<<

656 J. Opt. Soc. Am. B/Vol. 19, No. 4/April 2002 Abouraddy et al.

Entangled-photon ellipsometry

Ayman F. Abouraddy, Kimani C. Toussaint, Jr., Alexander V. Sergienko, Bahaa E. A. Saleh, and Malvin C. Teich Quantum Imaging Laboratory, Departments of Electrical & Computer Engineering and Physics, Boston University, Boston, Massachusetts 02215-2421

Received May 25, 2001; revised manuscript received October 10, 2001 Performing reliable measurements in optical metrology, such as those needed in ellipsometry, requires a cali- brated source and detector, or a well-characterized reference sample. We present a novel interferometric tech- nique to perform reliable ellipsometric measurements. This technique relies on the use of a nonclassical op- tical source, namely, polarization-entangled twin photons generated by spontaneous parametric downconversion from a nonlinear crystal, in conjunction with a coincidence-detection scheme. Ellipsometric measurements acquired with this scheme are absolute, i.e., they require neither source nor detector calibra- tion, nor do they require a reference. © 2002 Optical Society of America OCIS codes: 120.2130, 270.0270, 190.0190, 120.3940, 000.1600, 350.4600.

1. INTRODUCTION the use of this nonclassical light source to the field of el- lipsometry. A question that arises frequently in metrology is the fol- In Section 3 we propose two different experimental lowing: How does one measure reliably the reflection or implementations of twin-photon ellipsometry. The first transmission coefficient of an unknown sample? The makes use of a twin-photon interferometer that has been outcome of such a measurement depends on the reliability previously used for testing the foundations of quantum of both the source and the detector used to carry out the mechanics. The second technique makes direct use of measurements. If they are each absolutely calibrated, polarization-entangled photon pairs emitted through such measurements would be trivial. Since such ideal SPDC. This approach effectively comprises an interfero- conditions are never met in practice, and since high- metric ellipsometer, although none of the optical elements precision measurements are often required, a myriad of usually associated with constructing an interferometer experimental techniques, such as null and interferometric are utilized. Instead, polarization entanglement itself is approaches, have been developed to circumvent the im- harnessed to perform interferometry and to achieve ideal perfections of the devices involved in these measure- ellipsometry. The inherent limitations of the first tech- ments. nique are eliminated in the second. One optical metrology setting in which high-precision measurements are a necessity is ellipsometry,1–6 in which 2. IDEAL ELLIPSOMETRY the polarization of light is used to study thin films on sub- In an ideal ellipsometer, the light emitted from a reliable strates, a technique established more than a hundred optical source is directed into an unknown optical system 1,4,5 years ago. Ellipsometers have proven to be an impor- (which may simply be an unknown sample that reflects tant metrological tool in many arenas ranging from the the impinging light) and thence into a reliable detector. semiconductor industry to biomedical applications. To The practitioner keeps track of the emitted and detected carry out ideal ellipsometry, one needs a perfectly cali- radiation, and from this bookkeeping (s)he can infer infor- brated source and detector. Various approaches, such as mation about the optical system. This device may be null and interferometric techniques, have been commonly used as an ellipsometer if the source can emit light in any used in ellipsometers2,6 to approach this ideal. Section 2 specified state of polarization. The sample is character- of this paper describes the basic requirements for ideal el- ized by two parameters: ␺ and ⌬. The quantity ␺ is re- lipsometry and reviews some of the more common tech- lated to the magnitude of the ratio of the sample’s eigen- niques that have been used in conjunction with available polarization complex reflection coefficients, ˜r1 and ˜r2 , ␺ ϭ ͉ ͉ ⌬ detectors and sources. through tan ˜r1 /˜r2 ; is the phase shift between In this paper we propose a novel technique for obtain- them.2 ing reliable ellipsometric measurements based on the use Because of the high accuracy required in measuring of twin photons produced by the process of spontaneous these parameters, an ideal ellipsometric measurement optical parametric downconversion7–11 (SPDC). This would require absolute calibration of both the source and source has been used effectively in studies of the founda- the detector. Since this is not attainable in practical set- tions of quantum mechanics12,13 and in applications in tings, ellipsometry makes use of a myriad of experimental quantum metrology14–17; process- techniques developed to circumvent the imperfections of ing, such as quantum cryptography18–20; quantum the involved devices. The most common techniques are teleportation21,22; and quantum imaging.23–25 We extend null and interferometric ellipsometry.

0740-3224/2002/040656-07$15.00 © 2002 Optical Society of America Abouraddy et al. Vol. 19, No. 4/April 2002/J. Opt. Soc. Am. B 657

In the traditional null ellipsometer,2 depicted in Fig. 1, 3. TWIN-PHOTON ELLIPSOMETRY the sample is illuminated with a beam of light that can be All classical optical sources (including ideal amplitude- prepared in any state of polarization. The reflected light, stabilized lasers) suffer from unavoidable quantum fluc- which is generally elliptically polarized, is then analyzed. tuations even if all other extraneous noise sources are re- The polarization of the incident beam is adjusted to com- moved. Fluctuations in the photon number can only be pensate for the change in the relative amplitude and eliminated by constructing a source that emits nonover- phase, introduced by the sample, between the two eigen- lapping wave packets, each of which contains a fixed pho- polarizations; thus the resulting reflected beam is linearly ton number. Such sources have been investigated, and polarized. If passed through an orthogonal linear polar- indeed sub-Poisson light sources have been demon- izer, this linearly polarized beam will yield a null (zero) strated.26–28 measurement at the optical detector. The null ellipsom- One such source may be readily realized through the eter does not require a calibrated detector since it does process of spontaneous parametric downconversion not measure intensity but instead records a null. The (SPDC) from a second-order nonlinear crystal (NLC) principal drawback of null measurement techniques is when illuminated with a monochromatic laser beam the need for a reference to calibrate the null, for example, (pump).11 A portion of the pump photons disintegrate to find its initial location (the rotational axis of reference into photon pairs. The two photons that comprise the at which an initial null is obtained) and then to compare pair, known as signal and idler, are highly correlated this with the subsequent location upon inserting the since they conserve the energy (frequency matching) and sample into the apparatus. Such a technique thus alle- momentum (phase matching) of the parent pump photon. viates the problem of an unreliable source and detector In type II SPDC the signal and idler photons have or- but necessitates the use of a reference sample. The ac- thogonal polarizations, one extraordinary and the other curacy and reliability of all measurements depend on our ordinary. These two photons emerge from the NLC with knowledge of this reference sample. In this case, the a relative time delay due to the birefringence of the measurements are a function of ␺, ⌬, and the parameters NLC.29 Passing the pair through an appropriate bire- of the reference sample. fringent material of suitable length compensates for this Another possibility is to perform ellipsometry that em- time delay. This temporal compensation is required for ploys an interferometric configuration in which the light extracting ␺ and ⌬ from the measurements; we show sub- from the source follows more than one path, usually cre- sequently that when compensation is not employed, one ated by beam splitters, before reaching the detector. The may obtain ␺ but not ⌬. sample is placed in one of those paths. We can then es- The signal and idler may be emitted in two different di- timate the efficiency of the detector (assuming a reliable rections, a case known as noncollinear SPDC, or in the source) by performing measurements when the sample is same direction, a case known as collinear SPDC. In the removed from the interferometer. This configuration former situation, the SPDC state is polarization en- thus alleviates the problem of an unreliable detector but tangled; its is described by29 depends on the reliability of the source and suffers from the drawback of requiring several optical components 1 ͉⌿͘ ϭ ͉͑HV͘ ϩ ͉VH͘), (1) (beam splitters, mirrors, etc.). The ellipsometric mea- ͱ2 surements are a function of ␺, ⌬, source intensity, and the parameters of the optical elements. The accuracy of the where H and V represent horizontal and vertical polariza- 30 measurements are therefore limited by our knowledge of tions, respectively. It is understood that the first polar- the parameters characterizing these optical components. ization indicated in a ket is that of the signal photon, and The stability of the optical arrangement is also of impor- the second is that of the idler. Such a state may not be tance to the performance of such a device. written as the product of states of the signal and idler photons. Although Eq. (1) represents a pure quantum state, the signal and idler photons considered separately are each unpolarized.31,32 The state represented in Eq. (1) assumes that there is no relative phase between the two kets. Although the relative phase may not be zero, it can, in general, be arbitrarily chosen by making small ad- justments to the NLC. In the collinear case the SPDC state is in a polarization-product state, ͉⌿͘ ϭ ͉HV͘. (2) Because this state is factorizable (i.e., it may be written as the product of states of the signal and idler photons), it is not entangled. We first discuss a configuration based on the use of col- Fig. 1. Null ellipsometer: S is an optical source, P is a linear linear type II SPDC, which we call an unentangled twin- polarizer, ␭/4 is a quarter-wave plate (compensator), A is a linear ␪ photon ellipsometer. This configuration is introduced for polarization analyzer, and D is an optical detector; i is the angle of incidence. The sample is characterized by the ellipsometric pedagogical reasons as a precursor to the configuration of parameters ␺ and ⌬ defined in the text. principal interest to us, called the entangled twin-photon 658 J. Opt. Soc. Am. B/Vol. 19, No. 4/April 2002 Abouraddy et al.

ellipsometer, which makes use of polarization-entangled ˆ information of the field mode. The first element in J1 , photon pairs (noncollinear type II SPDC). Both arrange- Ϫ ˆ ␻ ϩ ˆ ␻ j ͕ As( ) Ai( Ј)͖, represents the annihilation opera- ments are described with a generalization of the Jones- tor of the field in beam 1, which is a superposition of sig- matrix formalism appropriate for twin-photon polarized ˆ nal and idler field operators. The second element in J1 , beams. ˆ ␻ ϩ ˆ ␻ As( ) Ai( Ј), is the annihilation operator of the field in beam 2. A. Unentangled Twin-Photon Ellipsometer We now define a twin-photon Jones matrix that repre- We now examine the use of collinear type II SPDC in a sents the action of linear deterministic optical elements, standard twin-photon polarization interferometer, previ- placed in the two beams, on the polarization of the field as ously used in numerous experiments30 and shown in Fig. follows: 2. The twin photons, with the state shown in Eq. (2), im- pinge on the input port of a nonpolarizing beam splitter, T T so that on 50% of the trials the two photons are separated ϭ ͫ 11 12ͬ 33 T , (4) into the two output ports of the beam splitter. In the T21 T22 remainder of the trials, the two photons emerge together ϭ ϫ from the beam splitter out of one of the ports, but such where Tkl (k, l 1, 2) is the familiar 2 2 Jones ma- cases do not contribute to coincidence measurements and trix that represents the polarization transformation per- thus may be ignored. Photons emerging from the one of formed by a linear deterministic optical element. The in- the output ports of the beam splitter are directed to the dices refer to the spatial modes of the input and output sample under test and are then directed to polarization beams. For example, T11 is the Jones matrix of an opti- cal element placed in beam 1 whose output is also in beam analyzer A1 followed by single-photon detector D1 . Pho- tons emerging from the other output port are directed to 1, whereas T21 is the Jones matrix of an optical element placed in beam 1 whose output is in beam 2, and similarly polarization analyzer A2 followed by single-photon detec- for T and T . In most cases, when an optical element tor D2 . A coincidence circuit registers the coincidence 12 22 is placed in beam 1 and another in beam 2, T ϭ T rate Nc of the detectors D1 and D2 , which is proportional 12 21 to the fourth-order function of the fields at the ϭ 0. An exception is, e.g., a beam splitter with beams 1 detectors.34,35 In this subsection, we demonstrate how and 2 incident on its two input ports, or other optical com- this unentangled twin-photon polarization interferometer ponents that mix the spatial modes of the two beams. yields ellipsometric measurements. The twin-photon Jones matrix T transforms a twin- ˆ ˆ ˆ ϭ ˆ We first introduce a matrix formalism that facilitates photon Jones vector J1 into J2 according to J2 TJ1 . the derivation of the fields at the detectors. We begin by Applying this formalism to the arrangement in Fig. 2, defining a twin-photon Jones vector that represents the assuming that beams 1 and 2 impinge on the two polar- field operators of the signal and idler in two spatially dis- ization analyzers A1 and A2 directly (in absence of the ␻ ␻ sample), the twin-photon Jones matrix is given by tinct modes. If aˆ s( ) and aˆ i( Ј) are the boson annihila- tion operators for the signal-frequency mode ␻ and idler- ␻Ј P͑Ϫ␪ ͒ 0 frequency mode , respectively, then the twin-photon ϭ ͫ 1 ͬ Tp ␪ , (5) Jones vector of the field following the beam splitter is 0P͑ 2͒

j͕ Ϫ Aˆ ͑␻͒ ϩ Aˆ ͑␻Ј͖͒ where Jˆ ϭ ͩ s i ͪ , (3) 1 ˆ ␻ ϩ ˆ ␻ As͑ ͒ Ai͑ Ј͒ cos2 ␪ cos ␪ sin ␪ ͑␪͒ ϭ ͫ ͬ ˆ ␻ ϭ ␻ 1 ˆ ␻Ј ϭ ␻Ј 0 36 P 2 , where As( ) aˆ s( )(0) and Ai( ) aˆ i( )(1). The cos ␪ sin ␪ sin ␪ 1 0 vectors (0) (horizontal) and (1) (vertical) are the familiar ␪ ␪ Jones vectors representing orthogonal polarization and 1 and 2 are the angles of the axes of the analyzers 37 ˆ ␻ ˆ ␻ states. The operators As( ) and Ai( Ј) thus are anni- with respect to the horizontal direction. In this case the hilation operators that include the vectorial polarization twin-photon Jones vector following the analyzers is there- fore

jP͑Ϫ␪ ͕͒ϪAˆ ͑␻͒ ϩ Aˆ ͑␻Ј͖͒ Jˆ ϭ T Jˆ ϭ ͩ 1 s i ͪ 2 p 1 ␪ ˆ ␻ ϩ ˆ ␻ P͑ 2͕͒As͑ ͒ Ai͑ Ј͖͒ cos ␪ ͕Ϫ ␪ ͑␻͒ ϩ ␪ ͑␻Ј͖͒ͩ 1 ͪ j cos 1aˆ s sin 1aˆ i Ϫ ␪ sin 1 ϭ ͩ ͪ . cos ␪ ͕ ␪ ͑␻͒ ϩ ␪ ͑␻Ј͖͒ͩ 2 ͪ cos 2aˆ s sin 2aˆ i ␪ sin 2 (6)

Using the twin-photon Jones vector Jˆ , one can obtain Fig. 2. Unentangled twin-photon ellipsometer: NLC stands for 2 nonlinear crystal; BS is a nonpolarizing beam splitter; A and A expressions for the fields at the detectors. The positive- 1 2 frequency components of the field at detectors D and D , are linear polarization analyzers; D1 and D2 are single-photon 1 2 ˆ ϩ ˆ ϩ detectors; and Nc is the coincidence rate. denoted E1 and E2 , respectively, are given by Abouraddy et al. Vol. 19, No. 4/April 2002/J. Opt. Soc. Am. B 659

N ϭ C͓tan2 ␺ cos2 ␪ sin2 ␪ ϩ sin2 ␪ cos2 ␪ ˆ ϩ͑ ͒ ϭ ͭ Ϫ ␪ ͵ ␻ ͑Ϫ ␻ ͒ ͑␻͒ c 1 2 1 2 E1 t j cos 1 d exp j t aˆ s Ϫ ␺ ⌬ ␪ ␪ ␪ ␪ 2 tan cos cos 1 cos 2 sin 1 sin 2͔, (13) cos ␪ ϩ ␪ ͵ ␻Ј ͑Ϫ ␻Ј ͒ ͑␻Ј͒ͮͩ 1 ͪ where the constant of proportionality C depends on the ef- sin 1 d exp j t aˆ i Ϫ ␪ , sin 1 ficiencies of the detectors and the duration of accumula- (7) tion of coincidences. One can obtain C, ␺, and ⌬ with a minimum of three measurements with different analyzer ϩ settings, e.g., ␪ ϭ 0°, ␪ ϭ 90°, and ␪ ϭ 45°, while ␪ Eˆ ͑t͒ ϭ ͭ cos ␪ ͵ d␻ exp͑Ϫj␻t͒aˆ ͑␻͒ 2 2 2 1 2 2 s remains fixed at any angle except 0° and 90°. If the sample is replaced by a perfect mirror, the coin- cos ␪ ϩ ␪ ͵ ␻Ј ͑Ϫ ␻Ј ͒ ͑␻Ј͒ͮͩ 2 ͪ cidence rate in Eq. (13) becomes a sinusoidal pattern of sin 2 d exp j t aˆ i ␪ , sin 2 2 ␪ Ϫ ␪ 100% visibility, C sin ( 1 2), as previously indicated. (8) In practice, by judicious control of the apertures placed in the downconverted beams, visibilities close to 100% can while the negative frequency components are given by be obtained. their Hermitian conjugates. With these fields one can To understand the need for temporal compensation dis- ϰ 2 ␪ Ϫ ␪ show that the coincidence rate Nc sin ( 1 2) using cussed previously, we rederive Eq. (13), which assumes the expressions developed in Appendix A. full compensation, when a birefringent compensator is Consider now that the sample, assumed to have placed in one of the arms of the configuration: frequency-independent reflection coefficients, is placed in ϭ 2 ␺ 2 ␪ 2 ␪ ϩ 2 ␪ 2 ␪ the optical arrangement illustrated in Fig. 2, and that the Nc C͓tan cos 1 sin 2 sin 1 cos 2 polarizations of the downconverted photons are along the Ϫ 2 tan ␺ cos ⌬ cos ␪ cos ␪ sin ␪ eigenpolarizations of the sample. The effect of the 1 2 1 sample, placed in beam 1, may be represented by the fol- ϫ ␪ ⌽ ␶ ␻ ␶ sin 2 ͑ ͒cos͑ 0 ͔͒. (14) lowing twin-photon Jones matrix: ␶ ␻ Here is the birefringent delay, 0 is half the pump fre- R0 quency, and ⌽(␶) is the Fourier transform of the SPDC ϭ ͫ ͬ Ts , (9) normalized power spectrum. When ␶ ϭ 0, we recover 0I Eq. (13), whereas when ␶ is larger than the inverse of the where SPDC bandwidth, the third term that includes ⌬ becomes zero, and thus ⌬ cannot be determined. ˜r 0 The drawback of the arrangement illustrated in Fig. 2 ϭ ͫ 1 ͬ R (10) is the requirement for a beam splitter, as in classical in- 0 ˜r2 terferometric ellipsometry. Any deviation from the as- (the justification for using this matrix to represent the ac- sumed symmetric reflectance/transmittance of this device tion of the sample is provided in Appendix A), I is the 2 will impair the measurements and necessitate the use of ϫ 2 identity matrix, and ˜r1 and ˜r2 are the complex re- a reference sample for calibration. flection coefficients of the sample described earlier. The twin-photon Jones vector after reflection from the sample B. Entangled Twin-Photon Ellipsometer and passage through the polarization analyzers is given As in classical interferometry, the configuration in the by previous subsection uses a beam splitter as a means of creating the multiple paths that lead to interference. We ˆ ϭ ˆ J3 TpTsJ1 now show that one can construct an interferometer that makes use of , which then dis- cos ␪ ͕Ϫ ␪ ͑␻͒ ϩ ␪ ͑␻Ј͖͒ͩ 1 ͪ penses with the beam splitter. This has the salutary ef- j ˜r1 cos 1aˆ s ˜r2 sin 1aˆ i Ϫ ␪ sin 1 ϭ ͩ ͪ , fect of keeping 100% of the incoming photon flux (rather cos ␪ ͕ ␪ ͑␻͒ ϩ ␪ ͑␻Ј͖͒ͩ 2 ͪ than 50%) while eliminating the requirement of charac- cos 2aˆ s sin 2aˆ i ␪ sin 2 terizing it. Moreover, no other optical elements are intro- (11) duced, so one need not be concerned with the character- ization of any components. This is a remarkable feature which results in of entanglement-based quantum interferometry. The NLC is adjusted to produce SPDC in a type II non- Eˆ ϩ͑t͒ ϭ jͭ Ϫ˜r cos ␪ ͵ d␻ exp͑Ϫj␻t͒aˆ ͑␻͒ collinear configuration, as illustrated in Fig. 3. Follow- 1 1 1 s ing the procedure discussed in the previous subsection, it is straightforward to show that the resulting coincidence cos ␪ ϩ ␪ ͵ ␻Ј ͑Ϫ ␻Ј ͒ ͑␻Ј͒ͮͩ 1 ͪ rate is given by ˜r2 sin 1 d exp j t aˆ i Ϫ ␪ , sin 1 ϭ ͓ 2 ␺ 2 ␪ 2 ␪ ϩ 2 ␪ 2 ␪ (12) Nc C tan cos 1 sin 2 sin 1 cos 2 ϩ ␺ ⌬ ␪ ␪ ␪ ␪ ͔ ˆ ϩ 2 tan cos cos 1 cos 2 sin 1 sin 2 . (15) with E2 (t) identical to Eq. (8), since there is no sample in this beam. This expression is virtually identical to the one presented Finally, using the expressions developed in Appendix A, in Eq. (13) (except for the substitution of the plus sign for it is straightforward to show that the minus sign in the last term). An interesting feature 660 J. Opt. Soc. Am. B/Vol. 19, No. 4/April 2002 Abouraddy et al.

exactly parallel. This leads to an error in the angle of in- cidence and, consequently, errors in the estimated param- eters. In our case no optical components are placed be- tween the source (NLC) and the sample; any desired polarization manipulation may be performed in the other arm of the entangled twin-photon ellipsometer. Further- more, one can change the angle of incidence to the sample easily and repeatedly. A significant drawback of classical ellipsometry is the difficulty of fully controlling the polarization of the incom- ing light. A linear polarizer is usually employed at the input of the ellipsometer, but the finite extinction coeffi- cient of this polarizer causes errors in the estimated parameters.2 In the entangled twin-photon ellipsometer the polarization of the incoming light is dictated by the phase-matching conditions of the nonlinear interaction in Fig. 3. Entangled twin-photon ellipsometer. the NLC. The polarizations defined by the orientation of the optical axis of the NLC play the role of the input po- larization in classical ellipsometry. The NLC is aligned for type II SPDC so that only one polarization component of the pump generates SPDC, whereas the orthogonal (undesired) component of the pump does not (since it does not satisfy the phase-matching conditions). The advan- tage is therefore that the downconversion process ensures the stability of polarization along a particular direction.

4. CONCLUSION Classical ellipsometric measurements are limited in their accuracy by virtue of the need for an absolutely calibrated source and detector. Mitigating this limitation requires the use of a well-characterized reference sample in a null Fig. 4. Unfolded version of the entangled twin-photon ellipsom- configuration. eter displayed in Fig. 3. Twin-photon ellipsometry, which makes use of simulta- neously emitted photon pairs, is superior because it re- of this interferometer is that it is not sensitive to an over- moves the need for a reference sample. Nevertheless, all mismatch in the length of the two arms of the setup, the unentangled twin-photon ellipsometer requires that and this increases the robustness of the arrangement. the optical components employed in the interferometric An illuminating way of representing the action of the arrangement be well characterized. entangled twin-photon quantum ellipsometer is readily We have demonstrated that entangled twin-photon el- achieved by redrawing Fig. 3 in the unfolded configura- lipsometry is self-referencing and therefore eliminates tion shown in Fig. 4. Using the advanced-wave interpre- the necessity of constructing an interferometer alto- tation, which was suggested by Klyshko in the context of gether. The underlying physics that leads to this re- twin-photon imaging,38 the coincidence rate for photons markable result is the presence of fourth-order (coinci- at D1 and D2 may be obtained by tracing light waves dence) quantum interference of the photon pairs in originating from D2 to the NLC and then onto D1 upon re- conjunction with nonlocal polarization entanglement. flection from the sample. With this interpretation, the Our proposed entangled twin-photon ellipsometer is configuration in Fig. 4 becomes geometrically similar to subject to the same shot-noise-limited, as well as angu- the classical ellipsometer. Although none of the optical larly resolved, precision that is obtained with traditional components usually associated with interferometers ellipsometers (interferometric and null systems, respec- (beam splitters and wave plates) are present in this tively), but removes the limitation in accuracy that re- scheme, interferometry is still effected through the en- sults from the necessity of using a reference sample in tanglement of the source. traditional ellipsometers. An advantage of this setup over its idealized null ellip- Since the SPDC source is inherently broadband, sometric counterpart, discussed in Section 2, is that the narrow-band spectral filters must be used to ensure that two arms of the ellipsometer are separate, and the light the ellipsometric data are measured at a specific fre- beams traverse them independently in different direc- quency. Spectroscopic data can be obtained by employing tions. This allows various instrumentation errors of the a bank of such filters. Alternatively, techniques from classical setup to be circumvented. For example, placing Fourier-transform spectroscopy may be used to directly optical elements before the sample causes beam deviation make use of the broadband nature of the source in ellip- errors39 when the faces of the optical components are not sometric measurements. Abouraddy et al. Vol. 19, No. 4/April 2002/J. Opt. Soc. Am. B 661

APPENDIX A two corresponding complex reflection coefficients) are taken into consideration. Note that the detectors actu- We investigate the effect that reflection from a sample ally record a time-averaged coincidence rate N since the has on the quantized-field operators. We model the c response time for optical detectors is usually much longer sample as a lossless beam splitter, with complex reflection than the inverse bandwidth of the function ␸(␻, ␻ coefficient ˜r and complex transmission coefficient ˜t, that p Ϫ ␻) (see Ref. 35 for details). transforms the input field operators aˆ 1 and aˆ v into output ˆ ˆ field operators b1 and bv according to ACKNOWLEDGMENTS bˆ ϭ jr˜aˆ ϩ ˜taˆ , bˆ ϭ ˜taˆ ϩ jr˜aˆ ,(A1) 1 1 v v 1 v This work was supported by the National Science Foun- where ͉˜t͉2 ϩ ͉˜r͉2 ϭ 1 (so that the bosonic commutation dation (NSF) and by the Center for Subsurface Sensing ˆ ˆ and Imaging Systems (CenSSIS), an NSF engineering re- relations are preserved for b1 and bv), aˆ 1 is the annihila- tion operator of a single mode of the incident optical field, search center. and aˆ is the annihilation operator of the vacuum enter- v A. V. Sergienko’s e-mail address is [email protected]. ing the other port of the beam splitter. The coincidence rate at the detectors at times t1 and t2 is given by REFERENCES ͑ ͒ ϭ ͗⌿͉ ˆ ͑Ϫ͒͑ ͒ ˆ ͑Ϫ͒͑ ͒ ˆ ͑ϩ͒͑ ͒ ˆ ͑ϩ͒͑ ͉͒⌿͘ 1. P. Drude, ‘‘Bestimmung optischer Konstanten der Metalle,’’ G t1 , t2 E1 t1 E2 t2 E2 t2 E1 t1 , Ann. Physik Chemie 39, 481–554 (1890). (A2) 2. R. M. A. Azzam and N. M. Bashara, Ellipsometry and Po- ˆ (ϩ) ϭ ͐ ␻ Ϫ ␻ ˆ ␻ ˆ (ϩ) larized Light (North-Holland, Amsterdam, 1977). where E1 (t) d exp( j t)b1( ), E2 (t) 3. H. G. Tompkins and W. A. McGahan, Spectroscopic Ellip- ϭ ͐ ␻ Ϫ ␻ ␻ ͉⌿͘ d exp( j t)aˆ 2( ), and is the twin-photon state sometry and Reflectometry (Wiley, New York, 1999). at the output of the NLC: 4. A. Rothen, ‘‘The ellipsometer, an apparatus to measure thicknesses of thin surface films,’’ Rev. Sci. Instrum. 16, ͉⌿͘ ϭ ͵ ␻␸͑␻ ␻ Ϫ ␻͉͒ ͘ 26–30 (1945). d , 1␻,1␻ Ϫ␻,0␻ .(A3) p p 5. A. B. Winterbottom, ‘‘Optical methods of studying films on reflecting bases depending on polarisation and interference The first element in the ket corresponds to the signal phenomena,’’ Trans. Faraday Soc. 42, 487–495 (1946). mode at frequency ␻, the second element corresponds to 6. M. Mansuripur, ‘‘Ellipsometry,’’ Opt. Photon. News 11(4), the idler mode at frequency ␻ Ϫ ␻ (conservation of en- 52–56 (2000). p 7. D. N. Klyshko, ‘‘Coherent decay of photons in a nonlinear ergy ensures that the signal and idler frequencies add up medium,’’ Pis’ma Zh. Eksp. Teor. Fiz. 6,490–492 (1967) ␻ to the pump frequency p), and the third element corre- [Sov. Phys. JETP Lett. 6,23–25 (1967)]. sponds to the vacuum mode at frequency ␻ at the other 8. S. E. Harris, M. K. Oshman, and R. L. Byer, ‘‘Observation of input port of the beam splitter that represents the tunable optical parametric fluorescence,’’ Phys. Rev. Lett. ␸ ␻ ␻ Ϫ ␻ 18,732–735 (1967). sample. The function ( , p ) is the probability 9. T. G. Giallorenzi and C. L. Tang, ‘‘Quantum theory of spon- amplitude of the possible combinations of frequencies for taneous parametric scattering of intense light,’’ Phys. Rev. pairs of signal and idler modes emitted by the NLC. 166,225–233 (1968). Inserting the identity operator 10. D. A. Kleinman, ‘‘Theory of optical parametric noise,’’ Phys. Rev. 174, 1027–1041 (1968). ͚ ͉ ͉ n ,n ,n n1 , n2 , n3͗͘n1 , n2 , n3 11. D.N.Klyshko,Photons and Nonlinear (Gordon and 1 2 3 Breach, New York, 1988). (represented in the Fock basis of the Hilbert space 12. A. Zeilinger, ‘‘Experiment and the foundations of quantum spanned by the signal, idler, and vacuum fields) into Eq. physics,’’ Rev. Mod. Phys. 71, S288–S297 (1999). 13. E. S. Fry and T. Walther, ‘‘Fundamental tests of quantum (A2) gives mechanics,’’ in Advances in Atomic, Molecular, and Optical Physics, B. Bederson and H. Walther, eds. (Academic, Bos- ϭ ⌿͉ ˆ ͑Ϫ͒ ˆ ͑Ϫ͒ ͉ ton, 2000), Vol. 42, pp. 1–27. G͑t1 , t2͒ ͚ ͗ E1 ͑t1͒E2 ͑t2͒ n1 , n2 , n3͘ n1 ,n2 ,n3 14. D. N. Klyshko, ‘‘Utilization of vacuum fluctuations as an op- tical brightness standard,’’ Kvant. Elektron. (Moscow) 4, ϫ ͗n , n , n ͉Eˆ ͑ϩ͒͑t ͒Eˆ ͑ϩ͒͑t ͉͒⌿͘ 1056–1062 (1977) [Sov. J. Quantum Electron. 7, 591–595 1 2 3 2 2 1 1 (1977)]. 15. A. Migdall, R. Datla, A. V. Sergienko, and Y. H. Shih, ‘‘Ab- ϭ ͉ ͉ ˆ ͑ϩ͒ ˆ ͑ϩ͒ ͉⌿ ͉2 solute detector quantum efficiency measurements using ͚ ͗n1 , n2 , n3 E2 ͑t2͒E1 ͑t1͒ ͘ . correlated photons,’’ Metrologia 32,479–483 (1995). n1 ,n2 ,n3 16. D. Branning, A. L. Migdall, and A. V. Sergienko, ‘‘Simulta- (A4) neous measurement of group and phase delay between two photons,’’ Phys. Rev. A 62, 063808 (2000). It is straightforward to show that using the state in Eq. 17. A. F. Abouraddy, K. C. Toussaint, Jr., A. V. Sergienko, B. E. (A3) results in all terms in the summation in Eq. (A4) Saleh, and M. C. Teich, ‘‘Ellipsometric measurements by ϭ ϭ ϭ vanishing except for the term where n1 n2 n3 0, use of photon pairs generated by spontaneous parametric so that G(t , t ) ϭ ͉͗0, 0, 0͉Eˆ (ϩ)(t )Eˆ (ϩ)(t )͉⌿͉͘2, and downconversion,’’ Opt. Lett. 26, 1717–1719 (2001). 1 2 2 2 1 1 18. A. K. Ekert, J. G. Rarity, P. R. Tapster, and G. M. Palma, ˜ also results in the terms containing taˆ v vanishing. Thus ‘‘Practical based on two-photon in- in coincidence measurements, the effect of reflection from terferometry,’’ Phys. Rev. Lett. 69, 1293–1295 (1992). the sample appears as a direct multiplication of the rel- 19. A. V. Sergienko, M. Atatu¨ re, Z. Walton, G. Jaeger, B. E. A. Saleh, and M. C. Teich, ‘‘Quantum cryptography using evant operators by the suitable complex reflection coeffi- femtosecond-pulsed parametric down-conversion,’’ Phys. cient, which justifies the use of the matrix in Eq. (10) Rev. A 60, R2622–R2625 (1999). when two orthogonal polarizations of the fields (and thus 20. T. Jennewein, C. Simon, G. Weihs, H. Weinfurter, and A. 662 J. Opt. Soc. Am. B/Vol. 19, No. 4/April 2002 Abouraddy et al.

Zeilinger, ‘‘Quantum cryptography with entangled pho- 30. A. V. Sergienko, Y. H. Shih, and M. H. Rubin, ‘‘Experimen- tons,’’ Phys. Rev. Lett. 84, 4729–4732 (2000). tal evaluation of a two-photon wave packet in type-II para- 21. D. Bouwmeester, J.-W. Pan, K. Mattle, M. Eibl, H. Wein- metric downconversion,’’ J. Opt. Soc. Am. B 12, 859–862 furter, and A. Zeilinger, ‘‘Experimental quantum teleporta- (1995). tion,’’ Nature 390, 575–579 (1997). 31. U. Fano, ‘‘Description of states in by 22. D. Boschi, S. Branca, F. De Martini, L. Hardy, and S. Pope- and operator techniques,’’ Rev. Mod. Phys. scu, ‘‘Experimental realization of teleporting an unknown 29,74–93 (1957). pure quantum state via dual classical and Einstein- 32. A. F. Abouraddy, B. E. A. Saleh, A. V. Sergienko, and M. C. Podolsky-Rosen channels,’’ Phys. Rev. Lett. 80, 1121–1125 Teich, ‘‘Degree of entanglement for two qubits,’’ Phys. Rev. A (1998). 64, 050101(R) (2001). 23. B. E. A. Saleh, S. Popescu, and M. C. Teich, ‘‘Generalized 33. R. A. Campos, B. E. A. Saleh, and M. C. Teich, ‘‘Quantum- entangled-photon imaging,’’ in Proceedings of the Ninth An- mechanical lossless beam splitter: SU(2) symmetry and nual Meeting of the IEEE Lasers and Electro-Optics Society, photon statistics,’’ Phys. Rev. A 40, 1371–1384 (1989). P. Zory, ed. (IEEE, Piscataway, N.J., 1996), Vol. 1, pp. 362– 34. R. J. Glauber, ‘‘The quantum theory of optical coherence,’’ 363. Phys. Rev. 130, 2529–2539 (1963). 24. M. B. Nasr, A. F. Abouraddy, M. C. Booth, B. E. A. Saleh, A. 35. B. E. A. Saleh, A. F. Abouraddy, A. V. Sergienko, and M. C. V. Sergienko, M. C. Teich, M. Kempe, and R. Wollenschen- Teich, ‘‘Duality between partial coherence and partial en- sky, ‘‘Biphoton focusing for two-photon excitation,’’ Phys. tanglement,’’ Phys. Rev. A 62, 043816 (2000). Rev. A 65, 023816 (2002). 36. R. J. Glauber, ‘‘Optical coherence and photon statistics,’’ in 25. A. F. Abouraddy, B. E. A. Saleh, A. V. Sergienko, and M. C. and Electronics, Les Houches, C. DeWitt, Teich, ‘‘Entangled-photon Fourier optics,’’ J. Opt. Soc. Am. B A. Blandin, and C. Cohen-Tannoudji, eds. (Gordon and (to be published). Breach, New York, 1965). 26. M. C. Teich and B. E. A. Saleh, ‘‘Observation of sub-Poisson 37. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics Franck–Hertz light at 253.7 nm,’’ J. Opt. Soc. Am. B 2, 275– (Wiley, New York, 1991). 282 (1985). 38. D. N. Klyshko, ‘‘A simple method of preparing pure states of 27. M. C. Teich and B. E. A. Saleh, ‘‘Photon bunching and anti- an optical field, of implementing the Einstein-Podolsky- bunching,’’ Prog. Opt. 26,1–104 (1988). Rosen experiment, and of demonstrating the complementa- 28. M. C. Teich and B. E. A. Saleh, ‘‘Squeezed and antibunched rity principle,’’ Usp. Fiz. Nauk 154, 133–152 (1988) [Sov. light,’’ Phys. Today 43(6), 26–34 (1990). Phys. Usp. 31(1), 74–85 (1988)]. 29. P. G. Kwiat, K. Mattle, H. Weinfurter, A. Zeilinger, A. V. 39. J. R. Zeidler, R. B. Kohles, and N. M. Bashara, ‘‘Beam de- Sergienko, and Y. Shih, ‘‘New high-intensity source of viation errors in ellipsometric measurements; an analysis’’ polarization-entangled photon pairs,’’ Phys. Rev. Lett. 75, Appl. Opt. 13, 1938–1945 (1974). 4337–4341 (1995).