T= 50 °C).The main drawback is that these crystals are ...
  34. [34]
    A survey of research on KTP and its analogue crystals - ScienceDirect
    Based on the calculation of the refractive indices and phase matching of KTP crystals, it is shown that non-critical phase matching can be achieved ...
  35. [35]
  36. [36]
    [PDF] Second Harmonic Generation, Fall 2016 | OPTI 511L
    In Type II phase matching the fundamental is polarized at 45° to the ′ x , y axes and the second harmonic is polarized along y; an ordinary and an extraordinary ...
  37. [37]
    [PDF] Collinear phase matching for second harmonic generation using ...
    In this article, we propose a technique to address this problem of finding PM directions for biaxial crystals by us- ing simple vector analysis to project PM ...Missing: critical | Show results with:critical
  38. [38]
    [PDF] L8_ Phase matching.pptx - CREOL #ucf
    The rediced coupling equations (5.4) of Lecture 5 ignore the position dependence of the electrical fields, specifically their phases.
  39. [39]
    [PDF] Nonlinear Optics (WiSe 2018/19)
    Nov 2, 2018 · tanθ . Therefore, we obtain for the walk-off angle between Poynting vector and wave ... The walk-off angle ρ at phase matching is tanρ = (n.
  40. [40]
    High-efficiency, multicrystal, single-pass, continuous-wave second ...
    We describe the critical design parameters and present detailed experimental and theoretical studies for efficient, continuous-wave (cw), single-pass second ...
  41. [41]
    Thermal dephasing in an enhanced cavity based high-power ...
    We report on the performance of an enhanced-cavity (EC) designed for obtaining high-power and efficiency second-harmonic generation (SHG).
  42. [42]
    LBO Crystals for Second and Third Harmonic Generation - Thorlabs
    Apr 24, 2025 · SHG is a nonlinear process through which two photons with the same wavelength, called the fundamental, are converted into a single SHG photon ...<|control11|><|separator|>
  43. [43]
    Characterization of semiconductor interfaces by second-harmonic ...
    At a surface or interface, the bulk symmetry is broken and electric dipole effects are allowed. Cross sections for SHG and SFG processes are small, giving ...
  44. [44]
    [PDF] Second Harmonic Generation of Molecules Located at the Air ...
    Therefore it is possible to distinguish surface and bulk species even when they are the same molecules. ... The signal levels are 10-100 times larger than that ...
  45. [45]
    Multi‐Scale Modeling of Surface Second‐Harmonic Generation in ...
    May 6, 2024 · However, a signal is observed from interfaces where the symmetry is broken. Whereas the effect can be phenomenologically accommodated, a ...Missing: mechanism | Show results with:mechanism
  46. [46]
    Second-Harmonic Enhancement with Mie Resonances in Perovskite ...
    The SHG signal is enhanced at the linear Mie resonances of single nanoparticles. Therefore, it is important to understand which factors influence the ...Results and Discussion · Methods · Supporting Information · References
  47. [47]
    Surface and interface resonances in second harmonic generation ...
    Mar 29, 2006 · The separation of surface and interface contributions to optical second harmonic generation from thin Ag and Au films on ...
  48. [48]
    Trends Surface studies with optical second-harmonic generation
    This introductory review covers the application of the non-linear optical processes of second-harmonic and sum-frequency generation to the study of surfaces ...<|control11|><|separator|>
  49. [49]
    An Optical Frequency Domain Angle Measurement Method Based ...
    Jan 19, 2021 · The sinc function is determined by the length of the nonlinear optical crystal and the phase mismatch (Δk) as a function of incident angle to ...
  50. [50]
    Phase-Matching Controlled Orbital Angular Momentum Conversion ...
    Sep 29, 2020 · Here, we demonstrate experimentally that OAM transfer depends strongly on the phase-matching condition defined by both linear and orbital angular momenta.
  51. [51]
    Characteristics of optical second-harmonic generation due to ...
    Characteristics of optical second-harmonic generation due to Čerenkov-radiation-type phase matching ... A far-field pattern from a rectangular channel ...Missing: sinc | Show results with:sinc
  52. [52]
    Ultrafast optical vortex beam generation in the ultraviolet
    The asymmetry in the power of the two lobes can be attributed to spatial walk- off in the crystal. As is evident from Fig. 5, the angular band- width of the ...<|separator|>
  53. [53]
    Frequency-conversion of vector vortex beams with space-variant ...
    Jul 30, 2019 · The asymmetric intensity distribution of the SH beam especially at higher vortex orders can be attributed to the spatial walk-off effects of ...
  54. [54]
    (PDF) Walk-off-compensated type-I and type-II SHG using twin ...
    Aug 9, 2025 · The walkoff-compensated (WOC) twin-crystal device enables to reduce the aperture effects in critically phase-matched parametric generation.
  55. [55]
    Planar second-harmonic generation with noncollinear pumps in ...
    We show that the second-harmonic radiation is emitted in the form of two cones as well as in a plane representing the cross-correlation of the two fundamental ...
  56. [56]
    Large acceptance of non-collinear phase-matching second ...
    Nov 6, 2014 · We investigate several bandwidths of non-collinear phase-matching second harmonic generation, which is generated by sum-frequency of the ...
  57. [57]
    (PDF) Cone-angle tunable second-harmonic generation at the edge ...
    A simple method to generate second-harmonic conical waves with tunable cone angles (apex angles) only using a cube uniaxial nonlinear crystal is proposed.
  58. [58]
  59. [59]
    Controlling second-harmonic diffraction by nano-patterning MoS 2 ...
    Nov 19, 2019 · With the help of an additional lens in the SH beam path, we can also measure the angular spectrum of the SH, i.e. the angular distribution of ...
  60. [60]
    Far-field polarization signatures of surface optical nonlinearity in ...
    Jun 29, 2020 · ... far-field pattern with lobes ... Tailorable second-harmonic generation from an individual nanowire using spatially phase-shaped beams.Missing: sinc | Show results with:sinc
  61. [61]
    Angular dependence of phase-matched second-harmonic ...
    Recently, it was shown that phase-matched second-harmonic generation (SHG) in the dipole approximation is possible in centrosymmetric photonic crystals because ...Missing: sinc | Show results with:sinc
  62. [62]
    Spatial hole burning, nonlinear phase, and SHG acceptance ...
    Spatial hole burning, nonlinear phase, and SHG acceptance bandwidth effects in ... Second harmonic generation · PDF Article. Abstract. Poster Presentation ...
  63. [63]
    From coherent instabilities to spatial hole burning | Phys. Rev. A
    May 7, 2008 · ... spatial hole burning (SHB) and the other one is related to the ... second-harmonic generation (SHG) component to be detected. One can ...
  64. [64]
    Passive FM laser operation and the stability of intracavity-doubled ...
    A time-domain description of second-harmonic generation and steady-state spatial hole burning is used to model the performance of multimode intracavity ...
  65. [65]
    Heat coupled Gaussian continuous-wave double-pass type-II ...
    ... thermal lensing. Mohammad Sabaeian, Fatemeh Sedaghat Jalil-Abadi, Mostafa ... A model describing the thermal effects in type II second harmonic generation ...
  66. [66]
    Thermal Effects in High-Power Continuous-Wave Single-Pass ...
    Dec 18, 2013 · Thermal Effects in High-Power Continuous-Wave Single-Pass Second Harmonic Generation ... crystal to provide green output at 532 nm.
  67. [67]
    Heat coupled Gaussian continuous-wave double-pass type-II ...
    A model describing the thermal effects in type II second harmonic generation (SHG) of Gaussian continuous-wave (CW) in a double-pass cavity is presented. ... thermal lensing. Opt Express. 2014 Oct 20;22(21):25615-28. doi ...
  68. [68]
    Theoretical and experimental researches on the walk-off ...
    Dec 14, 2021 · Accordingly, the beam quality factors in the X and Y directions improved to M x 2 = 1.2 and M y 2 = 1.4 under the same pump power, respectively.
  69. [69]
    [PDF] Second harmonic generation with focused beams in a pair of walkoff ...
    We have presented mathematical expressions based on the heuristic method pioneered by Boyd and. Kleinman [6] for second harmonic generation in two walkoff- ...
  70. [70]
    Quality enhancement of a 355 nm sum-frequency light generated ...
    Oct 18, 2019 · This work provides a comprehensive study around the effect of a quality-deteriorated fundamental beam on the quality of a 355 nm wavelength beam experimentally ...
  71. [71]
    Spatial–temporal evolution of ultrashort laser pulse second ...
    Jun 15, 2021 · Such a phase speed mismatch and group velocity mismatch will cause the presence and action of many negative factors, such as the wide frequency ...
  72. [72]
    High-efficiency generation of ultrashort second-harmonic pulses ...
    With this group-velocity mismatch, the limiting crystal length Lmax for avoiding pulse broadening is 4.79 and 2.81 cm/ps for 5% period modulation fluctuation in ...
  73. [73]
    Second-harmonic generation of ultrashort pulses in refractive index ...
    The broadband extension is possible using crystals with opposite group velocity or temporal walk-off. For short pulses, the temporal walk-off increases the ...
  74. [74]
    Engineered aperiodically poled nonlinear crystal for high-power ...
    We demonstrate the use of an engineered aperiodically poled (APP) crystal grating to generate a high-efficiency second-harmonic radiation of about 66% at 50 ...
  75. [75]
    Proposal of high quality walk-off compensated sum frequency ...
    As key materials in the optical communication and laser industry, birefringent crystals are traditionally utilized to fabricate polarizers, optical isolators, ...
  76. [76]
    Super-tunable, broadband up-conversion of a high-power CW laser ...
    Apr 13, 2017 · The fan-out and temperature-controlled QPM devices have also been used for up-conversion bandwidth broadening by mechanical displacement and ...
  77. [77]
    Newlight Photonics Inc. – Your source for BBO, LBO, BiBO, KTP ...
    at 1064 nm ... A laser pulse as short as 10 fs can be efficiently frequency doubled with a thin BBO crystal in terms of both phase-velocity and group-velocity ...
  78. [78]
    BBO Crystal & its Applications - OptoCity
    Oct 20, 2025 · - BBO is one of the best NLO crystal for frequency doubling, tripling and quadrupling of high power acousto-optic, electro-optic Q-switched and ...
  79. [79]
    [PDF] Lithium Triborate Crystal (LiB - ,LBO) - Laser Components
    LBO is featured by: ▫ Broad transparency range from 160 nm to 2600 nm (see Figure 1);. ▫ High optical homogeneity and being free of inclusi-.
  80. [80]
    Lbo Crystal - Crylight
    1 Broad transparency range from 160nm to 2600nm (SHG range from 550nm to 2600nm). · 2 Type I and type II non-critical phase-matching (NCPM) over a wide ...
  81. [81]
    Properties of KDP, DKDP and ADP crystal - United Crystals
    KDP, DKDP, and ADP crystals are used in laser systems. DKDP has a 200-2150nm range, KDP 180-1550nm, and ADP 180-1500nm. DKDP has a 0.40 nonlinear coefficient,  ...
  82. [82]
    KDP/DKDP Crystals & Applications - OptoCity
    Oct 16, 2025 · KDP and KD*P are hygroscopic, a sealed housing with transmitting windows is recommended for protection of the crystal surfaces.
  83. [83]
    [PDF] Crystals Selection Guide - EKSMA Optics
    critical phase-matching is highly desirable for fre quency conversion of ... Phase matching angle for type 1 sHG at 10.6 µm, deg. 76. 55. 67. 14. Walk-off ...
  84. [84]
    Nonlinear optical properties of urea - ScienceDirect
    We report measurements on the second and third order nonlinear hyperpolarizabilities of the urea molecule using the techniques of dc field induced ...
  85. [85]
  86. [86]
    Mechanism of linear and nonlinear optical properties of the urea ...
    May 26, 2011 · The origin of the linear and nonlinear optical (NLO) properties of the urea crystal family has been analyzed by coupling the calculated electronic structure ...
  87. [87]
    Quasi-phase-matched second-harmonic generation by use of a total ...
    Second-harmonic generation experiments demonstrated angle-tuned output at 4.6 to µ 5.3 µ m in GaAs and 1.7 to µ 2.0 µ m in ZnSe crystals when pulsed infrared ...Missing: mid- | Show results with:mid-
  88. [88]
    Broadband second-harmonic generation in a double-tapered ...
    This article analytically describes broadband second-harmonic generation in a double-tapered gallium arsenide (GaAs) slab using total internal reflection ...
  89. [89]
    Mid-infrared nonlinear phenomena in polycrystalline semiconductors
    We have studied second-harmonic generation by 30 ps and 70 ns 10 μm CO2 laser pulses in nonlinear polycrystalline ZnSe, CdTe and GaAs samples.
  90. [90]
    Comprehensive Guide to Growth Method an of BBO crystals - CryLink
    Jun 21, 2023 · The flux method, also known as the solution growth method, is another common technique for BBO crystal growth. It involves dissolving the raw ...
  91. [91]
  92. [92]
    Temperature-insensitive second-harmonic generation based on ...
    Here we introduce a temperature-insensitive noncollinear (TIN) phase-matching scheme, which can significantly improve the performance of high-power SHG devices ...<|separator|>
  93. [93]
    [PDF] Improved second-harmonic generation by selective Yb ion doping in ...
    KDP is the only crystal capable of being grown to large sizes [8]. However, it has a small and is hygroscopic. KTP, on the other hand, has a large nonlinear ...
  94. [94]
    Non-linear Optical Crystals, BBO, LBO, DKDP, KDP, ADP, LiNbO3 ...
    KTP is the most common NLO crystal for SHG or YAG lasers. The high conversion efficiency, non-hygroscopicity, and cost-efficiency make KTP the first choice for ...
  95. [95]
    3 W, 300 μJ, 25 ns pulsed 473 nm blue laser based on actively Q ...
    A laser output of 8 W, 10 kHz, and 30 ns at 946 nm is reported. The laser is extracavity frequency doubled in a BiBO crystal to obtain 3 W and 300 μJ of blue ...
  96. [96]
    103W high beam quality green laser with an extra- cavity second ...
    In this paper, we reported a high efficiency and high beam quality green laser with an extracavity second harmonic generation (SHG). A high power Q-switched ...
  97. [97]
    High Efficiency and High Stability for SHG in an Nd:YVO4 Laser with ...
    Power of the laser with KTP intra-cavity. The intracavity conversion efficiency was found by relating the optical power at 532 nm to the optical power at 1064 ...
  98. [98]
    Efficient and compact intracavity-frequency-doubled YVO 4 /Nd:YVO ...
    With an incident pump power of 27.5 W, as high as 7.2 W of CW output power at 532 nm was achieved. The optical-to-optical conversion efficiency of 26.2% was ...
  99. [99]
    103W high beam quality green laser with an extra- cavity second ...
    ... green problem.” [14]. B. Yong et al. reported a green laser with the beam ... thermal induced phase mismatching for SHG. In order to compensate this ...
  100. [100]
    Review of cw, high-power, diode-pumped, green lasers
    ... laser light, similar to that produced by argon ion lasers. Here a solution to the green problem is mandatory. The green problem solutions can be categorize ...<|separator|>
  101. [101]
    Green Lasers – pointers, frequency-doubled, Nd:YAG laser
    Figure 2: A high-power thin disk laser with intracavity frequency doubling generates 1.3 kW of green light with excellent efficiency. Source: Max Kovalenko, ...Missing: QPM | Show results with:QPM
  102. [102]
    130-W picosecond green laser based on a frequency-doubled ...
    A second-harmonic conversion efficiency of 54% is achieved with a 15-mm-long noncritically phase-matched lithium triborate (LBO) crystal from a 240-W 8-ps 78- ...
  103. [103]
  104. [104]
  105. [105]
  106. [106]
  107. [107]
  108. [108]
  109. [109]
    Video-rate two-photon microendoscopy using second harmonic ...
    Oct 11, 2024 · We present a 2.5-mm-diameter resonant fiber scanning two-photon microendoscope with a 30-mm long forward-viewing rigid probe tip that enables video-rate ...
  110. [110]
    Nonlinear Optical Methods for Characterization of Molecular ... - NIH
    Second harmonic generation (SHG) is a parametric process that has all the virtues of the two-photon version of MPEF, yielding a signal at twice the frequency ...
  111. [111]
    On the determination of χ(2) in thin films: a comparison of one-beam ...
    Mar 20, 2017 · The technique is based on Maker's finding that the pattern of the transmitted second-harmonic (SH) power generated in a bulk nonlinear crystal ...
  112. [112]
    Optical second harmonic generation as a probe of surface chemistry
    Second-Harmonic Generation Provides Insight into the Screening Response of the Liquid Water Interface. The Journal of Physical Chemistry C 2023, 127 (30) , ...
  113. [113]
    Nonlinear Optical Characterization of 2D Materials - PMC
    The second-order nonlinearity (three wave mixing) is determined by χ ( 2 ) of the second term in Equation (1), including SHG, sum and difference frequency ...
  114. [114]
    [PDF] Characterization of domain distributions by second harmonic ...
    Jun 13, 2018 · Here, an optimized second harmonic generation method has been explored for ferroelectric domain characterization. Combing a unique ...
  115. [115]
    Characterization of thymine microcrystals by CARS and SHG ...
    Oct 13, 2020 · Second harmonic (SHG) and coherent anti-Stokes Raman scattering (CARS) spectroscopy can be used to reveal arrangement of thymine molecules, one of the DNA ...
  116. [116]
    Second Harmonic Generation characterization of SOI wafers: Impact ...
    In this work, we investigate Second Harmonic Generation (SHG) as a non-destructive characterization method for Silicon-On-Insulator (SOI) materials.
  117. [117]
    Optical second-harmonic generation as a surface probe for ...
    We show that optical second-harmonic generation can also be used as an effective surface probe with a submonolayer sensitivity for media without inversion ...Missing: review | Show results with:review
  118. [118]
    High‐Throughput Scanning Second‐Harmonic‐Generation ...
    Mar 14, 2023 · In this study, we developed a high-throughput scanning SHG (HTS-SHG) microscopy method with automatic partitioning, accurate positioning, and ...
  119. [119]
    Quantitative measurement and interpretation of optical second ...
    In this review, we try to assess recent developments in the SHG experimental methodologies towards quantitative analysis of the nonlinear optical properties.<|separator|>