Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Absolute instrument spectral response measurements using angle-resolved parametric fluorescence

Open Access Open Access

Abstract

The broadband parametric fluorescence from a nonlinear crystal can be used as a compact primary source instead of a blackbody for absolute measurements of instrument spectral efficiency. We describe such a setup for measuring the instrument spectral response function in the wavelength range from 450 to 1000 nm. We perform angle–resolved imaging spectroscopy of conical parametric fluorescence in a beta-barium borate crystal pumped by a 405-nm diode laser. The experimental angle–resolved spectra and the generation efficiency of parametric down–conversion agree with a plane-wave theoretical analysis.

© 2013 Optical Society of America

1. Introduction

Radiometry is important to a wide range of applications requiring accurate or absolute measurements of radiation spectral responsivity. These include lighting, optical sources, detectors, optical components, spectrometers, and fluorescence microscopy. Blackbody sources and electron storage rings (synchrotron radiation) are accepted primary radiation source standards in radiometry because their spectral characteristics can be calculated from fundamental physical processes and parameters [14]. Synchrotron radiation sources are limited to large national laboratories such as BESSY [5]; the radiometric standard based on high-temperature blackbodies and cryogenic radiometers at NIST [3, 6] is not available for in situ calibration in users’ laboratories. Spontaneous parametric down-conversion (SPDC), in which a single photon is down-converted into a correlated photon pair is an alternative primary source for use in radiometry [2, 79].

SPDC, also known as parametric fluorescence or parametric scattering, is a second-order optical process in which a driving pump photon is scattered into signal–idler photon pairs subject to energy and momentum conservation. This spontaneous parametric emission can be described properly only by field quantization [1015]. It offers a convenient mechanism of measuring detector efficiency in the photon counting regime where blackbodies are not suitable. Measurements made with such a SPDC sources are intrinsically absolute, not relying on an externally calibrated radiometric standard. By replacing the specimen/samples with a SPDC source at the focal plane of a microscope spectroscopy system, one can correct the spectral response including transmittance and geometric losses due to all optical components used in the set-up. One can further identify or correct transmission losses due to reflection and absorption, and geometric losses due to limiting apertures and detector area in situ. However, the absolute spectral distribution of a SPDC source has not been characterized as comprehensively as in the case of blackbodies and electron storage rings.

Here we experimentally measure angular spectral distributions of parametric fluorescence and describe a compact set-up to obtain the absolute spectral efficiency of an imaging spectrometer equipped with a CCD camera in the visible wavelength range (450 nm to 1000 nm). In a focal area of 100 μm2 in a BBO (beta-barium borate) crystal pumped by a 100 mW 405-nm pump laser, the parametric fluorescence spectral flux density at 810 nm is comparable to that from a T = 1500 K ideal blackbody source. The conical parametric fluorescence is directional and can be focused or relayed through a complex microscopic and spectroscopic system. The compact set-up is portable and can in principle be calibrated against a NIST standard radiometer. The wavelength limit is largely imposed by the responsivity of the silicon-based CCD camera. Our method can also be extended to the near infrared wavelength range (900 nm to 1700 nm) for a spectroscopy or microscopy system equipped with an InGaAs array.

Although a detailed analysis of the angle-resolved parametric fluorescence spectrum is presented here, calibration of the instrument spectral response function can be performed without a full angle-resolved spectrum as long as the scattered pump laser light is rejected properly. Using the Fourier transform imaging method and simple models described here, one can select suitable lenses and fibers, and BBO crystals with matching dimensions or numerical apertures to optimize the overall system efficiency. The method here can be also used to monitor angular distributions of SPDC and optimize the photon pair generation rates and coupling efficiency into single-channel avalanche photodiodes or photon counters via fibers.

The use of SPDC as a primary source in radiometry relies on the fact that the fundamental physical processes of SPDC are well known [4]. Since its prediction and observation in the 1960s, parametric fluorescence has become a technique for measuring second-order nonlinear susceptibilities [1518] and for developing tunable light sources via parametric oscillation or amplification processes. In 1969, Zeldovich and Klyshko first proposed the use of parametric fluorescence (luminescence) as a nonclassical source of photon pairs in [19]. This description was experimentally verified by Burnham et al. in [20]. Signal–idler photon pairs generated by SPDC have been used to address fundamental issues of quantum theory and have found applications in quantum entanglement and quantum information processing [2124].

The possible wave vectors of the signal–idler photon pairs are determined by energy and momentum conservation, a constraint referred to as phase-matching, leading to highly directional parametric emission. The phase-matching condition can be met by selecting birefringent crystals with appropriate refractive indices or by designing waveguides or periodic structures of specific wavelengths. There are two major types of phase-matching schemes for parametric down-conversions: Type-I, where signal–idler photons have the same polarization (co-linearly polarized photons), and Type-II, where the signal–idler photons have orthogonal polarization (cross-linearly polarized photons). Both types of parametric processes have been used to generate photon pairs, sometimes referred to as biphoton states, which exhibit correlation/entanglement for variables including polarization, momentum, time, energy, and angular momentum. When the phase-matching condition is met, the signal and idler radiation form a conical pattern independent of the intensity of the pump source. The angular distribution of parametric fluorescence is determined by the energy of the pump, signal, and idler waves, subject to the dispersion of the crystal and walk-off angles of these three waves. The magnitude of the second-order nonlinear susceptibility χ(2) is a typical selection criterion for parametric downconversion. Many uniaxial or biaxial nonlinear crystals have been used for parametric down conversion: for example, KD*P (potassium dideuterium phosphate, KD2PO4) [25, 26], BBO (beta-barium borate, β-BaB2O4) [27, 28], and LBO (lithium niobate, LiNbO3) [13]. Among these, LBO and BBO crystals have been studied for harmonic frequency generation, optical parametric oscillation, and generation of bi-photon states. Theoretical modeling and numerical calculation of phase-matching in these nonlinear crystals are well established [2931].

In this article, we first present experimental angle-resolved images and spectra of parametric fluorescence (SPDC) from a BBO crystal which is widely used for the generation of entangled biphoton states [2124, 32]. The experimental angle-resolved spectra and the generation efficiency of parametric down conversion agree with a plane-wave theoretical analysis. Finally, we describe procedures to use parametric fluorescence as a broadband light source for the calibration of the instrument spectral response function in the wavelength range from 450 to 1000 nm.

2. Experimental methods

We measure the angular distribution and photon flux of parametric fluorescence from a 3-mm thick BBO crystal. The BBO crystal is cut at an angle of θm = 29 ± 0.5° with respect to the optical axis as illustrated in Fig. 1. This cut angle θm is chosen for the Type-I (eo + o) degenerate parametric down-conversion at λ = 810 nm with a pump λp = 405 nm. The crystal is mounted on a three-axis rotary mount with the crystal’s optical axis (OA) in the horizontal plane when the parametric fluorescence signal is maximized. The angle formed by the OA and the pump’s propagation wave vector can be finely tuned by tilting the crystal to satisfy the phase-matching condition for various θm near the crystal cut angle. We can thus adjust the pump and signal Poynting vectors from collinear to non-collinear and generate parametric fluorescence with varying conical emission angles.

 figure: Fig. 1

Fig. 1 Schematic diagram of the crystal and laboratory frame coordinates for parametric down-conversion in a BBO crystal. The phase-matching angle θm is defined as the angle formed by the crystal optical axis (z′) and the pump wave vector (z). The angles θ′s and θs are, respectively, internal and external angles formed by the signal and pump wave vectors. Here the incident pump wave is horizontally polarized, leading to a vertically polarized down-converted signal and idler waves.

Download Full Size | PDF

The pump is a violet diode laser with a TEM00 linearly polarized 2-mm 1/e2 diameter output beam at a wavelength λp = 405 nm (CNI Laser MLL-III-405). The pump beam is focused on the crystal through a lens (L1) with a focal length of 500 mm. Lens L1 and the objective are positioned to form a telescope such that the residual pump beam is collimated with a reduced beam radius below 100 μm. By passing the pump beam through a pair of a half-wave plate (HWP) and a Glan–Taylor polarizer (P1), we can vary the incident pump intensity by rotating the HWP while maintaining the degree of linear polarization better than 99.9%.

The angle-resolved images and spectra of parametric fluorescence are measured by a Fourier transform optical system, including a 20× long-working-distance objective and an imaging spectrometer as shown in Fig. 2. The BBO crystal is positioned at the focal plane of the objective lens (effective focal length fo = 10 mm). The parametric fluorescence with an amplitude distribution F(x, y) at the crystal is collected by a 20× microscope objective with a 10-mm effective focal length (numerical aperture N.A. = 0.26). The back focal plane of the objective is the Fourier transform plane with coordinates (u, v) = (fo ×sin(θx), fo ×sin(θy)). The collection angle is within ±15°, limited by the objective. The objective lens converges parallel rays emanating from the crystal to the back focal plane of the objective. In this plane, the fluorescence image in the crystal is transformed into a far-field image in spatial frequency that is related to the emission angle as described above. The spatial intensity distribution of parametric fluorescence in the back focal plane of the objective lens thus corresponds to the angular distribution of radiation. This Fourier transform plane is placed at the front focal plane of lens L2 (focal length f = 100 mm). Lenses L2 and L3 are identical and separated by a distance of 2f, and they relay the Fourier transformed images to the entrance plane and then onto the charge-couple device (CCD) through the zero-order diffraction off the grating of the imaging spectrometer (PI-Acton SpectroPro 2750i, focal length 750 mm). In this way, we measure the angular distribution of parametric fluorescence. When lens L2 is removed, we project the real-space spatial intensity distribution of parametric fluorescence from the BBO crystal onto the CCD.

 figure: Fig. 2

Fig. 2 Schematic diagram of the experimental setup. The transmitted and scattered pump photons are rejected by a miniature beam blocker, a thin-film notch filter, and long-pass filters. The angular and spectral distributions of conical parametric fluorescence are measured by an imaging spectrometer through a Fourier transform optical system. L1, L2 and L3 are convergent lenses with focal length f = 500, 100, and 100 mm, respectively; Obj is an objective (20× N.A.=0.26) with an effective focal length of 10 mm (Mitutoyo Plan Apo infinity-corrected long-working-distance objective); P1 and P2 are Glan–Taylor and Glan–Thompson polarizers; M1 and M2 are silver mirrors; HWP is a half-wave plate for λ = 405 nm; BB is a miniature pump beam blocker; NF is a notch filter (Semrock 405-nm StopLine single-notch filter); and LP represents longpass filters (a Semrock 409-nm blocking edge BrightLine long-pass filter and a Schott GG435 glass filter). When L2 is removed, a real-space fluorescence image is formed at the entrance of the spectrometer with an imaging magnification of 10×. Examples of real-space and angle-resolved fluorescence images are shown with actual dimensions.

Download Full Size | PDF

The fluorescence image is recorded by a CCD positioned in a conjugate imaging plane of the Fourier plane. The resultant intensity distribution is related to the Fourier transform of the intensity of the parametric fluorescence I(x, y) = |F(x, y)|2, where F(x, y) is the electromagnetic field distribution at the crystal. By projecting this far-field image through the entrance slit and the first-order diffraction of a 300 lines/mm grating, we obtain angle-resolved spectra as an image by taking the spectral dispersion of the parametric fluorescence as a function of angle. The spectral resolution of 0.1 nm is determined by the dispersion of the grating and the width of the entrance slit (≈ 100μm). The spatial and angular resolutions of the system are approximately 2 μm and 2 mrad, respectively, limited by the pixel size of the liquid-nitrogen-cooled CCD camera.

In a parametric scattering process, only about one out of 1010 incident photons is parametrically down-converted. It is essential to prevent the transmitted and scattered pump photons from entering the spectrometer. For Type-I phase matching in a negative uniaxial BBO crystal (e-o-o case), the polarization of the high-frequency pump (e-wave) is orthogonal to the polarization of the signal and idler (o-waves). Thus, the transmitted and scattered pump photons can be suppressed by approximately 4–5 orders of magnitude by a pair of polarizers (P1 and P2) with orthogonal polarization orientations for pump and signal/idler waves, respectively. The pump photons are further rejected by a thin-film notch filter (Semrock 405-nm StopLine single-notch filter) and two longpass filters (a Semrock 409-nm blocking edge BrightLine long-pass filter and a Schott GG435 glass filter). The filters are arranged in the sequence shown in Fig. 2 to suppress fluorescence from filters induced by the transmitted violet pump laser. The combination of polarizers and filters allows for a rejection of pump photons by approximately a factor of 1010. This rejection ratio of 1010 can be further improved above 1014 by a miniature beam blocker made of a ∼ 0.5 mm-diameter silver-paste dot on a microscope cover positioned in front of the notch filter (NF)/objective. Depending on the signal wavelength, the measured angle-resolved spectra may still contain residual transmitted and scattered pump and background fluorescence from filters. We measure such a background emission spectrum in the absence of Type-I e-o-o parametric fluorescence by rotating the BBO crystal 90 degrees azimuthally.

3. Experimental results and modeling

3.1. Phase matching

The three-wave parametric processes are calculated according to the conservation of energy and momentum, commonly referred to as phase matching. The angle-resolved spectra of the parametric fluorescence are consistent with the tuning curves calculated for the phase-matching condition under a plane-wave approximation.

The energy conservation condition is expressed as

ωp=ωs+ωi,
where ωp is the frequency of the incident pump wave and ωs and ωi are the frequencies of the signal and idler waves.

The momentum conservation condition can be expressed as

kp=ks+ki,
where kp, ks, and ki are the pump, signal, and idler wave vectors, respectively. For Type-I down-conversion in a BBO, the signal and idler labels are arbitrary. In the case of degenerate down conversion, ks = ki, and Eq. (2) reduces to
np=nscos(θs),
where np and ns are the indices of refraction of the pump and signal, and θ′s is the angle formed by the propagation directions of the signal and pump waves inside the crystal.

Down-converted signal/idler photons are co-linearly polarized but orthogonal to the polarization of the pump wave. The wavelengths and wave vectors of the parametric fluorescence are determined by the phase-matching angle θm, the angle formed by the optical axis of the crystal (z′-axis), and the wave vector of the pump wave (z-axis) as shown in Fig. 1. By tilting the crystal, we can vary the phase-matching condition from collinear to non-collinear, leading to a conical angle up to 5 degrees for degenerate parametric fluorescence near 810 nm.

For Type-I phase matching, the incident pump photons are subject to the extraordinary index of refraction ñ(θm), while the down-conversion photons are subject to the ordinary index of refraction. The extraordinary index of refraction, ñ, depends on the phase-matching angle θm and follows the relationship:

n˜(θm,λ)=(cos(θm)2no(λ)2+sin(θm)2ne(λ)2)12.

In the parametric process, a pump wave of wavelength λp creates signal waves at λs, and angles θs. In the Type-I e-o-o case, ns = no(λs) and np = ñ(θm, λp) [Eq. (4)]. Here the labeling of signal and idler waves is arbitrary (θs = θi). A continuum of phase-matching functions Φ(λs, θs) for parametric fluorescence can be obtained using the aforementioned equations and indices of refraction no(λs) and ne(λs). Indices of refraction of wavelengths ranging from 0.3 μm to 5 μm are extracted from ”NIST Noncollinear Phase Matching in Uniaxial and Biaxial Crystals Program” as described in [29]. We calculate the phase matching functions (tuning curves) for down-converted signal/idler waves ranging from 430 to 1000 nm for a pump wave λp = 405 nm.

3.2. Angle-resolved imaging

We adopt the plane-wave analysis developed in [12,30,33] to determine the angular distribution of parametric fluorescence. The effects of a finite pump beam size have also been considered, for example, in [34, 35]. Under a plane-wave approximation, the parametric fluorescence forms a conical angular distribution. The diameter and axis of the conical emission are determined by the wavelengths of the pump, signal, and idler waves, the Poynting vector walk-off angles of the three interacting waves, and the dispersion of the nonlinear crystal. In such a parametric process, the angular spread of the conical emission is determined by the conservation of transverse and longitudinal momenta of interacting waves. The transverse momentum induced by focusing the pump into the crystal contributes to a finite angular spread of the cones. In our experiments, we use a lens with a long focal length (f = 500 mm), resulting in a focal spot with a 1/e2 radius larger than 100 μm. Thus, for the experiments reported here, the effects of walk-off and finite pump beam size are negligible compared with the spectral and angular resolution of the optical system.

We can determine the angular distribution of parametric fluorescence using the following phase-matching function for a finite crystal length L and a pump Gaussian beam profile with a 1/e2 radius W[29, 36, 37]:

Φ=exp(12W2(Δkx2+Δky2))(sin(12LΔkz)12LΔkz)2=exp(12κ2W2)sinc2(12LΔkz).

The mismatch wave vector, Δk, is decomposed into longitudinal ( || kp) and transverse (in x–y plane) parts: Δkz and κ=Δkx2+Δky2. The phase-matching tolerances can be considered in terms of the angular spread (Δθs) and spectral bandwidth (Δλs), defined as the full-width-at-half-maximum (FWHM) for the above function. For a pump wave with a focal radius W ≈ 50μm, only the tolerances from the sinc2 part can be measured in our optical system. For this situation, considering a Taylor series expansion of Δkz near the perfect phase-matching point (Δkz = 0), we can analytically deduce the angular spread (ΔθFWHM) and spectral bandwidth (ΔλFWHM) for the degenerate case (ks = ki):

ΔθFWHM(θs)=2×0.886πL×|Δkz/θs|2.783×nsL×ks×θs,
and
ΔλFWHM(λs)=2×0.886πL×|Δkz/λs|0.443×λs2L×nscos(θs/ns),
where Δkz = |kp − 2ks cos(θs)|, ks = nsωs/c = 2πnss, and θs ≪ 1.

The angle-resolved images of parametric fluorescence at λ = 810 nm are shown in Fig. 3. These false-color images, taken through a 1-nm band-pass filter, represent the angular intensity distributions of parametric fluorescence at λ = 810 ± 0.5 nm for the phase-matching angle θm = 28.6°, 28.8°, 29.1°, and 29.4°. The BBO crystal is cut at the designed angle with about 1° tolerance. To determine the phase-matching angle θm with better precision, we first set the phase-matching angle for the collinear case by comparing the simulated and experimental angular distributions. We then deduce the phase-matching angle for the non-collinear case from the tilting angle of the crystal relative to that for the collinear case. The conical signal angle (θs) of degenerate parametric fluorescence at λ = 810 nm increases with θm. In Fig. 4, we plot the angular spread (ΔθFWMH) as a function of the inverse of the signal angle (1/θs). The angular spread decreases with θs for small angles when sin(θs) ≈ θs. We attribute the discrepancy between experimental data and Eq. (5) to a limited experimental angular resolution (≈ 2 mrad), finite pump beam size and spatial coherence, and the birefrigent walk-off.

 figure: Fig. 3

Fig. 3 Angular intensity distribution of parametric fluorescence at 810 nm. (a)–(d) Angle-resolved images of parametric fluorescence for λ = 810 ± 0.5 nm and θm as indicated. The external angle θs is formed by the pump and signal wave vectors in air [see also Fig. 1]. The color palette represents the intensity of the single radiation. Parametric fluorescence is spectrally filtered through a 1-nm bandpass filter with central wavelength λ = 800 nm. The phase-matching angle is adjusted by tilting the BBO crystal with respect to the pump wave vector. The collinear phase-matching angle is set to θm = 28.6° according to a theoretical calculation using indices of refraction given in [29]. (e)–(g) Experimental (blue dashed line) and theoretical (red solid line) cross-sections. The non-collinear phase-matching angle relative to the collinear one can be determined experimentally by measuring the tilting angle of the crystal surface with respect to the pump wave vector.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Angular spreads ΔθFWHMof parametric fluorescence at 810 nm. ΔθFWHM is plotted as a function of the inverse of the signal angle (1/θs). Experimental data are represented with error bars as solid circles. The red solid line is the theoretical curve according to Eq. 5, while the dashed line is the theoretical curve including a finite angular resolution of 2 mrad. Selected experimental angular intensity profiles for A (θs = 1.3° = 0.023 rad), B (2.8° = 0.049 rad), and C (3.7° = 0.065 rad) are shown in the inset.

Download Full Size | PDF

3.3. Angle-resolved spectroscopy

The parametric fluorescence flux per unit frequency is

Ns(ωs,κs)=deff2ωsωiωpL2Np2π4c3ε0nsninpdωsd2κsd2ξexp(12ξ2)sinc2(12LΔkz),
where deff is the effective second-order nonlinear coefficient, L the interacting crystal length, Np = Pp/ħωp the pump flux, ξ = κW = (κs + κi)W the dimensionless transverse momentum associated with the signal and idler waves, and W the 1/e2 pump beam radius. The integration over ξ is detailed in Sec. 3.5.

Experimental angle-resolved spectra Ns*(λs,θs) and corresponding calculated Ns(λs, θs) are shown in Figs. 5 (a)–(d). The tuning curves, corresponding to the perfect phase-matching condition, are shown as white dashed lines on the experimental spectra. The experimental parameters are Np = 1.63 × 1017/s (Pp = 80 mW), L = 3 mm, and W2 = 60 μm ×30 μm. The indices of refraction ns, ni, and np are evaluated using the database in [29] (ns,i ≈ 1.66 at 810 nm). The effective nonlinear coefficient of a BBO crystal can be deduced from its d-matrix using deff = d31 sin(θm + ρ) − d22 cos(θm + ρ) sin(3ϕ) for Type-I phase matching. θm is the phase-matching angle, ϕ the azimuthal angle, and ρ the birefringent walk-off angle. We use deff =1.75 pm/V for λp = 405 nm adopting from [18, 24, 38].

 figure: Fig. 5

Fig. 5 Angle-resolved spectra of parametric fluorescence. (a)–(d) Experimental angle-resolved parametric fluorescence images (left panel) and spectra (right panel) for a phase-matching angle θm = 28.6°, 28.8°, 29.1°, and 29.4°, respectively. (e)–(h) Theoretical fluorescence flux calculated according to Eq. (6). The tuning curves for the perfect phase-matching condition are indicated by the white dashed lines on experimental imaging spectra. The color palette represents the calculated photon flux with Δλs = 1 nm and Δθs = 0.5 mrad on a logarithmic scale.

Download Full Size | PDF

Computationally, the angular distributions of parametric fluorescence are calculated for signal wavelengths from 430 to 1000 nm with step size δλ = 1 nm and δθs = 0.03° ≈ 0.5 mrad. Selected calculated angular distributions of fluorescence flux per 1 nm are shown in Figs. 5(e)–5(h) for θm = 28.6°, 28.8°, 29.1°, 29.4°. The shortest signal wavelength that appears in the simulation is ≈ 440 nm, limited by the availability of refractive indices between λ = 0.3 μm and 5 μm [31]. Experimentally, parametric fluorescence with a wavelength as short as 431 nm can be observed near θs = 0.

The experimental angle-resolved parametric fluorescence images are shown in the left panel of Fig. 5. These images are taken through a notch filter and longpass filters (cutoff wavelength of 420 nm) as shown in Fig. 2. Experimental angle-resolved spectra are acquired by spectrally resolving the parametric fluorescence across the fluorescence cone center through the entrance slit with an opening of 100 μm, corresponding to Δθx ≈ 0.6°. Angle-resolved fluorescence images [Figs. 5(a)–5(d), left panel] are taken for the zero-order diffraction of a holographic grating with 1800 lines/mm, while angle-resolved spectra (right panel) are dispersed by a grating with 300 lines/mm and a blaze wavelength of 1000 nm. The angular resolution is about 2 mrad, while the spectral resolution is about 0.1 nm. Residual stray or scattered pump laser light can be measured by rotating the BBO crystal 90° azimuthally. Such background ’noise’ is subtracted. Note that the measured CCD intensity is subject to the nonuniform spectral sensitivity and collection efficiency of the optical system including the effects of lens coating, optical filters, and gratings, and the spectral response of the CCD camera. The spectra branches out from θs = 0 at λ ≈ 433 nm. The secondary weak arcs in the wavelength range above 870 nm outside of the expected parametric fluorescence peaks are due to the second-order diffraction of the grating for the fluorescence from approximately 435 nm to 500 nm. The inner arcs branching from 435 nm and closing near 530 nm are due to parametric fluorescence from a Type-II parametric process (eo + e). The turning curves for such Type-II parametric fluorescence are indicated by black dashed-doted lines in experimental spectral images.

3.4. Fluorescence photon flux

The integrated parametric fluorescence photon flux can be obtained by the integrating over ξ and κs in Eq. (6) [30]. For values of θm or ωs such that the parametric fluorescence cone has a radius sufficiently large with negligible emission at the cone center (i.e. non-collinear casese), the resultant integrated fluorescence flux is

Ns=deff2Lωs2ωi2πc4ε0np2Npdωs=(2π)42cdeff2Lε0np2λs4λi2Npdλs.
The efficiency ηsNs/Np is a coefficient depending largely on the material properties such as the second-order nonlinear coefficient, interacting crystal length, and index of refraction for the pump wave. Assuming a bandwidth Δλ = 1 nm and L = 3 mm, the efficiency coefficient ηs = 1.3 × 10−10 for the degenerate parametric fluorescence at λs = 810 nm under λp = 405 nm. Specifically, we evaluate ηs for θm = 29.12° [Fig. 5(g)]. We integrate the photon flux of simulated angle-resolved spectra for λs = 809.5 → 810.5 nm, θs = 0 → 7.5°, and ϕ = 0 → 2π. The total parametric fluorescence photon flux, including both degenerate signal and idler waves, is 2Ns ≈ 4.2 × 107/s.

The wavelength of parametric fluorescence generated here ranges from ≈ 430 nm to above 1000 nm. The collection efficiency and spectral response of the optical systems could vary more than an order of magnitude in such a broad wavelength range. Therefore we consider the degenerate parametric fluorescence at λ = 810 nm to compare the experimentally determined fluorescence photon flux with the theoretically integrated photon flux [Eq. (7)].

The integrated degenerate parametric fluorescence flux at λs = 810 nm as a function of the incident pump flux are shown in Fig. 6. First, we determine the overall collection efficiency and the spectral response of the imaging and spectroscopy system by passing a laser beam of λ = 810 nm with a known photon flux through the same optical path as the parametric fluorescence. We then integrate the signal in angle-resolved images as shown in Fig. 3. Taking into account the system response at λ = 810 nm, we can determine the absolute fluorescence flux experimentally within roughly 20% error. The relative fluorescence fluxes, determined by the linearity of the CCD camera, is within 1% error. The absolute value of the pump flux is known to within roughly 10% error. The relative pump fluxes, as varied by a combination of a half-wave plate and a polarizer and limited by the linearity of the power meter, are known within a few percent. Thus we can investigate the pump flux (power) dependence of parametric fluorescence flux with precision. Fluorescence signal flux is linearly proportional to the pump flux over two order of magnitude, confirming that the dominant signal is spontaneous parametric fluorescence as described by Eq. (7). The slopes correspond to the efficiency coefficients. Considering both signal and idler fluorescence, we determined η = 2ηs to be approximately 2.8 × 10−10. Experimental and theoretical values of η = 2ηs for selected phase-matching angles are listed in Table 1.

 figure: Fig. 6

Fig. 6 Parametric fluorescence flux at 810nm. Integrated degenerate parametric fluorescence flux at λs = 810 nm as a function of the incident pump flux for the collinear case and three phase-matching angles θm of the angle-resolved images shown in Fig. 3. Signal flux is linearly proportional to the pump flux over two order of magnitude, confirming that the dominant signal is spontaneous parametric fluorescence as described by Eq. (7). The slopes η = Ns/Np are 1.2 × 10−10 for θm = 28.6°, 2.6 × 10−10 for θm = 28.8°, and 2.8 × 10−10 for θm = 29.1° and 29.4°.

Download Full Size | PDF

Tables Icon

Table 1. Parametric fluorescence efficiency η = 2Ns/Np for 810 ± 0.5 nm

3.5. Calculation of fluorescence flux

In this section we describe the integration over ξ of Eq. (6) for the calculation of the angular distribution of fluorescence flux here. The mismatch wave vector Δk = k′s + k′ik′p can be decomposed into longitudinal (Δkz) and transverse (κ) parts:

Δkz=ksz+kizkpz=ks2κs2+ki2κi2kpz,andκ=κs+κi.
Here k′s = nsωs/c, k′i = niωi/c, and k′p = npωp/c are the wave numbers for the signal, idler, and pump, respectively. κs (κi) is the transverse wave vector of the signal (idler) wave. The phase-matching function Φ in Eq. (5) is a function of three variables: ωs, κs and κi. Considering energy conservation ωp = ωs + ωi, we can carry out the integration over ξ for two independent variables, ωs and κs. Using ξ = ξs + ξi = (κs + κi)W, we rewrite Eq. (6) as
Ns(ωs,κs)=deff2ωsωiωpL2Np8π4c3ε0nsninpdωsd2κsd2ξiexp(12|ξs+ξi|2)sinc2(12LΔkz)
We further simplify the numerical integration by (a) considering that the sinc2 term is a constant for a given set of ωs and κs, and (b) applying a saddle-point approximation. κs and κi form an angle φ in the xy plane of the crystal, where φ = 0 in the anti-parallel case [corresponding to ϕsϕi = π in Fig. 1]. Under a pump wave vector along , the phase-matching condition is met mostly for κ′s + κ′i = 0; i.e., φ ≈ 0. The Gaussian term of the integrand can thus be separated and integrated with a saddle-point approximation for φ ≈ 0 and ξsξi:
d2ξiexp(ξs2+ξi22ξsξicos(φ)2)ξidξiexp((ξsξi)22)dφexp(ξsξiφ22)2πdξiexp((ξsξi)22).
The integrand is a maximum at ξs = ξi, where pairs of signal and idler photons are emitted in approximately opposite conical directions. Applying the above two approximations, we obtain
Ns(ωs,κs,φ)=deff2ωsωiωpL2Np8π4c3ε0nsninp2πdωsκsdκsdϕdξie12(ξsξi)2sinc2(12LΔkz),
where d2κs = κss, ϕ is the azimuthal angle.

Ns(ωs, κs, ϕ) is isotropic in ϕ. It can be expressed as Ns(λs, θs), a function of experimentally measurable signal angle θs and wavelength λs by considering that κs = ks sinθs and ωs = 2πc/λs:

Ns(λs,θs)=2π2πdeff2ωpL2Npε0nsninpλs5λisin(2θs)dλsdθsdξie12(ξsξi)2sinc2(LΔkz2).

The equation for Ns(λs, θs) above, together with ωp = ωs +ωi and kp = ks +ki, is used in the numerical integration to obtain the theoretical angular distribution of parametric fluorescence shown in Fig. 5.

4. Instrument spectral response function

Parametric fluorescence spectra can also be used to calibrate the instrument spectral response function (ISRF) of the imaging spectroscopy system. According to Eq. (7), which is valid for non-collinear cases, we can deduce a parameter SNs×λs4λi2=2(2π)4cdeff2LNpdλs/(ε0np2)[30]. S is a wavelength-independent constant for a given pump wavelength and geometry. We define a generalized spectral function, S(λs)Ns(λs)λs4λi2, for both calculated and experimental angle-resolved spectra. The calculated Ssim(λs) exhibit less than 1% variation between λ = 460 and 1000 nm. Experimentally, Sexp*(λs)=Ns*(λ)λs4λi2 can be determined from the integration of an angle-resolved spectrum Ns*(λ,θs) over θs. Sexp(λ) represents the relative ISRF of the imaging spectroscopy system, including the optical components, grating, and CCD camera along the fluorescence collection optical path. The value of the ISRF at a fixed wavelength can then be used to determine the absolute ISRF across the parametric fluorescence wavelength range. The effective efficiency, defined as (# of photo-generated electrons / # of photons), is approximately 20% at λ = 810 nm in our experiments. Sexp*(λ) and Ssim(λ) for phase-matching angles θm = 29.1° and 29.4° are shown in Fig. 7. The stray scattered pump laser signal becomes increasingly difficult to subtract from these angle-resolved spectra, leading to a distorted Sexp*(λ) due to overlapping of parametric fluorescence and scattered pump laser near θ = 0. We thus use Sexp*(λ) at θm = 29.1° to deduce the ISRF of our optical setup shown in Fig. 2.

 figure: Fig. 7

Fig. 7 Instrument spectral response function. The normalized spectral density function S(λ) for θm = 29.1° and 29.4°. Ssim is determined from the integration over θs = −7.5° to 7.5° of the calculated angle-resolved spectra as shown in Fig. 5 [see also Eq. (6)]. The black and red curves are Sexp* for θm = 29.4° and 29.1°, respectively. The value of Ssim is a constant with 1% standard deviation across wavelengths from 500 to 1000 nm, validating the calculated angular fluorescence spectra and Eq. (7). Sexp*(λ)=Ns*(λ)λs4λi2, where Ns*(λ) is the integration over θs ≈ −15° to 15° of the experimental imaging spectra Ns*(λs,θs). Sexp*(λ) represents a relative instrument spectral response function (ISRF) of the optical spectroscopy system, including optical filters and a Glan-Thompson polarizer along the path of the fluorescence, a liquid-nitrogen-cooled CCD (PI-Acton Spec-10:400BR), and a 300g/mm plane ruled reflectance grating with 1000-nm blaze wavelength (PI-Acton 750-1-030-1). The polarization of the fluorescence is vertically polarized. The absolute ISRF (or collection efficiency) can be deduced by calibrating Sexp*(λ) at a fixed wavelength such as λ = 810 nm in our experiments.

Download Full Size | PDF

5. Conclusion

The angular and spectral distributions of parametric fluorescence in a BBO crystal pumped by a 405-nm diode laser are measured by employing angle-resolved imaging spectroscopy. The experimental angle-resolved spectra and the generation efficiency of parametric down conversion are compared with a plane-wave theoretical analysis. Parametric fluorescence in a BBO crystal pump by a 405-nm diode laser is used as a broadband light source for the calibration of the instrument spectral response function in the wavelength range from 450 to 1000 nm, limited by the spectral sensitivity of the silicon-based CCD camera in our spectroscopy system. Parametric fluorescence can be used as a pseudo standard light source for the calibration of instrument spectral response function of a spectroscopy system.

Acknowledgments

We thank Brage Golding and John A. McGuire for valuable discussions and comments. This work was supported by the National Science Foundation (grant DMR-0955944) and Michigan State University.

References and links

1. N. P. Fox, “Primary radiometric quantities and units,” Metrologia 37, 507 (2000) [CrossRef]  .

2. J. Hollandt, J. Seidel, R. Klein, G. Ulm, A. Migdall, and M. Ware, Primary sources for use in radiometry (Academic Press, 2005), vol. 41 of Experimental Methods in the Physical Sciences, pp. 213–290.

3. S. W. Brown, G. P. Eppeldauer, and K. R. Lykke, “Facility for spectral irradiance and radiance responsivity calibrations using uniform sources,” Appl. Opt. 45, 8218–8237 (2006) [CrossRef]   [PubMed]  .

4. J. C. Zwinkels, E. Ikonen, N. P. Fox, G. Ulm, and M. L. Rastello, “Photometry, radiometry and ’the candela’: evolution in the classical and quantum world,” Metrologia 47, R15 (2010) [CrossRef]  .

5. R. Klein, R. Thornagel, G. Ulm, J. Feikes, and G. Wüstefeld, “Status of the metrology light source,” J. Electron. Spectrosc. Relat. Phenom. 184, 331–334 (2011) [CrossRef]  .

6. T. C. Larason and J. M. Houston, “Spectroradiometric detector measurements: Ultraviolet, visible, and near-infrared detectors for spectral power,” NIST Special Publication 250, 41 (2008).

7. J. G. Rarity, K. D. Ridley, and P. R. Tapster, “Absolute measurement of detector quantum efficiency using parametric downconversion,” Appl. Opt. 26, 4616–4619 (1987) [CrossRef]   [PubMed]  .

8. A. L. Migdall, R. U. Datla, A. Sergienko, J. S. Orszak, and Y. H. Shih, “Absolute detector quantum-efficiency measurements using correlated photons,” Metrologia 32, 479 (1995) [CrossRef]  .

9. A. Migdall, “Correlated-photon metrology without absolute standards,” Phys. Today 52, 41 (1999) [CrossRef]  .

10. W. H. Louisell, A. Yariv, and A. E. Siegman, “Quantum fluctuations and noise in parametric processes. I.” Phys. Rev. 124, 1646–1654 (1961) [CrossRef]  .

11. J. P. Gordon, W. H. Louisell, and L. R. Walker, “Quantum fluctuations and noise in parametric processes. II.” Phys. Rev. 129, 481–485 (1963) [CrossRef]  .

12. D. A. Kleinman, “Theory of optical parametric noise,” Phys. Rev. A 174, 1027–1041 (1968).

13. S. E. Harris, M. K. Oshman, and R. L. Byer, “Observation of tunable optical parametric fluorescence,” Phys. Rev. Lett. 18, 732–734 (1967) [CrossRef]  .

14. T. G. Giallorenzi and C. L. Tang, “Quantum theory of spontaneous parametric scattering of intense light,” Phys. Rev. 166, 225–233 (1968) [CrossRef]  .

15. R. L. Byer and S. E. Harris, “Power and bandwidth of spontaneous parametric emission,” Phys. Rev. 168, 1064–1068 (1968) [CrossRef]  .

16. M. M. Choy and R. L. Byer, “Accurate second-order susceptibility measurements of visible and infrared nonlinear crystals,” Phys. Rev. B 14, 1693–1706 (1976) [CrossRef]  .

17. I. Shoji, T. Kondo, A. Kitamoto, M. Shirane, and R. Ito, “Absolute scale of second-order nonlinear-optical coefficients,” J. Opt. Soc. Am. B: Opt. Phys. 14, 2268–2294 (1997) [CrossRef]  .

18. I. Shoji, T. Kondo, and R. Ito, “Second-order nonlinear susceptibilities of various dielectric and semiconductor materials,” Opt. Quantum Electron. 34, 797–833 (2002) [CrossRef]  .

19. B. Y. Zeldovich and D. N. Klyshko, “Field statistics in parametric luminescence,” JETP Lett. 9, 40 (1969).

20. D. C. Burnham and D. L. Weinberg, “Observation of simultaneity in parametric production of optical photon pairs,” Phys. Rev. Lett. 25, 84–87 (1970) [CrossRef]  .

21. Y. H. Shih and C. O. Alley, “New type of Einstein-Podolsky-Rosen-Bohm experiment using pairs of light quanta produced by optical parametric down conversion,” Phys. Rev. Lett. 61, 2921–2924 (1988) [CrossRef]   [PubMed]  .

22. Y. H. Shih, A. V. Sergienko, M. H. Rubin, T. E. Kiess, and C. O. Alley, “Two-photon entanglement in type-II parametric down-conversion,” Phys. Rev. A 50, 23–28 (1994) [CrossRef]   [PubMed]  .

23. P. G. Kwiat, K. Mattle, H. Weinfurter, A. Zeilinger, A. V. Sergienko, and Y. Shih, “New high-intensity source of polarization-entangled photon pairs,” Phys. Rev. Lett. 75, 4337–4341 (1995) [CrossRef]   [PubMed]  .

24. Y. Shih, “Entangled biphoton source - property and preparation,” Rep. Prog. Phys. 66, 1009 (2003) [CrossRef]  .

25. S. Friberg, C. K. Hong, and L. Mandel, “Measurement of time delays in the parametric production of photon pairs,” Phys. Rev. Lett. 54, 2011–2013 (1985) [CrossRef]   [PubMed]  .

26. C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett. 59, 2044–2046 (1987) [CrossRef]   [PubMed]  .

27. C. Chen, B. Wu, A. Jiang, and G. You, “A new-type ultraviolet SHG crystal: β-BaB2O4,” Sci. Sin. Ser. B 28, 235–243 (1985).

28. D. V. Strekalov, A. V. Sergienko, D. N. Klyshko, and Y. H. Shih, “Observation of two-photon ghost interference and diffraction,” Phys. Rev. Lett. 74, 3600–3603 (1995) [CrossRef]   [PubMed]  .

29. N. Boeuf, D. Branning, I. Chaperot, E. Dauler, S. Guerin, G. Jaeger, A. Muller, and A. Migdall, “Calculating characteristics of noncollinear phase matching in uniaxial and biaxial crystals,” Opt. Eng. 39, 1016–1024 (2000) [CrossRef]  .

30. K. Koch, E. C. Cheung, G. T. Moore, S. H. Chakmakjian, and J. M. Liu, “Hot spots in parametric fluorescence with a pump beam of finite cross section,” IEEE J. Quantum Electron. 31, 769–781 (1995) [CrossRef]  .

31. F. Devaux and E. Lantz, “Spatial and temporal properties of parametric fluorescence around degeneracy in a type-I LBO crystal,” Eur. Phys. J. D 8, 117–124 (2000) [CrossRef]  .

32. D. Dehlinger and M. W. Mitchell, “Entangled photons, nonlocality, and bell inequalities in the undergraduate laboratory,” Am. J. Phys. 70, 903–910 (2002) [CrossRef]  .

33. C. K. Hong and L. Mandel, “Theory of parametric frequency down conversion of light,” Phys. Rev. A 31, 2409–2418 (1985) [CrossRef]   [PubMed]  .

34. A. Ling, A. Lamas-Linares, and C. Kurtsiefer, “Absolute emission rates of spontaneous parametric down-conversion into single transverse gaussian modes,” Phys. Rev. A 77, 043834 (2008) [CrossRef]  .

35. M. W. Mitchell, “Parametric down-conversion from a wave-equation approach: Geometry and absolute brightness,” Phys. Rev. A 79, 043835 (2009) [CrossRef]  .

36. M. H. Rubin, “Transverse correlation in optical spontaneous parametric down-conversion,” Phys. Rev. A 54, 5349–5360 (1996) [CrossRef]   [PubMed]  .

37. M. V. Fedorov, M. A. Efremov, P. A. Volkov, E. V. Moreva, S. S. Straupe, and S. P. Kulik, “Spontaneous parametric down-conversion: Anisotropical and anomalously strong narrowing of biphoton momentum correlation distributions,” Phys. Rev. A 77, 032336 (2008) [CrossRef]  .

38. R. S. Klein, G. E. Kugel, A. Maillard, A. Sifi, and K. Polgár, “Absolute non-linear optical coefficients measurements of BBO single crystal and determination of angular acceptance by second harmonic generation,” Opt. Mater. 22, 163–169 (2003) [CrossRef]  .

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Schematic diagram of the crystal and laboratory frame coordinates for parametric down-conversion in a BBO crystal. The phase-matching angle θm is defined as the angle formed by the crystal optical axis (z′) and the pump wave vector (z). The angles θ′s and θs are, respectively, internal and external angles formed by the signal and pump wave vectors. Here the incident pump wave is horizontally polarized, leading to a vertically polarized down-converted signal and idler waves.
Fig. 2
Fig. 2 Schematic diagram of the experimental setup. The transmitted and scattered pump photons are rejected by a miniature beam blocker, a thin-film notch filter, and long-pass filters. The angular and spectral distributions of conical parametric fluorescence are measured by an imaging spectrometer through a Fourier transform optical system. L1, L2 and L3 are convergent lenses with focal length f = 500, 100, and 100 mm, respectively; Obj is an objective (20× N.A.=0.26) with an effective focal length of 10 mm (Mitutoyo Plan Apo infinity-corrected long-working-distance objective); P1 and P2 are Glan–Taylor and Glan–Thompson polarizers; M1 and M2 are silver mirrors; HWP is a half-wave plate for λ = 405 nm; BB is a miniature pump beam blocker; NF is a notch filter (Semrock 405-nm StopLine single-notch filter); and LP represents longpass filters (a Semrock 409-nm blocking edge BrightLine long-pass filter and a Schott GG435 glass filter). When L2 is removed, a real-space fluorescence image is formed at the entrance of the spectrometer with an imaging magnification of 10×. Examples of real-space and angle-resolved fluorescence images are shown with actual dimensions.
Fig. 3
Fig. 3 Angular intensity distribution of parametric fluorescence at 810 nm. (a)–(d) Angle-resolved images of parametric fluorescence for λ = 810 ± 0.5 nm and θm as indicated. The external angle θs is formed by the pump and signal wave vectors in air [see also Fig. 1]. The color palette represents the intensity of the single radiation. Parametric fluorescence is spectrally filtered through a 1-nm bandpass filter with central wavelength λ = 800 nm. The phase-matching angle is adjusted by tilting the BBO crystal with respect to the pump wave vector. The collinear phase-matching angle is set to θm = 28.6° according to a theoretical calculation using indices of refraction given in [29]. (e)–(g) Experimental (blue dashed line) and theoretical (red solid line) cross-sections. The non-collinear phase-matching angle relative to the collinear one can be determined experimentally by measuring the tilting angle of the crystal surface with respect to the pump wave vector.
Fig. 4
Fig. 4 Angular spreads ΔθFWHMof parametric fluorescence at 810 nm. ΔθFWHM is plotted as a function of the inverse of the signal angle (1/θs). Experimental data are represented with error bars as solid circles. The red solid line is the theoretical curve according to Eq. 5, while the dashed line is the theoretical curve including a finite angular resolution of 2 mrad. Selected experimental angular intensity profiles for A (θs = 1.3° = 0.023 rad), B (2.8° = 0.049 rad), and C (3.7° = 0.065 rad) are shown in the inset.
Fig. 5
Fig. 5 Angle-resolved spectra of parametric fluorescence. (a)–(d) Experimental angle-resolved parametric fluorescence images (left panel) and spectra (right panel) for a phase-matching angle θm = 28.6°, 28.8°, 29.1°, and 29.4°, respectively. (e)–(h) Theoretical fluorescence flux calculated according to Eq. (6). The tuning curves for the perfect phase-matching condition are indicated by the white dashed lines on experimental imaging spectra. The color palette represents the calculated photon flux with Δλs = 1 nm and Δθs = 0.5 mrad on a logarithmic scale.
Fig. 6
Fig. 6 Parametric fluorescence flux at 810nm. Integrated degenerate parametric fluorescence flux at λs = 810 nm as a function of the incident pump flux for the collinear case and three phase-matching angles θm of the angle-resolved images shown in Fig. 3. Signal flux is linearly proportional to the pump flux over two order of magnitude, confirming that the dominant signal is spontaneous parametric fluorescence as described by Eq. (7). The slopes η = Ns/Np are 1.2 × 10−10 for θm = 28.6°, 2.6 × 10−10 for θm = 28.8°, and 2.8 × 10−10 for θm = 29.1° and 29.4°.
Fig. 7
Fig. 7 Instrument spectral response function. The normalized spectral density function S(λ) for θm = 29.1° and 29.4°. Ssim is determined from the integration over θs = −7.5° to 7.5° of the calculated angle-resolved spectra as shown in Fig. 5 [see also Eq. (6)]. The black and red curves are S exp * for θm = 29.4° and 29.1°, respectively. The value of Ssim is a constant with 1% standard deviation across wavelengths from 500 to 1000 nm, validating the calculated angular fluorescence spectra and Eq. (7). S exp * ( λ ) = N s * ( λ ) λ s 4 λ i 2, where N s * ( λ ) is the integration over θs ≈ −15° to 15° of the experimental imaging spectra N s * ( λ s , θ s ). S exp * ( λ ) represents a relative instrument spectral response function (ISRF) of the optical spectroscopy system, including optical filters and a Glan-Thompson polarizer along the path of the fluorescence, a liquid-nitrogen-cooled CCD (PI-Acton Spec-10:400BR), and a 300g/mm plane ruled reflectance grating with 1000-nm blaze wavelength (PI-Acton 750-1-030-1). The polarization of the fluorescence is vertically polarized. The absolute ISRF (or collection efficiency) can be deduced by calibrating S exp * ( λ ) at a fixed wavelength such as λ = 810 nm in our experiments.

Tables (1)

Tables Icon

Table 1 Parametric fluorescence efficiency η = 2Ns/Np for 810 ± 0.5 nm

Equations (14)

Equations on this page are rendered with MathJax. Learn more.

ω p = ω s + ω i ,
k p = k s + k i ,
n p = n s cos ( θ s ) ,
n ˜ ( θ m , λ ) = ( cos ( θ m ) 2 n o ( λ ) 2 + sin ( θ m ) 2 n e ( λ ) 2 ) 1 2 .
Φ = exp ( 1 2 W 2 ( Δ k x 2 + Δ k y 2 ) ) ( sin ( 1 2 L Δ k z ) 1 2 L Δ k z ) 2 = exp ( 1 2 κ 2 W 2 ) sinc 2 ( 1 2 L Δ k z ) .
Δ θ FWHM ( θ s ) = 2 × 0.886 π L × | Δ k z / θ s | 2.783 × n s L × k s × θ s ,
Δ λ FWHM ( λ s ) = 2 × 0.886 π L × | Δ k z / λ s | 0.443 × λ s 2 L × n s cos ( θ s / n s ) ,
N s ( ω s , κ s ) = d eff 2 ω s ω i ω p L 2 N p 2 π 4 c 3 ε 0 n s n i n p d ω s d 2 κ s d 2 ξ exp ( 1 2 ξ 2 ) sinc 2 ( 1 2 L Δ k z ) ,
N s = d eff 2 L ω s 2 ω i 2 π c 4 ε 0 n p 2 N p d ω s = ( 2 π ) 4 2 c d eff 2 L ε 0 n p 2 λ s 4 λ i 2 N p d λ s .
Δ k z = k s z + k i z k p z = k s 2 κ s 2 + k i 2 κ i 2 k p z , and κ = κ s + κ i .
N s ( ω s , κ s ) = d eff 2 ω s ω i ω p L 2 N p 8 π 4 c 3 ε 0 n s n i n p d ω s d 2 κ s d 2 ξ i exp ( 1 2 | ξ s + ξ i | 2 ) sinc 2 ( 1 2 L Δ k z )
d 2 ξ i exp ( ξ s 2 + ξ i 2 2 ξ s ξ i cos ( φ ) 2 ) ξ i d ξ i exp ( ( ξ s ξ i ) 2 2 ) d φ exp ( ξ s ξ i φ 2 2 ) 2 π d ξ i exp ( ( ξ s ξ i ) 2 2 ) .
N s ( ω s , κ s , φ ) = d eff 2 ω s ω i ω p L 2 N p 8 π 4 c 3 ε 0 n s n i n p 2 π d ω s κ s d κ s d ϕ d ξ i e 1 2 ( ξ s ξ i ) 2 sinc 2 ( 1 2 L Δ k z ) ,
N s ( λ s , θ s ) = 2 π 2 π d eff 2 ω p L 2 N p ε 0 n s n i n p λ s 5 λ i sin ( 2 θ s ) d λ s d θ s d ξ i e 1 2 ( ξ s ξ i ) 2 sinc 2 ( L Δ k z 2 ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.