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

Polarization effects in nonlinear interference of down-converted photons

Open Access Open Access

Abstract

We study polarization effects in the nonlinear interference of photons generated via frequency nondegenerate spontaneous parametric-down conversion. Signal and idler photons, which are generated in the visible and infrared (IR) range, respectively, are split into different arms of a nonlinear Michelson interferometer, and the interference pattern for signal photons is detected. Due to the effect of induced coherence, the interference pattern for the signal photons depends on the polarization rotation of idler photons, which are introduced by a birefringent sample. Based on this concept, we realize two new methods of measuring sample retardation in the IR range by using well-developed and inexpensive components for visible light. The methods’ accuracy reaches specifications that are reported for industrial-grade optical elements. The developed IR polarimetry technique is relevant to material research, optical inspection, and quality control.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Interference effects in spontaneous parametric down-conversion (SPDC) have been extensively studied over the last few decades. In addition to the fundamental interest, revealing counterintuitive features of quantum mechanics, they also find practical applications in quantum communication [1], computation [2] and metrology [3–7]. In this work we consider the effect of the nonlinear interference of down-converted photons, also known as “induced coherence” [8,9]. When signal and idler photons generated in two nonlinear crystals are superimposed, the interference is observed in the intensity and coincidence counts [9]. In contrast to the classical case, the interference pattern for signal/idler photons depends on phases and amplitudes of all the three interacting photons: the signal, the idler, and the pump [10,11]. This effect is particularly useful in metrology applications when the sample properties are to be measured in the detection challenging spectral range, for instance in the infrared (IR). When the unknown sample is inserted in the path of idler photons, one can infer its properties in the IR range from the interference pattern for signal photons in the visible range. This concept was recently implemented in several practical applications, including IR imaging [12], IR spectroscopy [13–17] and IR tunable optical coherence tomography (OCT) [18].

Earlier works studied influence of various experimental factors on the nonlinear interference, including the effects of spatial and temporal overlap of SPDC modes [19–21], linear losses and dispersion [8,9,22] introduced in the interferometer. Analysis of relevant polarization effects, however, is less extensive [23–25]. Grayson et al. demonstrated that signal photons acquire the non-local Pancharatnam phase, which was introduced into the path of idler photons by a set of retardation elements [24]. Recently, Lahiri et al. used the nonlinear Mach-Zehnder interferometer and introduced a polarizer into the path of signal photons and an attenuator into the path of idler photons to study the degree of polarization in such a system [25].

In this work, we perform a systematic analysis of polarization effects in the nonlinear interference of down-converted photons. We derive explicit relations between polarization transformations of idler photons and the interference pattern of signal photons. Based on this principle, we propose and experimentally realize two new methods for characterization of retardation properties of a sample in the IR range using well-developed components for visible light.

2. Theory

We consider the nonlinear Michelson interferometer, as shown in Fig. 1(a) [17,18]. SPDC photons (signal and idler) generated at the first pass of the pump through a nonlinear crystal enter the interferometer. The pump, signal and idler photons travel in different paths after being separated by corresponding dichroic mirrors (DM1, DM2). The photons are reflected by the mirrors and recombined at the crystal. The reflected pump generates another pair of down-converted photons, which interferes with the pair traveled in the interferometer.

 figure: Fig. 1

Fig. 1 (a) The nonlinear Michelson interferometer. The pump beam (green) generates SPDC photons (yellow and red), which are separated into different arms by the dichroic mirror DM1. The dichroic mirror DM2 separates signal and pump photons. Mirrors Ms, Mp and Mi reflect all the photons into the crystal, where the pump generates another pair of photons. Then, the interference of signal photons is detected. (b) The beam splitter model which accounts for the double pass through the sample and reflection from the mirror Mi; τi is the amplitude transmission of the beam splitter. A mode ai1 transforms into a mode ai2 by injecting vacuum modes a0 and a0” from open ports of the beam splitter. The mirror Mi inverts the Cartesian coordinate system from the right-handed to the left-handed (x-y).

Download Full Size | PDF

The state vector of SPDC photons generated in the single crystal can be written as [26]:

|ψ=|vac+ηks,kiσs,σiF(ks,ki)aks,σs+aki,σi+|vac
where η<<1 is the SPDC conversion factor; aks,σs+, aki,σi+ are photon creation operators for signal and idler modes with wave vectors ks and ki, polarizations σs and σi, respectively; F(ks,ki) is the two-photon amplitude, which defines spatial and spectral properties of SPDC photons; sub-indices s and i indicate signal and idler modes, respectively [27,28].

Let us consider type-0 phase matching when pump, signal and idler photons have the same linear polarization σs = σi = σ (the theory also applies to type-I and type-II phase matching). According to Eq. (1), for the single spatial and temporal mode, the state vector of SPDC photons created at the first |ψ1 and at the second |ψ2 passes of the pump through the crystal is given by the superposition state [8,9,17,18]:

|Ψ=|ψ1+|ψ2=|vac+ηas1,σ+ai1,σ+|vac+ηeiφpas2,σ+ai2,σ+|vac,
where φp is the phase acquired by the pump in the interferometer [29,30].

When the sample is inserted into the path of idler photons, the state vector in Eq. (2) changes. According to the beam splitter model [31], and assuming that idler modes i1 and i2 are matched, the photon annihilation operator for the idler mode i2 is given by:

ai2,σ=eiφi(τi2ai1,σ+τi1τi2a0,σ+1τi2a0,σ),
where a0,σ and a0,σ denote vacuum fields entering from open ports of the beam splitters, τi is the amplitude transmission coefficient of the beam splitter, φi is the acquired phase. The count rate for signal photons is given by PsΨ|Es()Es(+)|Ψ. Following calculations, described in details in [17,18], we obtain the following expression:
Ps2[1+|τi|2|μ(Δt)|cos(φpφsφi+argτi2+argμ(Δt))],
μ(Δt)=|F(Ω)|2eiΩ(Δt)dΩ
where µ(Δt) is the normalized correlation function of the SPDC, and Δt is the time delay between propagation times of signal and idler photons in the interferometer, Ω is the frequency detuning [32]. The count rate Ps for the signal photons depends on the transmission coefficient of idler photons, and the phases of the signal, the idler, and the pump photons.

Let us now redefine the factor |τi|2, taking into account the polarization properties of the sample. Without the loss of generality, we consider the sample to be a generic retardation waveplate. We use the Jones matrices formalism [33,34] and introduce the corresponding transformation matrix Twp=(τmeiδ/200τmeiδ/2), where δ is the retardation between extraordinary and ordinary waves, and τm is the transmission coefficient of the sample (it accounts for reflection, absorption, and scattering in the sample). Then, the rotation of the Cartesian coordinate system before and after the sample is given by: J=R(θ)TwpR(θ), where R(θ)=(cosθsinθsinθcosθ) is the coordinate rotation matrix, and θ is the orientation of the optical axis of the sample. Figure 1(b) shows the detailed description of the propagation of the idler mode i1. It accounts for a double pass of the photons through the sample and reflection by the mirror Mi. The Jones matrix for this system is given byJ=R(θ)TwpMTwpR(θ)=R(θ)TwpTwpR(θ), where M is the transformation matrix of the mirror Mi. We can then re-write the resulting Jones matrix in the following form:

J=τm2(trr*t*),
where t and r are the complex amplitude transmission and reflection coefficients of the sample, respectively:

t=cosδ+isinδcos2θ,r=isinδsin2θ

We assume that horizontally polarized (along the x-axis) idler photons are created in the first and the second pass of the pump through the nonlinear crystal. The idler photon in mode i1 has an initial polarization vector e=(10) [33,34]. After propagation through the sample, the polarization vector is modified as follows: J(10)=τm2(tr*). The modulus of the amplitude transmission function for the horizontally polarized component is given by:

|τi|2=|τm|2|t|=|τm|2cos2δ+sin2δcos22θ

Since the interference can only be observed for the horizontally polarized component of the idler photon, we can substitute the transmission function |τi|2 in Eq. (4a) by Eq. (7). Thus we obtain the following expression for the count rate of signal photons:

Ps2[1+|τm|2|t||μ(Δt)|cos(φpφsφi+argτm2+argμ(Δt)+argt)]

Let us analyze this expression in more details and show how it can be used for measurement of the sample retardation.

Method 1. Relative phase shift of interference fringes. We set the fast optical axis of the sample parallel to the initial horizontal polarization of idler photons (θ = 0°). Next, we rotate the sample at 90 degrees, so that the slow axis becomes parallel to the polarization of idler photons (θ = 90°). In these two cases, the idler photon experiences refractive indices no, and ne, respectively (we assume no<ne). According to Eq. (8), the relative phase shift between interference fringes of signal photons is proportional to the sample retardation δ:

δ=φi(θ=90°)φi(θ=0°)=2π(neno)λilm,
where lm is the length of the sample accounting for a double pass of idler photons, and λi is the wavelength of idler photons.

Method 2. Visibility of interference fringes. From Eq. (8) we obtain the following expression for the visibility (contrast) of the nonlinear interference:

V=|τm|2|μ(Δt)||t|=|τm|2|μ(Δt)|cos2δ+sin2δcos22θ

The visibility is proportional to the modulus of the transmission function of the waveplate at the wavelength of idler photons. Note, that the similar dependence is valid for signal photons. From Eq. (10) it follows that the maximum and minimum values of the visibility are observed at θ = 0°/90° and θ = 45°, respectively. Then, the retardation can be directly found from the ratio of minimum and maximum values of the visibility:

VminVmax=cos2δcos2δ+sin2δ=|cosδ|

The two methods allow measuring the optical retardation of the sample at the wavelength of idler photons from the measurements of the interference of signal photons. Note that Eqs. (9) and (11) account for the double pass of idler photons through the sample, and δ/2 gives sample retardation in the single pass. Next, we describe the experimental realization of the two methods and discuss the obtained results.

3. Experiment

The experimental setup is shown in Fig. 2. A continuous wave (cw) laser with 532 nm wavelength is used as a pump. The laser is focused by a lens F1 (f = 200 mm) into a periodically poled Lithium Niobate (PPLN) crystal, where SPDC occurs. The phase matching conditions are set to generate the signal photons at λs = 809.2 nm and the idler photons at λi = 1553 nm (ΛPPLN = 7.4 μm poling period; temperature TPPLN = 403 K). Signal and idler beams are separated by a dichroic mirror DM2 into visible and IR arms, respectively, and collimated by lenses F’ (both f = 75 mm). Then pump and signal beams are split by the dichroic mirror DM3 into separate channels. All the three beams are reflected into the crystal by mirrors Ms,p (silver coated) and Mi (gold coated). The reflected pump beam generates another pair of SPDC photons. In the detection part, the pump and idler photons are filtered out by the dichroic mirror (DM1), the notch filter (NF) and the bandpass filter (BP, 809.2 ± 0.6 nm). The signal beam is collimated by the lens F2 (f = 200 mm) and detected by an avalanche photodiode (APD) or a CCD camera preceded by lenses F3 (f = 50 mm) and F4 (f = 100 mm), respectively. The CCD camera is used to facilitate the setup alignment.

 figure: Fig. 2

Fig. 2 The experimental setup. The cw-laser pumps the PPLN crystal, where SPDC occurs. The PPLN is set to generate signal and idler photons in the visible and IR range, respectively. The photons are split by the dichroic mirror DM2 into different arms. Pump and signal photons are separated by the dichroic mirror DM3. All the photons are reflected by the mirrors Ms, Mp, and Mi. Filters DM1, NF, and BP filter the detected signal photons. The interference is detected either by the avalanche photodiode (APD) or by the CCD camera. Mirror Ms is mounted on the translation stage for adjustment of the optical path Δzs. The sample is inserted into the path of idler/ signal photons. Mirrors Mi and Mp are placed on piezo-translators for fine scans of interference fringes.

Download Full Size | PDF

The interference of the signal photons is observed once the interferometer arms are equalized within the coherence length of the SPDC, refer Appendix for the details. The mirror Ms, mounted on a translation stages (step size ~1 μm), is used to equalize the interferometer arms (Δzs). Test samples are mounted into a rotation stage and inserted into the interferometer arms. Fine scans of the interference of signal photons are performed by translating the mirror Mi or Mp (Δzi,p), mounted onto the piezo stages (step size ~2 nm). The interference fringes are measured at different orientations of the sample θ. The interferometer was not actively stabilized and was observed to be stable during the measurements.

4. Results and discussion

4.1 Sample in the path of idler photons

First, we test our method with samples of known retardation, namely quarter- (QWP) and half- (HWP) waveplates designed for 1550 nm. For the initial calibration, we measure the interference fringes without a sample in the IR arm, refer to details in the Appendix. Then, the sample is inserted into the IR arm with its optical axis set parallel to the polarization of the idler photons (θ = 0°). The change of the optical path length is compensated by translation of the mirror Ms. Once the optimal position of the mirror is found, we perform fine scans by translating the mirror Mi in the idler channel using a piezo-stage. Measurements are taken at different orientations of the optical axis of the sample θ.

Figures 3(a) and 3(b) show interference fringes of signal photons measured for the orientation of the sample at θ = 0°, 45°, and 90°. The relative phase shift between patterns at θ = 0° and 90° is equal to δ. As expected from Eq. (9), the shift of the interference pattern is λi/2 and λi for the QWP and HWP, respectively (idler photons travel twice through the sample). This experiment demonstrates the first method for measurement of the sample retardation. Summary of the obtained results is given in Table 1.

 figure: Fig. 3

Fig. 3 The count rate of signal photons at λs = 809.2 nm versus translation of the mirror Mi in the idler channel for (a) QWP at 1550 nm and (b) HWP at 1550 nm. The orientation of the optical axis at 0° (black squares), 45° (red dots) and 90° (blue triangles). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99). The relative phase shift between interference patterns at θ = 0° and 90° is equal to retardation δ. (c) The dependence of the visibility on the sample orientation θ for QWP (blue triangles) and HWP (red circles) at 1550 nm. Points are experimental data, and solid lines are fits by Eq. (10) (R2>0.99). The inset shows zoomed results for QWP at 1550 nm at visibility values close to zero.

Download Full Size | PDF

Tables Icon

Table 1. Results of retardation measurements at 1553 nm by the two methods (idler mirror scan).

Figure 3(c) shows the dependence of the visibility of the interference on the orientation of the sample θ. In accordance with Eq. (10) the interference fringes have maximum visibility at θ = 0° and 90°, while at θ = 45° the visibility reaches its minimum. For the QWP at θ = 45° the visibility is nearly equal to zero (V = 0.005 ± 0.005). Solid lines in Fig. 3(c) correspond to theoretical curves for waveplates designed for 1550 nm. The sample’s absorption data was obtained in independent measurements and is summarized in the Appendix. Thus we realize the second method for measurement of the sample retardation. The measurement results are summarized in Table 1.

Next, we perform measurements of samples with arbitrary retardations. We used HWP and QWP designed for operation at 532 nm. These waveplates operate as unconventional retarders for the probing beam at λi = 1553 nm. Figures 4(a) and 4(b) show interference fringes of signal photons measured for the two samples at orientations θ = 0°, 45°, and 90°. Corresponding retardation values are calculated by Eq. (9), and the results are shown in Table 1. Figure 4(c) shows the dependence of the visibility on the orientation of the two samples. Solid lines are fits with Eq. (10). Sample absorption data was obtained in independent measurements, and it is summarized in the Appendix. The retardation is inferred from the ratio of the minimum and maximum values of the visibility, see Eq. (11). The measurement results are summarized in Table 1.

 figure: Fig. 4

Fig. 4 The count rate of signal photons at λs = 809.2 nm versus translation of the mirror Mi in the idler channel for (a) QWP and (b) HWP at 532 nm inserted in the path of idler photons. The orientation of the optical axis of the sample is at 0° (black squares), 45° (red dots) and 90° (blue triangles). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99). The relative phase shift between interference patterns at θ = 0° and 90° is equal to δ. (c) The dependence of the visibility on the orientation of the sample θ for QWP (blue triangles) and HWP (red circles) designed for 532 nm. The solid curves show fits by Eq. (10) (R2 = 0.98).

Download Full Size | PDF

Then, we investigate the dependence of interference fringes on the translation of the mirror in the pump arm. Figure 5 shows interference fringes obtained for QWP and HWP designed for operation at 1550 nm and 532 nm for different orientation of the optical axis (at θ = 0°, 45°, and 90°). We perform fine scans of the mirror Mp, placed on the piezo-stage. The pump wavelength λp now defines the periodicity of the pattern. Similar to the above procedure, we infer the sample retardation from the phase shift and the visibility ratio. The results are summarized in Table 2. They agree with the ones reported in Table 2.

 figure: Fig. 5

Fig. 5 The count rate measured by translating pump mirror Mp for (a) QWP at 1550 nm, (b) HWP at 1550 nm, (c) QWP at 532 nm, (d) HWP at 532 nm inserted in the idler arm with orientations of the optical axis at 0° (black), 45° (red) and 90° (blue). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99).

Download Full Size | PDF

Tables Icon

Table 2. Results of retardation measurements at 1553 nm by the two methods (pump mirror scan).

4.2 Sample in the path of signal photons

Next, we study polarization transformations by the sample, when it is placed in the signal channel. We use the zero-order QWP designed for 800 nm and perform scans of the mirror in the idler channel. Similar to previous measurements, we observe the phase shift of the interference fringes at different orientations of the sample θ, see Fig. 6(a). Also, we measure the visibility dependence at each orientation of the sample (Method 2), see Fig. 6(b). The retardation is then inferred from the shift of the interference fringes (Method 1) and the ratio of minimum/ maximum visibilities (Method 2). The results are summarized in Table 3.

 figure: Fig. 6

Fig. 6 (a) The count rate measured by translating the mirror Mi in the idler channel when the QWP at 800 nm is inserted in the signal channel. Orientations of the optical axis are at 0° (black squares), 45° (red dots) and 90° (blue triangles). (b) The dependence of the visibility of the interference on the orientation of the sample. Points are experimental data, and solid lines are fits with Eq. (8) in (a) and Eq. (11) in (b) (both R2 > 0.99). The inset shows zoomed visibility values near zero for QWP at 800 nm (yellow rhombuses and line), and for QWP at 1550 nm (blue line) for reference.

Download Full Size | PDF

Tables Icon

Table 3. Results of retardation measurements at 800 nm by the two methods (idler mirror scan).

4.3 Discussion

Based on the obtained experimental data we are now able to compare the two methods for characterization of samples retardation. The “phase shift” method (Method 1) is fast, as we infer the retardation from two sets of fringes at different orientations of the sample. However, the interferometer should be well stabilized during the measurements to minimize drift of the fringes due to thermal and mechanical fluctuations, see Appendix A3 for the analysis of introduced uncertainties. The best experimental accuracy in determination of the retardation δ is about ± 0.006π.

In contrast, the “visibility method” (Method 2) is more tolerable to the phase drifts, as they do not strongly affect the visibility. However, the measurement takes longer time than the “phase shift” method. The accuracy of the method is defined by the contrast between minimum and maximum visibility values attained in the experiment. In this method, the best achieved experimental accuracy is at the order of ± 0.002π, which also accounts for the accuracy in setting up the sample orientation. The accuracy of the two methods reaches specifications reported for industrial-grade optical elements

As our results show, the period of the interference fringes is defined by the scanning configuration: it is given either by λp or by λi, depending on which mirror is scanned. The configuration with the scan of the mirror in the pump beam may represent a certain advantage when the translation range of the piezo-stage is limited. We also note that when the sample is inserted in the path of the signal photon, the retardation can be measured at the visible wavelength with the same configuration. Moreover, our method allows us to distinguish fast and slow axes of the sample, see Appendix A4.

5. Conclusions

In conclusion, we performed a detailed study of polarization effects in the nonlinear interferometer. We showed that the change of the polarization of idler photons affects both the visibility and phase of the interference fringes of signal photons. This effect allows us inferring the sample retardation at the frequency of idler photons from (a) shift of the interference fringes and (b) from the visibility ratio. The suggested “phase shift” and the “visibility” methods allow characterization of the sample retardation with the accuracy up to Δδ = ± 0.002π, which meets the requirements of the optical industry.

In comparison with conventional polarimetry systems, which require an IR light source and a detector, our technique has a benefit for wavelength tunability and the use of visible range source and detector. An extension of the operating wavelengths of the presented methods can be easily achieved by choice of the periodic poling of the crystals and/or by tuning crystal temperature. Operation within the range of 1.5–4.3 um was shown in our earlier works with Lithium Niobate crystals [15–18]. The retardation measurement for a given sample takes about 4 minutes using the “phase shift” method and about 20 minutes using the “visibility method”. These measurement times are comparable with corresponding values for the conventional techniques [35].

Our technique can be readily extended to the IR polarimetry of arbitrary waveplates and samples with optical activity, such as Faraday rotators and chiral media. Furthermore, it is possible to measure the spatial uniformity of the retardation properties of a sample using an additional lens in the path of probing photons [18].

Also, active control over polarization of probing photons allows enhancing the signal-to-noise ratio by compensating polarization changes in the realistic samples. This idea forms the basis for polarization-sensitive optical coherence tomography (PS-OCT) [36] and polarization-sensitive quantum optical coherence tomography (PS-QOCT) [37], which have already been set forth as high-contrast methods for birefringence measurements of layered samples. Consequently, our technique can be used for the PS-OCT development, extending the conventional methods to the mid-and far-IR range.

Appendix

A.1 Visibility of interference

The state vector of type-0 SPDC the in Eq. (2) is given by:

|Ψ=|ψ1+|ψ2=|vac+ei(φs+φi)|Hs1|Hi1+eiφp|Hs2|Hi2,
where |H is the single-photon Fock state with horizontal polarization. When the waveplate is introduced into the path of signal photons the state vector in Eq. (12) is given by:

|Ψ=|vac+ei(φs+φi)J|Hs1|Hi1+eiφp|Hs2|Hi2,

Then, the polarization state vector of the signal photons |Hs1 is transformed as J(10)s1=(tsrs*), where index s denote signal photons. Then, the state vector of the SPDC photons in Eq. (13) is given by:

|Ψ=ei(φs+φi)(ts|Hs1rs*|Vs1)|Hi1+eiφp|Hs2|Hi2

Assuming that idler photon modes are aligned i1=i2=i, the count rate of the signal photons is given by:

PsΨ|as+as|Ψ=|ei(φs+φi)ts+eiφp|2+|ei(φs+φi)rs*|2==2+2|ts|cos(φs+φiφp+argts)

The visibility of the interference fringes obtained from Eq. (15) is proportional to the modulus of the transmission function of the waveplate at the wavelength of signal photons V=|ts|.

A2. Alignment of interferometer

Initially, the interferometer is balanced, and the interference pattern is observed around the zero position of the translation stage of Ms. The reference interference pattern is shown by black squares in Fig. 7. Once the sample is inserted in the path of idler photons, the mirror Ms has to be moved to compensate for the introduced optical delay. Figure 7 also shows interference pattern after introducing samples into the path of idler photons.

 figure: Fig. 7

Fig. 7 The shift of the interference pattern after the introduction of the sample in the path of idler photons. Data for HWP at 1550 nm, QWP at 1550 nm, HWP at 532 nm and QWP at 532 nm with the orientation of the optical axis at θ = 0°.

Download Full Size | PDF

For the HWP at 1550 nm, QWP at 1550 nm, HWP at 532 nm phase shifts are approximately the same, as they are all purchased from the same supplier (Thorlabs) and have close values of the optical thicknesses. The QWP at 532 nm is purchased from another vendor (DayOptics) and the optical thickness is somewhat smaller.

Figure 8 shows the shift of the reference interference pattern due to the introduction of the QWP at 800 nm into the path of signal and idler photons. As the waveplate is designed for 800 nm, its absorption for signal photons is smaller than for idler photons.

 figure: Fig. 8

Fig. 8 The shift of the interference pattern due to the change of the optical path for the signal and idler photons after insertion of the HWP at 800 nm with the orientation of the optical axis at θ = 0°.

Download Full Size | PDF

By measuring the interference at the position of maximum visibility of the interference pattern, it is possible to define the transmission function |τm|2, see Table 4. This data is used to plot theoretical curves for the visibility function in Figs. 3, 4, and 6. The experimental data in Table 4 is in a good agreement with reference measurements, obtained with a conventional IR spectrophotometer (Shimadzu UV3600). The accuracy of the transmission coefficient measurement is ±0.012, which is higher than shown in [17] due to a nearly 5-times increase in the visibility of the interference pattern.

Tables Icon

Table 4. Transmission coefficient |τm|2 for different samples.

A3. Uncertainties of measurements

A3.1 “Phase shift” method

The accuracy of measurements in the “phase shift” method is determined by temperature fluctuations of the PPLN crystal, mechanical perturbations in the interferometer, and by the uncertainty of the position of the scanning piezo stage.

The temperature of the PPLN crystal is controlled with an accuracy of ±0.01K, which results in the uncertainty of generated wavelengths of 0.22 nm for the idler photons at 1550 nm and 0.1 nm for the signal photons at 810 nm. Therefore, the uncertainty of the phase estimation due to temperature fluctuation is about 3.2×10−4π.

To study the effect of mechanical perturbations, we perform measurements of the phase fluctuations at a fixed temperature of the crystal, see Fig. 9. Considering that the signal, shown in Fig. 9, is at the slope of the interference fringe, the drift of the phase is proportional to the standard deviation of the signal detected by the APD Δδ=σAA0.0056π, where A is the amplitude of the interference fringe. The standard deviation is taken during one measurement cycle of 50 seconds.

 figure: Fig. 9

Fig. 9 Measurement of the phase drift in the interferometer.

Download Full Size | PDF

During the experiment, we first measure the interference fringe with the orientation of the waveplate at θ=0°. We take the first period of the interference fringe (see region “1” in Fig. 10) as a reference point for future measurements at orientations θ=45° and 90°. Next, we start to scan the interference fringe again at θ=0°, and change the orientation of the waveplate to θ=45°/90° (region “2” in Fig. 10). By matching the initial point of the interference fringe at region “1”, we are able to compensate for the long term phase drift in the interferometer.

 figure: Fig. 10

Fig. 10 The experimental procedure of the “phase shift” measurements with the QWP designed for 532 nm wavelength. Region “1” corresponds to the reference point, where each scan starts from the θ = 0° orientation of the waveplate; “2” corresponds to the region, where the orientation of the waveplate is changed; “3” is the region where the retardation of the waveplate is measured.

Download Full Size | PDF

The uncertainty of the position of the scanning piezo stage is determined by the fitting of the obtained interference fringes by a cosine function (see Eq. (8)). Fitting of the interference fringes is done using Origin software with a built-in least-square method. The phase of the interference fringes is determined as:

δ=2πλ2x,
where x is the position of the piezo stage, additional factor 2 takes into account the double pass of the photons in the interferometer. The phase uncertainty is determined as:

Δδ2πλ2Δx=Δxλ/4π

For idler photons, the accuracy is Δδ1.3389π=0.003π, see Figs. 3, 4, and 6. And for pump photons, the accuracy is Δδ0.45133π=0.003π, see Fig. 5.

From the above analysis, one concludes that the uncertainty due to temperature fluctuations can be neglected. Finally, taking into account the phase drift in the interferometer and the fitting errors, the total uncertainty of the “phase shift” method is approximately 0.006 π.

A3.2 “Visibility” method

In the “visibility” method the accuracy is determined by the fluctuation of the photocount rate in the APD detector. According to Eq. (11) the retardation δ is determined as:

δ=arccos(VminVmin).

Therefore, the uncertainty of the retardation measurement is determined as:

Δδ=ΔVmin2Vmax2Vmin2+ΔVmax2Vmax2Vmin2Vmin2Vmax2,
where ΔVmin and ΔVmax are uncertainties for the minimum and maximum visibilities. These uncertainties are calculated as fitting errors in the Origin software.

For example, the error for the QWP@1550 nm is given by:

Δδ=0.0044420.7484620.005112+0.0044420.7484620.0051120.0051120.7484620.002π

The error for the HWP@1550 nm is:

Δδ=0.0058420.7613220.754542+0.0058620.7613220.7545420.7545420.7613220.026π

We can see that the accuracy of this method depends on the ratio between the maximum and minimum values of the visibilities. The higher is the visibility ratio, the more accurate are the measurements.

Comparing these two methods of the retardation measurement, we can conclude that the “phase shift” measurements is faster, but it has a slightly worse accuracy. At the same time, the “visibility” method takes a longer time, but it yields a higher accuracy.

A4. Determination of the fast and slow axis of the waveplates

Our experimental configuration allows to distinguish slow and fast axis of a sample by applying the “phase shift” method. Indeed, any change in the optical path length in the interferometer arms results in the shift of interference fringes (see Fig. 7). We can determine the direction of the shift of the fringes which correspond to the increase or decrease of the optical path.

In Fig. 7, when we place waveplate in the path of the idler photons, the optical path for the signal photons should be increased. To compensate for the additional optical path we translate the stage with mirror Ms to the positive direction of z (we can also consider moving mirror Mi to the negative direction –z).

For the arbitrary sample, we need to determine the positions of the highest visibility of the interference fringes. The highest visibility value corresponds to the fast or to the slow axis of the waveplate. Next, the waveplate is rotated by 90 degrees. If the sample is rotated in the direction to its slow axis (none), it results in an increase of the optical path length and a shift of interference fringes to a specific direction. And vice versa, moving from the fast to the slow axis (neno) results in a shift of the interference fringes in the opposite direction.

We can test this statement based on the results shown in Figs. 3, 4, 5, 6, and 10. We say that angle at θ=0° corresponds to the fast axis of the waveplate, which corresponds to the ordinary refractive index no (no <ne). When we change the orientation of the waveplate (θ=45° or θ=90°), the refractive index starts to increase (none), which leads to the shift of the interference fringes to the negative direction.

Therefore, by observing the direction of the shift of the interference fringes with the rotation of the sample, we can unambiguously determine its fast and slow axis.

Funding

Quantum Technology for Engineering program (QTE) of A*STAR (project # A1685b0005); Singapore International Graduate Award (SINGA) fellowship.

Acknowledgments

We acknowledge fruitful discussions with Sergei Kulik, Rainer Dumke, and Berthold-Georg Englert.

References

1. D. Bouwmeester, J.-W. Pan, K. Mattle, M. Eibl, H. Weinfurter, and A. Zeilinger, “Experimental quantum teleportation,” Nature 390(6660), 575–579 (1997). [CrossRef]  

2. S. Gasparoni, J.-W. Pan, P. Walther, T. Rudolph, and A. Zeilinger, “Realization of a photonic controlled-NOT gate sufficient for quantum computation,” Phys. Rev. Lett. 93(2), 020504 (2004). [CrossRef]   [PubMed]  

3. V. Giovannetti, S. Lloyd, and L. Maccone, “Quantum-enhanced measurements: beating the standard quantum limit,” Science 306(5700), 1330–1336 (2004). [CrossRef]   [PubMed]  

4. C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett. 59(18), 2044–2046 (1987). [CrossRef]   [PubMed]  

5. D. A. Kalashnikov, E. V. Melik-Gaykazyan, A. A. Kalachev, Y. F. Yu, A. I. Kuznetsov, and L. A. Krivitsky, “Quantum interference in the presence of a resonant medium,” Sci. Rep. 7(1), 11444 (2017). [CrossRef]   [PubMed]  

6. D. Branning, A. L. Migdall, and A. V. Sergienko, “Simultaneous measurement of group and phase delay between two photons,” Phys. Rev. A 62(6), 063808 (2000). [CrossRef]  

7. T. Ono, R. Okamoto, and S. Takeuchi, “An entanglement-enhanced microscope,” Nat. Commun. 4(1), 2426 (2013). [CrossRef]   [PubMed]  

8. X. Y. Zou, L. J. Wang, and L. Mandel, “Induced coherence and indistinguishability in optical interference,” Phys. Rev. Lett. 67(3), 318–321 (1991). [CrossRef]   [PubMed]  

9. L. J. Wang, X. Y. Zou, and L. Mandel, “Induced coherence without induced emission,” Phys. Rev. A 44(7), 4614–4622 (1991). [CrossRef]   [PubMed]  

10. H. M. Wiseman and K. Molmer, “Induced coherence with and without induced emission,” Phys. Lett. A 270(5), 245–248 (2000). [CrossRef]  

11. M. V. Chekhova and Z. Y. Ou, “Nonlinear interferometers in quantum optics,” Adv. Opt. Photonics 8(1), 104–155 (2016). [CrossRef]  

12. G. B. Lemos, V. Borish, G. D. Cole, S. Ramelow, R. Lapkiewicz, and A. Zeilinger, “Quantum imaging with undetected photons,” Nature 512(7515), 409–412 (2014). [CrossRef]   [PubMed]  

13. A. V. Burlakov, S. P. Kulik, A. N. Penin, and M. V. Chekhova, “Three-photon interference: spectroscopy of linear and nonlinear media,” Sov. Phys. JETP 86(6), 1090–1097 (1998). [CrossRef]  

14. S. P. Kulik, G. A. Maslennikov, S. P. Merkulova, A. N. Penin, L. K. Radchenko, and V. N. Krasheninnikov, “Two-photon interference in the presence of absorption,” Sov. Phys. JETP 98(1), 31–38 (2004). [CrossRef]  

15. D. A. Kalashnikov, A. V. Paterova, S. P. Kulik, and L. A. Krivitsky, “Infrared spectroscopy with visible light,” Nat. Photonics 10(2), 98–101 (2016). [CrossRef]  

16. A. Paterova, S. Lung, D. A. Kalashnikov, and L. A. Krivitsky, “Nonlinear infrared spectroscopy free from spectral selection,” Sci. Rep. 7(1), 42608 (2017). [CrossRef]   [PubMed]  

17. A. Paterova, H. Yang, Ch. An, D. Kalashnikov, and L. Krivitsky, “Measurement of infrared optical constants with visible photons,” New J. Phys. 20(4), 043015 (2018). [CrossRef]  

18. A. V. Paterova, H. Yang, Ch. An, D. A. Kalashnikov, and L. A. Krivitsky, “Tunable optical coherence tomography in the infrared range using visible photon,” Quantum Science and Technology 3(2), 025008 (2018). [CrossRef]  

19. T. P. Grayson and G. A. Barbosa, “Spatial properties of spontaneous parametric down-conversion and their effect on induced coherence without induced emission,” Phys. Rev. A 49(4), 2948–2961 (1994). [CrossRef]   [PubMed]  

20. M. Lahiri, A. Hochrainer, R. Lapkiewicz, G. B. Lemos, and A. Zeilinger, “Twin-photon correlations in single-photon interference,” Phys. Rev. A (Coll. Park) 96(1), 013822 (2017). [CrossRef]  

21. X. Y. Zou, T. Grayson, G. A. Barbosa, and L. Mandel, “Control of visibility in the interference of signal photons by delays imposed on the idler photons,” Phys. Rev. A 47(3), 2293–2295 (1993). [CrossRef]   [PubMed]  

22. M. Lahiri, R. Lapkiewicz, G. B. Lemos, and A. Zeilinger, “Theory of quantum imaging with undetected photons,” Phys. Rev. A 92(1), 013832 (2015). [CrossRef]  

23. L. C. Ryff, “Interference, distinguishability, and apparent contradiction in an experiment on induced coherence,” Phys. Rev. A 52(4), 2591–2596 (1995). [CrossRef]   [PubMed]  

24. T. P. Grayson, J. R. Torgerson, and G. A. Barbosa, “Observation of a nonlocal Pancharatnam phase shift in the process of induced coherence without induced emission,” Phys. Rev. A 49(1), 626–628 (1994). [CrossRef]   [PubMed]  

25. M. Lahiri, A. Hochrainer, R. Lapkiewicz, G. B. Lemos, and A. Zeilinger, “Partial polarization by quantum distinguishability,” Phys. Rev. A (Coll. Park) 95(3), 033816 (2017). [CrossRef]  

26. S. P. Walborn, C. H. Monken, S. Padua, and P. H. Souto Ribeiro, “Spatial correlations in parametric down-conversion,” Phys. Rep. 495(4–5), 87–139 (2010). [CrossRef]  

27. S. P. Walborn, A. N. de Oliveira, S. Pádua, and C. H. Monken, “Multimode Hong-Ou-mandel interference,” Phys. Rev. Lett. 90(14), 143601 (2003). [CrossRef]   [PubMed]  

28. J.-Ch. Lee and Y.-H. Kim, “Spatial and spectral properties of entangled photons from spontaneous parametric down conversion with a focused pump,” Opt. Commun. 366, 442–450 (2016). [CrossRef]  

29. T. J. Herzog, J. G. Rarity, H. Weinfurter, and A. Zeilinger, “Frustrated two-photon creation via interference,” Phys. Rev. Lett. 72(5), 629–632 (1994). [CrossRef]   [PubMed]  

30. A. Heuer, S. Raabe, and R. Menzel, “Phase memory across two single photon interferometers including wavelength conversion,” Phys. Rev. A 90(4), 045803 (2014). [CrossRef]  

31. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995).

32. S.-Y. Baek and Y.-H. Kim, “Spectral properties of entangled photons generated via type-I frequency-nondegenerate spontaneous parametric down-conversion,” Phys. Rev. A 80(3), 033814 (2009). [CrossRef]  

33. R. M. A. Azzam and N. M. Bashara, Ellipsometry and Polarized Light (North-Holland, 1987).

34. D. H. Goldstein, Polarized Light (CRC, 2011).

35. P. A. Williams, A. H. Rose, and C. M. Wang, “Rotating-polarizer polarimeter for accurate retardance measurement,” Appl. Opt. 36(25), 6466–6472 (1997). [CrossRef]   [PubMed]  

36. J. F. de Boer, C. K. Hitzenberger, and Y. Yasuno, “Polarization sensitive optical coherence tomography - a review [Invited],” Biomed. Opt. Express 8(3), 1838–1873 (2017). [CrossRef]   [PubMed]  

37. M. C. Booth, G. Di Giuseppe, B. E. A. Saleh, A. V. Sergienko, and M. C. Teich, “Polarization-sensitive quantum-optical coherence tomography,” Phys. Rev. A 69(4), 043815 (2004). [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 (10)

Fig. 1
Fig. 1 (a) The nonlinear Michelson interferometer. The pump beam (green) generates SPDC photons (yellow and red), which are separated into different arms by the dichroic mirror DM1. The dichroic mirror DM2 separates signal and pump photons. Mirrors Ms, Mp and Mi reflect all the photons into the crystal, where the pump generates another pair of photons. Then, the interference of signal photons is detected. (b) The beam splitter model which accounts for the double pass through the sample and reflection from the mirror Mi; τi is the amplitude transmission of the beam splitter. A mode ai1 transforms into a mode ai2 by injecting vacuum modes a0 and a0” from open ports of the beam splitter. The mirror Mi inverts the Cartesian coordinate system from the right-handed to the left-handed (x-y).
Fig. 2
Fig. 2 The experimental setup. The cw-laser pumps the PPLN crystal, where SPDC occurs. The PPLN is set to generate signal and idler photons in the visible and IR range, respectively. The photons are split by the dichroic mirror DM2 into different arms. Pump and signal photons are separated by the dichroic mirror DM3. All the photons are reflected by the mirrors Ms, Mp, and Mi. Filters DM1, NF, and BP filter the detected signal photons. The interference is detected either by the avalanche photodiode (APD) or by the CCD camera. Mirror Ms is mounted on the translation stage for adjustment of the optical path Δzs. The sample is inserted into the path of idler/ signal photons. Mirrors Mi and Mp are placed on piezo-translators for fine scans of interference fringes.
Fig. 3
Fig. 3 The count rate of signal photons at λs = 809.2 nm versus translation of the mirror Mi in the idler channel for (a) QWP at 1550 nm and (b) HWP at 1550 nm. The orientation of the optical axis at 0° (black squares), 45° (red dots) and 90° (blue triangles). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99). The relative phase shift between interference patterns at θ = 0° and 90° is equal to retardation δ. (c) The dependence of the visibility on the sample orientation θ for QWP (blue triangles) and HWP (red circles) at 1550 nm. Points are experimental data, and solid lines are fits by Eq. (10) (R2>0.99). The inset shows zoomed results for QWP at 1550 nm at visibility values close to zero.
Fig. 4
Fig. 4 The count rate of signal photons at λs = 809.2 nm versus translation of the mirror Mi in the idler channel for (a) QWP and (b) HWP at 532 nm inserted in the path of idler photons. The orientation of the optical axis of the sample is at 0° (black squares), 45° (red dots) and 90° (blue triangles). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99). The relative phase shift between interference patterns at θ = 0° and 90° is equal to δ. (c) The dependence of the visibility on the orientation of the sample θ for QWP (blue triangles) and HWP (red circles) designed for 532 nm. The solid curves show fits by Eq. (10) (R2 = 0.98).
Fig. 5
Fig. 5 The count rate measured by translating pump mirror Mp for (a) QWP at 1550 nm, (b) HWP at 1550 nm, (c) QWP at 532 nm, (d) HWP at 532 nm inserted in the idler arm with orientations of the optical axis at 0° (black), 45° (red) and 90° (blue). Points are experimental data, and solid lines are fits by Eq. (8) (R2>0.99).
Fig. 6
Fig. 6 (a) The count rate measured by translating the mirror Mi in the idler channel when the QWP at 800 nm is inserted in the signal channel. Orientations of the optical axis are at 0° (black squares), 45° (red dots) and 90° (blue triangles). (b) The dependence of the visibility of the interference on the orientation of the sample. Points are experimental data, and solid lines are fits with Eq. (8) in (a) and Eq. (11) in (b) (both R2 > 0.99). The inset shows zoomed visibility values near zero for QWP at 800 nm (yellow rhombuses and line), and for QWP at 1550 nm (blue line) for reference.
Fig. 7
Fig. 7 The shift of the interference pattern after the introduction of the sample in the path of idler photons. Data for HWP at 1550 nm, QWP at 1550 nm, HWP at 532 nm and QWP at 532 nm with the orientation of the optical axis at θ = 0°.
Fig. 8
Fig. 8 The shift of the interference pattern due to the change of the optical path for the signal and idler photons after insertion of the HWP at 800 nm with the orientation of the optical axis at θ = 0°.
Fig. 9
Fig. 9 Measurement of the phase drift in the interferometer.
Fig. 10
Fig. 10 The experimental procedure of the “phase shift” measurements with the QWP designed for 532 nm wavelength. Region “1” corresponds to the reference point, where each scan starts from the θ = 0° orientation of the waveplate; “2” corresponds to the region, where the orientation of the waveplate is changed; “3” is the region where the retardation of the waveplate is measured.

Tables (4)

Tables Icon

Table 1 Results of retardation measurements at 1553 nm by the two methods (idler mirror scan).

Tables Icon

Table 2 Results of retardation measurements at 1553 nm by the two methods (pump mirror scan).

Tables Icon

Table 3 Results of retardation measurements at 800 nm by the two methods (idler mirror scan).

Tables Icon

Table 4 Transmission coefficient |τm|2 for different samples.

Equations (22)

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

|ψ =| vac+η k s , k i σ s , σ i F( k s , k i ) a k s , σ s + a k i , σ i + | vac
|Ψ= |ψ 1 + |ψ 2 =| vac+η a s1,σ + a i1,σ + | vac+η e i φ p a s2,σ + a i2,σ + | vac,
a i2,σ = e i φ i ( τ i 2 a i1,σ + τ i 1 τ i 2 a 0 ,σ + 1 τ i 2 a 0 ,σ ),
P s 2[ 1+ | τ i | 2 | μ( Δt ) |cos( φ p φ s φ i +arg τ i 2 +argμ( Δt ) ) ],
μ( Δt )= | F( Ω ) | 2 e iΩ( Δt ) dΩ
J = τ m 2 ( t r r * t * ),
t=cosδ+isinδcos2θ, r=isinδsin2θ
| τ i | 2 = | τ m | 2 | t |= | τ m | 2 cos 2 δ+ sin 2 δ cos 2 2θ
P s 2[ 1+ | τ m | 2 | t || μ( Δt ) |cos( φ p φ s φ i +arg τ m 2 +argμ( Δt )+argt ) ]
δ= φ i ( θ=90° ) φ i ( θ=0° )= 2π( n e n o ) λ i l m ,
V= | τ m | 2 | μ( Δt ) || t |= | τ m | 2 | μ( Δt ) | cos 2 δ+ sin 2 δ cos 2 2θ
V min V max = cos 2 δ cos 2 δ+ sin 2 δ =| cosδ |
|Ψ= |ψ 1 + |ψ 2 =| vac+ e i( φ s + φ i ) |H s1 |H i1 + e i φ p |H s2 |H i2 ,
|Ψ=| vac+ e i( φ s + φ i ) J |H s1 |H i1 + e i φ p |H s2 |H i2 ,
|Ψ= e i( φ s + φ i ) ( t s |H s1 r s * |V s1 ) |H i1 + e i φ p |H s2 |H i2
P s Ψ| a s + a s |Ψ= | e i( φ s + φ i ) t s + e i φ p | 2 + | e i( φ s + φ i ) r s * | 2 = =2+2| t s |cos( φ s + φ i φ p +arg t s )
δ= 2π λ 2x,
Δδ 2π λ 2Δx= Δx λ/4 π
δ=arccos( V min V min ).
Δδ= Δ V min 2 V max 2 V min 2 + Δ V max 2 V max 2 V min 2 V min 2 V max 2 ,
Δδ= 0.00444 2 0.74846 2 0.00511 2 + 0.00444 2 0.74846 2 0.00511 2 0.00511 2 0.74846 2 0.002π
Δδ= 0.00584 2 0.76132 2 0.75454 2 + 0.00586 2 0.76132 2 0.75454 2 0.75454 2 0.76132 2 0.026π
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.