spectra of 110-cut GaAs, GaP, and Si near the two-photon absorption band edge
Abstract
Spectra of the degenerate two-photon absorption coefficient , anisotropy parameter , and dichroism parameter of crystalline 110-cut GaAs, GaP, and Si at 300 K were measured using femtosecond pump-probe modulation spectroscopy over an excitation range in the vicinity of each material’s half-band gap (overall eV, or nm). Together these three parameters completely characterize the three independent components of the imaginary part of the degenerate third-order nonlinear optical susceptibility tensor . In direct-gap GaAs, these components peak at which is close to the peak at predicted by the Jones-Reiss phenomenological model. The dispersion is comparable to ab initio calculations. In indirect-gap GaP and Si, these components tend to increase with over our tuning range. In Si, the dispersion differs significantly from predictions of semi-empirical models and ab initio calculations do not account for transitions below the two-photon direct band gap, motivating further investigation. Kleinman symmetry was observed to be broken in all three materials. We also note anomalies observed and their possible origins, emphasizing the advantages of a 2-beam experiment in identifying the contribution of various nonlinear effects.
I Introduction
Column III-V and IV semiconductors are not only building blocks of modern microelectronics, but have become important photonic materials for telecommunications and integrated light-wave systems.Asghari and Krishnamoorthy (2011) Direct-gap gallium arsenide (GaAs) is used in photovoltaicYoon et al. (2010); Bett et al. (1999) and optoelectronicYoon et al. (2010); Wada (1988) devices. Among indirect-gap semiconductors, gallium phosphide (GaP) is a key component of green LEDsPilkuhn and Foster (1966) and nanophotonic devices,Wilson et al. (2020) and an emerging platform for integrated nonlinear photonics,Wilson et al. (2020) while silicon (Si) has important applications in all-optical switches,Euser and Vos (2005) photonic crystals,Vlasov et al. (2001) lasers,Rong et al. (2005) and optical waveguides.Tsang et al. (2002) In photonic systems utilizing ultrashort light pulses, long propagation lengths, or high light intensities ( MW/cm2), nonlinear optical properties of these semiconductors come into play. In particular, two-photon absorption (2PA), which is proportional to the imaginary part of the third-order nonlinear optical susceptibility tensor , limits optical switching energyDvorak et al. (1994); Blanco et al. (2000) and affects long-haul pulse transmission at telecommunication wavelengths and nm.Bristow, Rotenberg, and van Driel (2007); Tsang et al. (2002) Nonlinear refraction, which is proportional to , but depends on the dispersion of the degenerate as well as the nondegenerate via nonlinear Kramers-Kronig relations,Bassani and Scandolo (1991); Caspers (1964); Hutchings et al. (1992); Kogan (1963); Gibbs (1972); Price (1964); Jr. and Jr. (1975); Sheik-Bahae, Hagan, and Styland (1990); Sheik-Bahae et al. (1991); Toll (1956) underlies all-optical switchingGibbs (1985); Stegeman and Wright (1990) and long-haul light pulse propagation.Reintjes and McGroddy (1973) Accurate measurement and calculation of is essential to understanding the operation and limits of diverse photonic systems.
Early characterization of in GaAs, GaP, and Si consisted of single-wavelength measurements of degenerate 2PA amplitude using pulsed lasers with picosecond durations.Reintjes and McGroddy (1973); Bechtel and Smith (1976); Bosacchi, Bessey, and Jain (1978) Reported values of the extracted 2PA coefficient at a given wavelength often varied over an order of magnitude or more.Stryland et al. (1994); Sheik-Bahae and Stryland (1999) With the advent of tunable optical parametric amplifiers (OPAs) providing femtosecond pulses at kHz repetition rate, 2PA spectra of Si were measured with improved accuracy using single-beam excitation over a range eV at peak intensities GW cm-2.Bristow, Rotenberg, and van Driel (2007) This intensity induced nonlinear aborption as large as , but can also introduce competing nonlinearities [e.g. three-photon absorption (3PA) and free-carrier absorption (FCA)Stryland et al. (1994)] that can be challenging to distinguish from 2PA.Sheik-Bahae and Stryland (1999) Above-gap 2PA spectra of Si were also measured over a range of eV.Reitze et al. (1990) 2PA spectra of GaAs were also measured using single-beam excitation over a limited range of eV,Hurlbut et al. (2007) and a degenerate pump-probe experiment over a limited range of eV.Fishman et al. (2011) To our knowledge two-beam 2PA anisotropy in these three semiconductors, i.e. the dependence of 2PA on the angle between incident light polarization and crystollographic axes, has only been completely characterized at a single wavelength in GaAs.Dvorak et al. (1994)
In this paper, we report spectroscopic measurements of the amplitude and anisotropy of degenerate 2PA in GaAs, GaP, and Si using two-beam pump-probe modulation spectroscopy.Shank, Ippen, and Shapiro (1978); Dvorak et al. (1994) In this technique the 1 kHz pulse train from a tunable fs OPA is split into a pump beam that is chopped at a sub-harmonic of its repetition rate and focused loosely into the samples and a weak sychronized probe beam of the same wavelength that intersects the central portion of the pump beam path at a small angle inside the sample. Although alignment is more complicated, 2-beam 2PA offers two advantages over single-beam measurements. First, by detecting pump-induced probe transmission changes with lock-in techniques, the setup becomes sensitive to sub-1% absorption, excited by pump intensity GW cm-2. The magnitude of 2PA in co-polarized beams is also doubled relative to the single-beam case.Dvorak et al. (1994); Sheik-Bahae and Stryland (1999) Competing higher-order nonlinearities are therby more effectively avoided than in single-beam measurements at higher intensity. Second, by scanning the time delay between pump and probe pulses, the shape of the resulting cross-correlation signal reveals signatures of competing nonlinear processes such as FCA, 3PA,Benis et al. (2020) and 2-beam pump-probe coupling, when present. Absence of such signatures, i.e. an instantaneous or nearly-instantaneous cross-correlation response, provides strong in-situ evidence that data is being acquired in a pure 2PA regime. From combined measurements of 2PA amplitude and anistropy spectra, we obtain not only degenerate , as in past work, but anisotropy parameter and as well, enabling us to extract complete spectra of GaAs, GaP, and Si near their respective 2PA band edges for the first time. spectra are ultimately better suited than alone for quantitative comparison with first-principles calculations.
The paper is organized as follows: Section II adapts the standard theory of single-beam 2PA to the current 2-beam context, expands it to include 2PA anisotropy, and defines approximations used in our data analysis. Section III describes our 2-beam experimental setup, presents and discusses experimental results, and examines anomalies and limitations of this technique. Section IV summarizes our main findings.
II Theory
II.1 Pump-probe two-photon absorption
Up to 2-photon interactions, loss processes in a 2-beam open-aperture experiment can be characterized by the differential equationsCirloganu (2010)
(1) | ||||
(2) | ||||
(3) |
where is the (probe, pump) intensity, is the (pump-probe, pump-pump) 2PA coefficient, is the FCA cross section, and is the free-carrier density. For , probe-probe 2PA is neglected and both probe-probe and probe-pump excitations of free-carriers can also be neglected. We also only consider depletion of the pump due to pump-pump 2PA, since that term dominates over all other terms. For excitation below the band gap , linear absorption is negligible. We also assume that the free-carrier recombination rate is , much greater than the laser reptition rate and much less than the inverse of the pulse duration, and thus relaxation on the timescale of a pulse is neglected but the system is assumed to be fully relaxed before the arrival of the subsequent pulse. Exceptional cases in which other 2-beam coupling processes (e.g. photorefractive phase gratings, free carrier gratings, nonlinear refractive index gratings) influence the results are discussed in Section III.3.1. Eqs. 1 – 3 are strictly valid for co-propagating pump and probe beams, but accurately approximate experimental geometries, such as those used here, with small intersection angles (here ), large pump/probe beam-size ratios (here , and interaction lengths much shorter than the pump beam size.
The incident pulse intensity profiles (at ) can be expressed as
(4) |
where labels the beams, the on-axis peak intensity of beam is
(5) |
is the beam radius, is the full-width at half-maximum (FWHM) Gaussian pulse duration, is the incident pulse energy, and is the Fresnel power reflection coefficient. The relative time delay between the pump and probe pulses is defined as
(6) |
where is the Kronecker delta. If the sample thickness is shorter than the Rayleigh range , then beam radii can be considered constant with respect to . In the limit that FCA is weak and only considered to first-order, solving Eqs. 1 - 3 for the intensity profile in Eq. 4, and integrating over from to , from to , and from to and expanding about gives an analytical expression for the change in probe transmission when the pump is on from when the pump is off, normalized to the probe transmission when the pump is off, . The result is
(7) |
where
(8) | ||||
(9) |
characterize the magnitude of pump-probe 2PA and
(10) | ||||
(11) |
characterize the magnitude of FCA. The symbols refer to the cases of co-polarized and cross-polarized beams, resepectively. The generalized binomial coefficients in the cross-polarized case are
(12) |
which are valid so long as for and . The quantity relates the pump-pump 2PA coefficient to the cross-polarized pump-probe 2PA coefficient by
(13) |
The quantity is the ratio of the probe to pump beam areas. Pump-probe 2PA produces a characteristic 2nd-order autocorrelation dip in probe transmission while FCA causes an approximately step-function decrease in probe transmission for positive delays. The undepleted pump regime is the case of and higher-order terms capture the effect of pump depletion by pump-pump 2PA.
II.2 2PA anisotropy
2PA is “anisotropic” when it depends on the relative orientations of incident fields and crystallographic axes. For energetically degenerate, nearly collinear beams co-propagating along the laboratory -axis, the probe and pump fields in laboratory coordinates are
(14) | ||||
(15) |
where are the complex field amplitudes, , , is the angle between linearly-polarized probe and pump fields, are the phases of the complex field amplitudes, and is an arbitrary phase shift between the pump and probe fields (assumed to not vary in time over the duration of the pulse such that the two pulses are mutually coherent). The field experienced by the nonlinear medium is just the sum field , and the intensity of the fields as defined isBoyd (2008)
(16) |
where is the vacuum permittivity, the refractive index, and the speed of light. 2PA is a third-order nonlinear optical process. If the beams are co-polarized, i.e. , then the Fourier component is
(17) |
which is the nonlinear polarization density induced by the incident field that governs nonlinear propagation of probe field . Here, the factor of 3 arises from the sum over allowed frequency permutations and is the degenerate third-order nonlinear optical susceptibility tensor describing a polarization density response at the same frequency as the incident frequency.
Similarly, the expression for the Fourier component for cross-polarized beams, i.e. , is
(18) |
Solving the nonlinear wave propagation equation for these fields in the plane-wave basis and approximating slowly-varying field amplitudes gives an expression relating to the imaginary part of the third-order nonlinear optical susceptiblity tensor by
(19) | ||||
(20) |
where is the refractive index, the speed of light, and refers to the two-beam degenerate 2PA coefficient for co-polarized beams () which is twice the 2PA coefficient for a single beam with the same polarization and propagation direction, . is the two-beam degenerate 2PA coefficient for cross-polarized beams (). See the Supplementary Material for the complete derivation of these relationships.
In our experiments, the beams propagate along in crystallographic coordinates, with and defining a basis of linear polarization directions. We therefore rotationally transform using the relation
(21) |
using rotation matrix
(22) |
where is the probe polarization angle relative to the -axis, or equivalently, the sample rotation angle, in order to express Eqs. 19 – 20 in terms of tensor components in crystallographic coordinates. This enables us to take advantage of the (GaP and GaAs) or (Si) crystal symmetry for which there are only three independent, non-vanishing components; , , and , where .Popov, Svirko, and Zheludev (1995) Here, intrinsic permutation (but not Kleinman)Boyd (2008); Kleinman (1962) symmetry has been applied. can then be written
(23) | ||||
(24) |
where to allow for an arbitrary initial orientation, and
(25) |
is the anisotropy parameter and
(26) |
is the indicated ratio of tensor components. The difference in magnitude of 2PA between linear and circular polarizations in single-beam experiments is related to the dichroism parameter, . Under the usual weak 2PA conditions there is no population inversion, which constrains and . See the Supplementary Material for the derivation of these constraints. Thus the independent components of of a or crystal can be completely characterized by measuring 2-beam 2PA while rotating the sample about in both co-polarized () and cross-polarized () geometries.
II.3 Dispersion and anisotropy models
Ab initio calculations using density functional theory (DFT) have been performed by Murayama and Nakayama for in GaAs and Si.Murayama and Nakayama (1995) The calculations are applicable for GaAs over the entire spectral range, but for Si only for direct transitions above the two-photon direct-gap. 2PA in Si between the two-photon indirect-gap eV and the two-photon direct-gap eV is most likely due to phonon-mediated transitions, as in linear absorption, but could also be affected by the presence of defect states or dopants.Noffsinger et al. (2012)
We can model the dispersion in using the phenomenological models developed by Jones and Reiss for direct band gap semiconductorsJones and Reiss (1977); Wherrett (1984) and Garcia and Kalyanaraman for indirect band gap semiconductors.Garcia and Kalyanaraman (2006) The Jones-Reiss model for direct-gap semiconductors (GaAs) has the form
(27) |
where and is a material-dependent constant which functions as a fit parameter. This function peaks at a value of . The Garcia-Kalyanaraman model for indirect-gap semiconductors (GaP, Si) considers parabolic electron and hole bands with indirect matrix elements and also accounts for “allowed-allowed” (a-a), “allowed-forbidden” (a-f), and “forbidden-forbidden” (f-f) transitions. 2PA dispersion takes the form
(28) |
in the limit where the phonon energy and in the absence of a DC-field. is a material-dependent constant which functions as a fit parameter. This function peaks at . The 2PA dispersion in the presence of a DC-field for both direct-gapGarcia (2006) and indirect-gapGarcia and Kalyanaraman (2006) semiconductors acquires an exponential tail for due to tunneling-assisted 2PA. While these models are not based on full band-structure calculations, they have achieved some success in reproducing the dispersion observed in published results using single-beam techniques.
The anisotropy parameter was estimated using a 3-band model where the dominant anisotropy is due to a-f transitions givingDvorak et al. (1994)
(29) |
where () are momentum matrix elements for () transitions and () are the direct band gaps between the valence and the (conduction, higher conduction) bands. The values of and for semiconductors are not reliably knownDvorak et al. (1994) but we still apply this model to GaAs for predicting the 2PA anisotropy near the -point which corresponds to the two-photon band edge. The model is still reasonable for GaP for transitions near the direct-gap energy of 2.78 eV at the -point, which is only 25% higher than the indirect-gap, as there are few available indirect transitions. It is not straightforward to apply this model to Si due to the nearly parallel band structure, necessitating a calculation of the mean contribution of anisotropy over the -space whose direct transitions and available indirect transitions at 300 K are comparable in energy within the bandwidth of the excitation source. We report the theoretical values for the anisotropy near the two-photon band edge for a range of values for the momentum matrix elements from literature for GaAs and GaP in Table 1.
Parameter | GaAs | GaP |
---|---|---|
0.60,Aspnes (1972) (0.71, 1.06)Moss, Sipe, and van Driel (1987) | 0.57,Aspnes (1972) (0.71, 1.06)Moss, Sipe, and van Driel (1987) | |
(eV) | 1.44Dvorak et al. (1994) | 2.78Zallen and Paul (1964) |
(eV) | 4.5Dvorak et al. (1994) | 3.71Zallen and Paul (1964) |
-0.38, (-0.45, -0.68) | -0.85, (-1.06, -1.59) |
III Experiment
III.1 Experimental technique
III.1.1 Pump-probe modulation spectroscopy setup
Fig. 1 shows the pump-probe modulation spectroscopy (PPMS) experimental setup. The light source is a Light Conversion TOPAS-C optical parametric amplifier (tuning range nm) pumped by a 1 kHz Coherent Libra HE USP titanium-doped sapphire regenerative amplifier (nominally nm, mJ, fs). Spectral filters attenuate undesired frequency components. A spatial filter consisting of two plano-convex mm lenses and a pinhole (of diameter selected based on beam parameters) homogenizes and shapes the transverse beam profile to approximate a Gaussian. These filters were configured for each excitation wavelength. A linear thin-film polarizer ensures definite polarization. An achromatic half-wave plate (HWP) controls probe polarization relative to the pump. Plano-convex lenses L1 and L2 (with same focal length or 500 mm, depending on beam parameters) focus pump and probe beams to a common spot on the sample. Temporal and spatial overlap was optimized by monitoring two-beam second-harmonic generation (SHG) from a barium borate crystal in the sample position. The optical chopper (OC) was sychronized to the fifth subharmonic of the laser repetition rate ( Hz) where spectral noise was relatively low, and could be configured to modulate only the pump beam (PPMS experiment) or both beams (probe absorption calibration). The outputs of all photodiode detectors (Thorlabs DET100A for 325 - 1100 nm, Thorlabs PDA30G for 1200 - 2000 nm) were integrated and held for each pulse until the next trigger by Stanford Research Systems SR250 gated integrators. The probe and pump integrated signals were the inputs for two Stanford Research Systems SR830 DSP lock-in amplifiers referenced to the optical chopper TTL output. A Tektronix DPO 2014 oscilloscope recorded the outputs of the lock-in amplifiers, which were digitally recorded by a computer. The computer controlled pump-probe time delay, sample Z-scan to optimize pump-probe spatial overlap, and sample rotation.

III.1.2 Pump and probe spatiotemporal profiles
The pulse duration was measured by performing a SHG second-order autocorrelation. Pump power was measured with a Coherent FieldMate Si photodiode (PD) head (650 - 1100nm) or with a reference PD calibrated with a thermal head (1200 - 2000 nm). In order to reduce sensitivity on overlap position and geometric effects, the probe radius was kept smaller than the pumpDiels and Rudolph (1996) (the beam waists were not at the sample plane) which necessitated attenuating the probe pulse energy with neutral density filters lower than the pump to keep the fluence ratio . Pump and probe beams impinged on each sample at from the surface normal, and at with respect to each other. The latter angle became (i.e. approximately collinear) inside each sample after taking refraction into account. The beam radii at the overlap position were measured using an automated knife-edge technique, and measurements at and mm from the overlap were used to measure the beam angles and Rayleigh ranges.
III.1.3 Samples
The samples were semi-insulating undoped n-type GaAs (110) with resistivity – E7 cm, intrinsic carrier concentration – E7 cm-3, carrier mobility – E3 cm2 V-1 s-1, etch pitch density cm-2, and thickness mm; undoped GaP (110) with mm; and undoped Si (110) with – cm and mm. The spatial and temporal overlap of the beams was fine-tuned for each sample using 2PA prior to all data acquisition scans. The linear optical properties of the samples were measured using a J.A. Wollam M2000 variable-angle spectroscopic ellipsometer (VASE) and are reported in the Supplementary Material. Linear optical properties for nm were obtained from literature.Skauli et al. (2003); Li (1980)
III.2 Data collection and analysis procedure
III.2.1 Time delay scans
The 2PA coefficient was determined by performing a time delay scan and fitting the normalized probe transmission to Eq. 7. The fitting algorithm compared the variation in the fitting parameter for the and cases, and if the variation was below a threshold of , then the result for the case was reported; if the variation was larger, the process was repeated, comparing successively higher-order terms. The normalized change in probe transmission was calculated by
(30) |
where is the probe lock-in amplitude during the time delay scan and is the mean lock-in amplitude corresponding to absorption when the probe is modulated by the optical chopper during a calibration scan. After fitting, was adjusted by the vertical offset parameter which removes the effect of any scattered pump light. When multiple scans were performed at a single excitation energy, the mean value of was reported and the uncertainty was conservatively quantified by the maximum of the set including the standard error of the mean, the propagated uncertainty of the mean, and the propagated uncertainty in the individual measurements.
III.2.2 Rotation scans
The values of and were determined by performing rotation scans of the sample fixed at the spatial overlap position and in both co-polarized and cross-polarized geometries. Power fluctations can sometimes complicate the signal, and so the magnitude of the pump-induced absorption was corrected according to
(31) |
and
(32) |
where is a reference voltage monitoring the incident power, is the lock-in amplitude of the probe at large negative time delay (and thus negligible 2PA, capturing any offset), and is the mean of the probe lock-in amplitude normalized to the reference amplitude during the calibration scan.
Zincblende semiconductors can exhibit the photorefractive effect which can lead to transient two-beam coupling.Dvorak et al. (1994) Transient energy transfer between beams propagating along the -direction due to the photorefractive effect is antisymmetric about rotations of , while 2PA is symmetric.Dvorak et al. (1994) Analyzing only the symmetric part removes any contribution from the photorefractive effect. In addition, random fluctuations are also reduced. The symmetric and antisymmetric parts of the change in probe transmission are
(33) |
where and are the symmetric and antisymmetric parts, respectively, of the change in probe transmission upon rotations of .
III.3 Results
III.3.1 Time delay scans
Fig. 2 (a) shows an example of a time delay scan for GaP at eV in co-polarized geometry. Baseline probe absorption returned to zero at both positive and negative delays fs, indicating negligible FCA. We observed the same symmetry in all scans presented in this study.

For fs, the width of the measured 2PA response (data points) fits very closely to the calculated response (solid black curve) with the pulse duration independently measured by SHG second-order autocorrelation. We obtained a similar match for most data presented here. For a small number of exceptional cases, we observed a 2PA response slightly wider than the SHG autocorrelation. For these cases, we considered the possiblity of saturation effect and this led to a widening of error bars in reported spectra (see Supplementary Material).
In addition to revealing complete information regarding the tensor structure, the 2-beam PPMS technique offers another advantage over single-beam techniques in its ability to discern anomalous effects from 2PA. Such effects include FCA, transient energy transfer originating from photorefractive phase gratings, nonlinear refractive index phase gratings, 2PA amplitude gratings, free-carrier gratings, and saturable absorption which can couple the two beams.Smirl et al. (1988); Valley and Smirl (1988)
The temporal width of the nonlinear loss was anomalously broadened in GaAs and GaP as the excitation neared their respective band gaps in both polarization geometries. Fig. 2 (b) shows this anomaly in GaP at eV (0.85) and it was also observed in GaAs at eV (0.96). Pump depletion by pump-pump 2PA cannot account for this effect. The observed temporal FWHM in the nonlinear loss was a factor of 2.3 – 3 greater than expected for 2PA alone. Asymmetry in the time delay scans also supports the presence of other effects not accounted for in our model. Both of these features are obvious against the second-order autocorrelation signal expected for 2PA (as measured by SHG) which is superimposed on the data.
A fs-scale time delay-dependent FCA cross-section could produce such an anomaly as free carriers excited predominantly by pump-pump 2PA diffuse through -space in the conduction band. Zero time delay in this experiment was determined by a fit parameter, but the peak could be shifted towards positive time delays as a result of this effect. Another possibility is that a photorefractive index grating could induce transient energy transfer and generate the observed anomaly. Both of these effects are encompassed more generally by any non-instantaneous response or memory effect as the 2PA process approaches resonance. If such an effect is present it can induce a polarization density responsive not only to the instantaneous value of the electric field but also to its history on fs timescales.Diels and Rudolph (1996) Measurements at different crystal orientations, pulse energies, pulse durations, and a nondegenerate two-photon absorption experimentFishman et al. (2011); Hutchings and Stryland (1992); Bolger et al. (1993); Hannes et al. (2019) could reveal more information about the nature of such anomalies.
Below the two-photon absorption band edge of GaAs and GaP, nonlinear loss was very small except at high intensities (a factor of 1 – 20 greater than intensities at excitation energies ). This is consistent with 2PA being energetically disallowed but a transient coherent artifactLebedev et al. (2005) was still observed near zero time delay in both polarization geometries. A transient coherent artifact was also observed in Si near to but slightly above at eV, where the magnitude of 2PA was also quite low. This coherent artifact was asymmetric and notably included an antisymmetric component (about ) where briefly for positive time delays. While rapidly at the two-photon band edge, in general does not. encodes the nonlinear refractive index which is related to by nonlinear Kramers-Kronig relations.Boyd (2008) Below resonance, in semiconductors and thus is the dominant third-order nonlinearity at low excitation energies.Boyd (2008)
This coherent artifact shown in Fig. 2 (c) likely arises from a nonlinear refractive index phase grating, which can transfer energy between the probe and pump and produce such antisymmetric time delay profiles.Palfrey and Heinz (1985) This effect is a background effect that likely occurs in all measurements, but is so weak that it is only visible when 2PA is no longer the dominant third-order nonlinearity.
III.3.2 2PA spectra
The 2PA spectrum for GaAs measured by PPMS in co-polarized geometry is shown in Fig. 3 along with the fit using the Jones-Reiss model in Eq. 27, published values using mostly single-beam techniques, and ab initio calculations. The best fit of this data to the Jones-Reiss model gives cm GW-1 and a 2PA peak of cm GW-1. The 2PA coefficient reported in literature has a large spread, mainly from single-beam experiments clustered around 1053 nm using picosecond lasers.



The 2PA spectrum for GaP measured by PPMS in co-polarized geometry is shown in Fig. 4 along with a fit using a cumulative model for direct and indirect transitions, a fit using the Jones-Reiss model for direct transitions near , and a comparison to published values using single-beam techniques. The cumulative model is defined as using the Garcia-Kalyanaraman model for , and a sum of this function and the Jones-Reiss model above . The best fit to this model gives cm GW-1 and cm GW-1, with 2PA contributions peaking at cm GW-1 for indirect transitions and cm GW-1 for direct transitions.
The 2PA spectrum for Si measured by PPMS in co-polarized geometry is shown in Fig. 5 along with the fit using the Garcia-Kalyanaraman model and a comparison to published values using single-beam techniques. The best fit of this data to the Garcia-Kalyanaraman model gives cm GW-1, peaking at cm GW-1, which is considerably lower than the fit to the data from Bristow, et al.Bristow, Rotenberg, and van Driel (2007) of cm GW-1, peaking at cm GW-1. The measured dispersion differs from the phenomenological model. These results are tabulated in the Supplementary Material. An inset in Fig. 5 shows the degenerate 2PA spectrum predicted by ab initio calculations valid above eV and published values using a two-beam technique.
III.3.3 Rotation scans
An example of a rotation scan for GaP at eV showing symmetric and antisymmetric parts of the normalized in co- and cross-polarized geometries is shown in Fig. 6. The fits of the co- and cross-polarized symmetric data share global parameters and , while the other parameters are independent. The antisymmetric part is typically of the magnitude of the symmetric part.

III.3.4 Anisotropy spectra
The spectra of anisotropy and dichroism parameters for GaAs, GaP, and Si are reported in Table 2, along with the calculations of the relative amplitudes of and to . The values of from this work are compared to the range of theoretical predictions of for GaAs and GaP from Table 1 as well as results reported in literature for GaAs and Si. is largest for GaP, which is consistent with theoretical predictions at the two-photon band edge. The values of from this work are compared to a previously reported value in literature for GaAs, and ab initio calculations for anisotropy in GaAs are also shown. See the Supplementary Material for additional information on comparing values between different sources in literature.
Sample | (eV) | Notes | ||||||||
GaAs | 0.72 | -0.38 | Theoretical: see Table 1 | |||||||
0.72 | -0.45 | Theoretical: see Table 1 | ||||||||
0.72 | -0.68 | Theoretical: see Table 1 | ||||||||
0.77 | -0.13 | 0.08 | 0.68 | 0.17 | 0.20 | 0.13 | 0.74 | 0.17 | ||
0.88 | -0.65 | -0.50 | 0.91 | -0.17 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
0.89 | -0.36 | 0.06 | 0.0 | 0.3 | 0.58 | 0.18 | 0.2 | 0.3 | ||
0.93 | -0.46 | -0.01 | 0.62 | 0.22 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
0.95 | -0.29 | 0.05 | 0.47 | 0.11 | 0.34 | 0.08 | 0.61 | 0.10 | ||
0.95 | -0.56 | 0.02 | 0.63 | 0.29 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.03 | -0.27 | 0.08 | 0.4 | 1.6 | 0.4 | 0.8 | 0.5 | 1.6 | ||
1.03 | -0.47 | 0.15 | 0.54 | 0.39 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.13 | -0.40 | 0.05 | 0.42 | 0.12 | 0.39 | 0.07 | 0.62 | 0.09 | ||
1.13 | -0.58 | 0.26 | 0.51 | 0.55 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.17 | -0.224 | 0.011 | Bepko, et al.Bepko (1975) | |||||||
1.17 | -0.3 | Bechtel, et al.Bechtel and Smith (1976) | ||||||||
1.17 | -0.74 | 0.18 | DeSalvo, et al.DeSalvo et al. (1993) | |||||||
1.17 | 0.00 | 0.15 | Said, et al.Said et al. (1992) | |||||||
1.17 | -0.57 | 0.28 | 0.50 | 0.55 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.23 | -0.64 | 0.29 | 0.51 | 0.61 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.24 | -0.23 | 0.05 | 0.64 | 0.01 | 0.24 | 0.06 | 0.75 | 0.07 | ||
1.31 | -0.76 | 0.08 | 0.22 | 0.24 | 0.58 | 0.08 | 0.60 | 0.20 | Dvorak, et al.Dvorak et al. (1994) | |
1.31 | -0.64 | 0.34 | 0.49 | 0.66 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
1.48 | -0.66 | 0.34 | 0.49 | 0.67 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) | |||||
GaP | 1.38 | -0.53 | 0.06 | 0.09 | 0.17 | 0.59 | 0.10 | 0.35 | 0.14 | |
1.39 | -0.85 | Theoretical: see Table 1 | ||||||||
1.39 | -1.06 | Theoretical: see Table 1 | ||||||||
1.39 | -1.59 | Theoretical: see Table 1 | ||||||||
1.55 | -0.88 | 0.10 | -0.13 | 0.23 | 0.79 | 0.14 | 0.31 | 0.18 | ||
1.77 | -0.61 | 0.09 | 0.1 | 0.7 | 0.6 | 0.3 | 0.4 | 0.6 | ||
Si | 0.58 | 0.00 | 0.05 | Bristow, et al.Bristow, Rotenberg, and van Driel (2007) | ||||||
0.77 | -0.55 | 0.09 | 0.1 | 0.4 | 0.60 | 0.20 | 0.4 | 0.3 | ||
0.89 | -0.03 | 0.04 | 0.95 | 0.08 | 0.04 | 0.05 | 0.96 | 0.06 | ||
0.89 | 0.00 | 0.05 | Bristow, et al.Bristow, Rotenberg, and van Driel (2007) | |||||||
0.95 | -0.15 | 0.14 | 0.77 | 0.23 | 0.16 | 0.15 | 0.84 | 0.16 | ||
1.03 | -0.29 | 0.08 | 0.2 | 0.6 | 0.5 | 0.3 | 0.3 | 0.6 | ||
1.65 | -0.08 | 0.50 | 0.18 | 0.54 | Ab initio: Murayama, et al.Murayama and Nakayama (1995) |
IV Discussion

The complete set of measurements performed ranged from 0.62 – 1.38 eV for GaAs, 0.77 – 1.91 eV for GaP, and 0.62 – 1.03 eV for Si. Long-lifetime time-independent FCA was observed to be negligible in the regime studied in this work as evidenced by the complete recovery of the transmission of the probe at sufficiently large positive time delays. The antisymmetric part of in the rotation scans was generally small in comparison to the symmetric part, and thus the photorefractive effect did not significantly affect the anisotropy measurements in the regime studied. However, other anomalies were observed within this range for which the model for 2PA considered in this work was unable to explain, as detailed in Section III.3.1. The anomaly observed in time delay scans characterized by asymmetric temporal broadening and strong nonlinear loss precluded an estimation of at 1.38 eV for GaAs and 1.91 eV for GaP, near their respective band gaps. This anomaly indicated the presence of a transient effect of the pump on the probe which depended on the history of the incident field. A distinctly different anomaly was observed in time delay scans characterized by an antisymmetric component with an increase in probe transmission for positive delays for which the likely source was a nonlinear refractive index phase grating. This precluded an estimation of at 0.62 – 0.69 eV for GaAs, 0.77 – 1.24 eV for GaP, and 0.62 – 0.69 eV for Si, near their respective two-photon band edges. The co-polarized rotation scans often did not follow the expected form of 2PA for or crystal symmetry when the antisymmetric time delay scan anomaly was observed. However, some of the rotation scans did still match the expected form even with the presence of a time delay scan anomaly which suggests that 2PA may still be dominant at zero time delays. The complete set of measurements is outlined in the Supplementary Material. The 2PA spectra presented here reflect only the data for which no anomaly was observed.
The dispersion observed in GaAs in Fig. 3 indicates that the spectral structure of 2PA is more complex than predicted by the Jones-Reiss model. The peak in 2PA measured at 1.13 eV is close to that expected from the Jones-Reiss model of 1.03 eV. The measured spectra is comparable in magnitude to ab initio calculations. The number of surviving spectral measurements of in GaP and Si are too few to make compelling statements about the agreement with phenomenological models, but this also illustrates the importance of identifying anomalous responses not attributable to 2PA which is often only possible with a 2-beam experiment. Measurements of 2PA in GaP at increased spectral density near could better constrain the relative strengths of direct and indirect two-photon transitions. 2PA in Si between eV and eV is due to either phonon-mediated transitions or transitions involving defect or dopant states. Measurements of the degenerate 2PA spectrum dependence on temperature in this spectral range could verify the relative contributions to 2PA by these two mechanisms. The ab initio calculations of 2PA in Si by Murayama, et al. consider only direct transitions and thus are applicable in describing 2PA above .Murayama and Nakayama (1995)
The parameters , , and characterize the three independent components , , and . The relative amplitudes of the off-diagonal tensor components to the component are shown in Table 2, and the absolute amplitude of the tensor spectra for GaAs, GaP, and Si are shown in Fig. 7 and compared with results from literature for GaAs.Dvorak et al. (1994) See the Supplementary Material for more information on comparing values of the tensor components to literature results. The relative amplitudes of the tensor components differ by more than their uncertainties for all samples and excitation energies except for GaAs at eV, where the anisotropy was only weakly constrained. The relative amplitudes of also differ by more than their uncertainties for all samples and excitation energies except for GaAs at eV again, and also for GaP at eV, and Si at and 0.95 eV. The relative amplitudes differ by more than their uncertainties for GaAs at , 0.95, 1.13, and 1.24 eV, for GaP at eV, and for Si at and 0.95 eV. Thus Kleinman symmetry is broken, a feature not discerned in the result from Dvorak, et al. for GaAs at 1.31 eV.Dvorak et al. (1994) Indeed, Kleinman symmetry should be broken in the presence of dispersion in .Kleinman (1962); Boyd (2008)
V Conclusions
We characterized the tensor spectrum by measuring the degenerate two-photon absorption coefficient and its anisotropy using a 2-beam PPMS experiment for bulk 110-cut GaAs, GaP, and Si at 300 K over an overall excitation range eV ( nm). The phenomenological 2PA dispersion models we used qualitatitvely agree with our experimental results for GaAs and GaP but there remain significant discrepancies, especially in Si. These could be due to the models lacking consideration of the full band structure of these materials including the effect of anisotropy or additional nonlinear effects. Ab initio calculations in GaAs roughly agree with the observed dispersion. We observed an anomalous asymmetric temporally-broadened nonlinear loss in GaAs and GaP when approached their respective band gaps which could be due to a variety of noninstantaneous nonlinear responses. We also observed a coherent artifact in all three samples near which was likely due to a nonlinear refractive index phase grating. Kleinman symmetry was observed to be broken in all three materials.
Our results motivate further development of ab initio theoriesStryland et al. (1985); Bechstedt, Adolph, and Schmidt (1999); Attaccalite and Grüning (2013); Grüning and Attaccalite (2020); Murayama and Nakayama (1995) such as those which use DFT in the length gauge,Anderson et al. (2015, 2016); Anderson (2016); Anderson and Mendoza (2016) and experimental measurements of the degenerate 2PA spectrum dependence on temperature in Si and GaP which could help elucidate the potential contributions to 2PA below the two-photon direct-gap. Experiments at higher spectral density, using nondegenerate beams, and investigating a larger range of pulse energy- or duration-dependence could identify other nonlinear effects that may be present in these materials at femtosecond timescales. If phonon-mediated transitions are indeed dominant below the two-photon direct-gap in indirect band gap semiconductors, this would motivate further development of ab initio calculations of the tensor spectra which account for such transitions.Noffsinger et al. (2012)
Acknowledgements.
This work was funded by Robert A. Welch Foundation Grant F-1038 and the majority of experimental work performed at the Laboratorio de Óptica Ultrarrápida at Centro de Investigaciones en Óptica, A.C. in León, México. B.S.M. acknowledges support from Consejo Nacional de Ciencia y Tecnología, México, Grant A1-S-9410. The authors would like to thank Enrique Noé-Arias (Centro de Investigaciones en Óptica) for data acquisition program development, Dr. Sean Anderson (Wake Forest University) and Dr. Christopher Reilly (École Polytechnique Fédérale de Lausanne) for assistance with data analysis, and Jack Clifford (University of Texas at Austin) and the Centro de Investigaciones en Óptica Machine Shop for machining assistance and part fabrication.Data Availability
The data that supports the findings of this study are available within the article and its supplementary material, and additional information that support the findings of this study are available from the corresponding author upon reasonable request.
References
References
- Asghari and Krishnamoorthy (2011) M. Asghari and A. Krishnamoorthy, Nat. Photonics 5, 268–270 (2011).
- Yoon et al. (2010) J. Yoon, S. Jo, I. Chun, I. Jung, H. Kim, M. Meitl, E. Menard, X. Li, J. Coleman, U. Paik, and J. Rogers, Nature 465, 329–333 (2010).
- Bett et al. (1999) A. Bett, F. Dimroth, G. Stollwerck, and O. Sulima, Appl. Phys. A 69, 119–129 (1999).
- Wada (1988) O. Wada, Opt. Quantum Electron. 20, 441–474 (1988).
- Pilkuhn and Foster (1966) M. Pilkuhn and L. Foster, IBM J. Res. Dev. 10, 122–129 (1966).
- Wilson et al. (2020) D. Wilson, K. Schneider, S. Hönl, M. Anderson, Y. Baumgartner, L. Czornomaz, T. Kippenberg, and P. Seidler, Nat. Photonics 14, 57–62 (2020).
- Euser and Vos (2005) T. Euser and W. Vos, J. Appl. Phys. 97, 043102 (2005).
- Vlasov et al. (2001) Y. Vlasov, X. Bo, J. Sturm, and D. Norris, Nature 414, 289–293 (2001).
- Rong et al. (2005) H. Rong, A. Liu, R. Jones, O. Cohen, D. Hak, R. Nicolescu, A. Fang, and M. Paniccia, Nature 433, 725–728 (2005).
- Tsang et al. (2002) H. Tsang, C. Wong, T. Liang, I. Day, S. Roberts, A. Harpin, J. Drake, and M. Asghari, Appl. Phys. Lett. 80, 416–418 (2002).
- Dvorak et al. (1994) M. Dvorak, W. Schroeder, D. Andersen, A. Smirl, and B. Wherrett, IEEE J. Quantum Electron. 30, 256–268 (1994).
- Blanco et al. (2000) A. Blanco, E. Chomski, S. Grabtchak, M. Ibisate, S. John, S. Leonard, C. Lopez, F. Meseguer, H. Miguez, J. Mondia, G. Ozin, O. Toader, and H. van Driel, Nature 405, 437–440 (2000).
- Bristow, Rotenberg, and van Driel (2007) A. Bristow, N. Rotenberg, and H. van Driel, Appl. Phys. Lett. 90, 191104 (2007).
- Bassani and Scandolo (1991) F. Bassani and S. Scandolo, Phys. Rev. B 44, 8446–8453 (1991).
- Caspers (1964) P. Caspers, Phys. Rev. A 133, 1249 (1964).
- Hutchings et al. (1992) D. Hutchings, M. Sheik-Bahae, D. Hagan, and E. V. Stryland, Opt. Quantum Electron. 24, 1–30 (1992).
- Kogan (1963) S. Kogan, Sov. Phys. JETP 16, 217 (1963).
- Gibbs (1972) H. Gibbs, Causality and Dispersion Relations (Academic Press, New York, 1972).
- Price (1964) P. Price, Phys. Rev. 130, 1792 (1964).
- Jr. and Jr. (1975) F. R. Jr. and R. G. Jr., Phys. Rev. B 11, 2768 (1975).
- Sheik-Bahae, Hagan, and Styland (1990) M. Sheik-Bahae, D. Hagan, and E. V. Styland, Phys. Rev. Lett. 65, 96–99 (1990).
- Sheik-Bahae et al. (1991) M. Sheik-Bahae, D. Hutchings, D. Hagan, and E. V. Stryland, IEEE J. Quantum Electron. 27, 1296–1309 (1991).
- Toll (1956) J. Toll, Phys. Rev. 104, 1760 (1956).
- Gibbs (1985) H. Gibbs, Optical Bistability: Controlling Light with Light (Academic Press, Orlando, 1985).
- Stegeman and Wright (1990) G. Stegeman and E. Wright, Opt. Quantum Electron. 22, 95–122 (1990).
- Reintjes and McGroddy (1973) J. Reintjes and J. McGroddy, Phys. Rev. Lett. 30, 901–903 (1973).
- Bechtel and Smith (1976) J. Bechtel and W. Smith, Phys. Rev. B 13, 3515–3522 (1976).
- Bosacchi, Bessey, and Jain (1978) B. Bosacchi, J. Bessey, and F. Jain, J. Appl. Phys. 49, 4609–4611 (1978).
- Stryland et al. (1994) E. V. Stryland, M. Sheik-Bahae, T. Xia, C. Wamsley, Z. Wang, A. Said, and D. Hagan, Int. J. Nonlinear Opt. Phys. 3, 489–500 (1994).
- Sheik-Bahae and Stryland (1999) M. Sheik-Bahae and E. V. Stryland, Semiconductors and Semimetals, edited by R. Willardson and E. Weber, Vol. 58 (Academic Press, San Diego, 1999) pp. 257–318.
- Reitze et al. (1990) D. Reitze, T. Zhang, W. Wood, and M. Downer, J. Opt. Soc. Am. B 7, 84–89 (1990).
- Hurlbut et al. (2007) W. Hurlbut, Y. Lee, K. Vodopyanov, P. Kuo, and M. Fejer, Opt. Lett. 32, 668–670 (2007).
- Fishman et al. (2011) D. Fishman, C. Cirloganu, S. Webster, L. Padilha, M. Monroe, D. Hagan, and E. V. Stryland, Nat. Photonics 5, 561–565 (2011).
- Shank, Ippen, and Shapiro (1978) C. Shank, E. Ippen, and S. Shapiro, eds., Picosecond Phenomena (Springer-Verlag, Hilton Head, 1978).
- Benis et al. (2020) S. Benis, C. Cirloganu, N. Cox, T. Ensley, H. Hu, G. Nootz, P. Olszak, L. Padilha, D. Peceli, M. Reichert, S. Webster, M. Woodall, D. Hagan, and E. V. Stryland, Optica 7, 888–899 (2020).
- Cirloganu (2010) C. Cirloganu, Experimental And Theoretical Approaches To Characterization Of Electronic Nonlinearities In Direct–gap Semiconductors, Ph.D. thesis, University of Central Florida (2010).
- Boyd (2008) R. Boyd, Nonlinear Optics, 3rd ed. (Academic Press - Elsevier, Inc., Burlington, 2008).
- Popov, Svirko, and Zheludev (1995) S. Popov, Y. Svirko, and N. Zheludev, Susceptibility Tensors for Nonlinear Optics, Optics and Optoelectronics Series (Institute of Physics Publishing, Bristol, 1995).
- Kleinman (1962) D. Kleinman, Phys. Rev. 126, 1977 (1962).
- Murayama and Nakayama (1995) M. Murayama and T. Nakayama, Phys. Rev. B 52, 4986–4997 (1995).
- Noffsinger et al. (2012) J. Noffsinger, E. Kioupakis, C. V. de Welle, S. Louie, and M. Cohen, Phys. Rev. Lett. 108, 167402–167407 (2012).
- Jones and Reiss (1977) H. Jones and H. Reiss, Phys. Rev. B 16, 2466–2473 (1977).
- Wherrett (1984) B. Wherrett, J. Opt. Soc. Am. B 1, 67–72 (1984).
- Garcia and Kalyanaraman (2006) H. Garcia and R. Kalyanaraman, J. Phys. B 39, 2737–2746 (2006).
- Garcia (2006) H. Garcia, Phys. Rev. B 74, 035212 (2006).
- Aspnes (1972) D. Aspnes, Phys. Rev. B 6, 4648–4659 (1972).
- Moss, Sipe, and van Driel (1987) D. Moss, J. Sipe, and H. van Driel, Phys. Rev. B 36, 9708–9721 (1987).
- Zallen and Paul (1964) R. Zallen and W. Paul, Phys. Rev. 134, A1628–A1641 (1964).
- Diels and Rudolph (1996) J. Diels and W. Rudolph, Ultrafaser Laser Pulse Phenomena: Fundamentals, Techniques, and Applications on a Femtosecond Time Scale (Academic Press, San Diego, 1996).
- Skauli et al. (2003) T. Skauli, P. Kuo, K. Vodopyanov, T. Pinguet, O. Levi, L. Eyres, J. Harris, M. Fejer, B. Gerard, L. Becouarn, and E. Lallier, J. Appl. Phys. 94, 6447–6455 (2003).
- Li (1980) H. Li, J. Phys. Chem. Ref. Data 9, 561–658 (1980).
- Smirl et al. (1988) A. Smirl, G. Valley, K. Bohnert, and J. T.F. Boggess, IEEE J. Quantum Electron. 24, 289–303 (1988).
- Valley and Smirl (1988) G. Valley and A. Smirl, IEEE J. Quantum Electron. 24, 304–310 (1988).
- Hutchings and Stryland (1992) D. Hutchings and E. V. Stryland, J. Opt. Soc. Am. B 9, 2065–2074 (1992).
- Bolger et al. (1993) J. Bolger, A. Kar, B. Wherrett, R. DeSalvo, D. Hutchings, and D. Hagan, Optics Commun. 97, 203 (1993).
- Hannes et al. (2019) W.-R. Hannes, L. Krauß-Kodytek, C. Ruppert, M. Betz, and T. Meier, in Ultrafast Phenomena and Nanophotonics XXIII, Vol. 10916, edited by M. Betz and A. Elezzabi, International Society for Optics and Photonics (SPIE, 2019) pp. 52–62.
- Lebedev et al. (2005) M. Lebedev, O. Misochko, T. Dekorsky, and N. Georgiev, J. Exp. Theor. Phys. 100, 272–282 (2005).
- Palfrey and Heinz (1985) S. Palfrey and T. Heinz, J. Opt. Soc. Am. B 2, 674–679 (1985).
- Penzkofer and Bugayev (1989) A. Penzkofer and A. Bugayev, Opt. Quantum Electron. 21, 283–306 (1989).
- Saissy et al. (1978) A. Saissy, A. Azema, J. Botineau, and F. Gires, Appl. Phys. 15, 99–102 (1978).
- Jayaraman and Lee (1972) S. Jayaraman and C. Lee, Appl. Phys. Lett. 20, 392–395 (1972).
- Kleinman, Miller, and Nordland (1973) D. Kleinman, R. Miller, and W. Nordland, Appl. Phys. Lett. 23, 243–244 (1973).
- Bepko (1975) S. Bepko, Phys. Rev. B 12, 669–672 (1975).
- Lee and Fan (1974) C. Lee and H. Fan, Phys. Rev. B 9, 3502–3516 (1974).
- Oksman et al. (1972) Y. Oksman, A. Semenov, V. Smirnov, and O. Smirnov, Sov. Phys. Semicond. 6, 629 (1972).
- Ralston and Chang (1969) J. Ralston and R. Chang, Appl. Phys. Lett. 15, 164–166 (1969).
- Arsenev et al. (1969) V. Arsenev, V. Dneprovskii, D. Klyshko, and A. Penin, Sov. Phys. JETP 29, 413–415 (1969).
- Basov et al. (1966) N. Basov, A. Grasiuk, I. Zubarev, V. Katulin, and O. Krokhin, Sov. Phys. JETP 23, 366 (1966).
- Grasyuk et al. (1973) A. Grasyuk, I. Zubarev, V. Lobko, Y. Matveets, A. Mironov, and O. Shatberashvili, JETP Lett. 17, 416–418 (1973).
- Zubarev, Mironov, and Mikhailov (1977) I. Zubarev, A. Mironov, and S. Mikhailov, Sov. Phys. Semicond. 11, 239 (1977).
- DeSalvo et al. (1993) R. DeSalvo, M. Sheik-Bahae, A. Said, D. Hagan, and E. V. Stryland, Opt. Lett. 18, 194–196 (1993).
- Said et al. (1992) A. Said, M. Sheik-Bahae, D. Hagan, T. Wei, J. Wang, J. Young, and E. V. Stryland, J. Opt. Soc. Am. B 9, 405–414 (1992).
- Zhi-hui et al. (2015) C. Zhi-hui, X. Si, H. Jun, and G. Bing, Chinese J. Lumin. 36, 969–975 (2015).
- Lautenschlager et al. (1987a) P. Lautenschlager, M. Garriga, S. Logothetidis, and M. Cardona, Phys. Rev. B 35, 9174–9189 (1987a).
- Hoffman and Mott (2012) R. Hoffman and A. Mott, Army Research Laboratory ARL-TR-6157 (2012).
- Dinu, Quochi, and Garcia (2003) M. Dinu, F. Quochi, and H. Garcia, Appl. Phys. Lett. 82, 2954–2956 (2003).
- Lautenschlager et al. (1987b) P. Lautenschlager, M. Garriga, L. Vina, and M. Cardona, Phys. Rev. B 36, 4821–4830 (1987b).
- Stryland et al. (1985) E. V. Stryland, M. Woodall, H. Vanherzeele, and M. Soileau, Opt. Lett. 10, 490 (1985).
- Bechstedt, Adolph, and Schmidt (1999) F. Bechstedt, B. Adolph, and W. Schmidt, Braz. J. Phys. 29, 643–651 (1999).
- Attaccalite and Grüning (2013) C. Attaccalite and M. Grüning, Phys. Rev. B 88, 235113 (2013).
- Grüning and Attaccalite (2020) M. Grüning and C. Attaccalite, “Nonlinear optics from first-principles real-time approaches,” (2020), APS March Meeting.
- Anderson et al. (2015) S. Anderson, N. Tancogne-Dejean, B. Mendoza, and V. Véniard, Phys. Rev. B 91, 075302 (2015).
- Anderson et al. (2016) S. Anderson, N. Tancogne-Dejean, B. Mendoza, and V. Véniard, Phys. Rev. B 93, 235304 (2016).
- Anderson (2016) S. M. Anderson, Theoretical Optical Second-Harmonic Calculations for Surfaces, Ph.D. thesis, Centro de Investigaciones en Óptica, A.C. (2016).
- Anderson and Mendoza (2016) S. Anderson and B. Mendoza, Phys. Rev. B 94, 115314 (2016).