throbber
Precise evaluation of polarization mode
`dispersion by separation of even- and
`odd-order effects in quantum
`interferometry
`
`A. Fraine1, D.S. Simon1,∗, O. Minaeva2, R. Egorov1, A.V. Sergienko1,3,4
`1 Dept. of Electrical and Computer Engineering, Boston University, 8 Saint Mary’s St.,
`Boston, MA 02215.
`2 Dept. of Biomedical Engineering, Boston University, 44 Cummington St., Boston, MA 02215.
`3 Photonics Center, Boston University, 8 Saint Mary’s St., Boston, MA 02215.
`4 Dept. of Physics, Boston University, 590 Commonwealth Ave., Boston, MA 02215.
`simond@bu.edu
`
`Abstract:
`The use of quantum correlations between photons to separate
`measure even- and odd-order components of polarization mode dispersion
`(PMD) and chromatic dispersion in discrete optical elements is investigated.
`Two types of apparatus are discussed which use coincidence counting of en-
`tangled photon pairs to allow sub-femtosecond resolution for measurement
`of both PMD and chromatic dispersion. Group delays can be measured with
`a resolution of order 0.1 fs, whereas attosecond resolution can be achieved
`for phase delays.
`© 2011 Optical Society of America
`OCIS codes: (260.2030) Dispersion; (120.3180) Interferometry; (270.0270) Quantum Optics.
`
`[2]
`
`[3]
`
`[4]
`
`References and links
`H. Kogelnik and R. Jopson, ”Polarization Mode Dispersion,” in Optical Fiber Telecommunications IVB System
`[1]
`and Impairments, I. Kaminow, T. Li, eds. (Academic, Press 2002), pp. 725-861.
`D. Andresciani, E. Curti, E. Matera, B. Daino, ”Measurement of the group-delay difference between the principal
`states of polarization on a low-birefringence terrestrial fiber cable,” Opt. Lett. 12 844-846 (1987).
`B. Costa, D. Mazzoni, M. Puleo, E. Vezzoni, ”Phase shift technique for the measurement of chromatic dispersion
`in optical fibers using LED’s,” IEEE J. Quantum Elect. 18, 1509-1515 (1982).
`C.D. Poole, C.R. Giles, ”Polarization-dependent pulse compression and broadening due to polarization disper-
`sion in dispersion-shifted fiber,” Opt. Lett. 13, 155-157 (1987).
`C.D. Poole, ”Measurement of polarization-mode dispersion in single-mode fibers with random mode coupling,”
`Opt. Lett. 14 523-525 (1989).
`D. Derickson, Fiber Optic Test and Measurement (Prentice Hall, 1998).
`B. Bakhshi, J. Hansryd, P.A. Andrekson, J. Brentel, E. Kolltveit, B.K. Olsson, M. Karlsson, ”Measurement of
`the differential group delay in installed optical fibers using polarization multiplexed solitons,” IEEE Phot. Tech.
`Lett. 11, 593-595 (1999).
`S. Diddams, J. Diels, ”Dispersion measurements with white-light interferometry,” J. Opt. Soc. Am. B 13, 1120-
`1129 (1996).
`P. Williams, ”PMD measurement techniques and how to avoid the pitfalls,” J. Opt. Fiber. Commun. Rep. 1,
`84-105 (2004).
`D. Branning, A.L. Migdall, A.V. Sergienko, ”Simultaneous measurement of group and phase delay between two
`photons,” Phys. Rev. A 62, 063808 (2000).
`E. Dauler, G. Jaeger, A. Muller, and A. Migdall, ”Tests of a Two-Photon Technique for Measuring Polarization
`Mode Dispersion With Subfemtosecond Precision,” J. Res. Natl. Inst. Stand. Technol. 104, 1-10 (1999).
`[12] M.H. Rubin, D.N. Klyshko, Y.H. Shih, A.V. Sergienko, ”Theory of two-photon entanglement in type-II optical
`parametric down-conversion,” Phys. Rev. A 50, 5122-5133 (1994).
`
`[5]
`
`[6]
`[7]
`
`[8]
`
`[9]
`
`[10]
`
`[11]
`
`0001
`
`
`
`
`
`Capella 2022
`Fujitsu v. Capella
`IPR2015-00727
`
`

`
`[13]
`[14]
`
`D.N. Klyshko, Photons and Nonlinear Optics (Gordon and Breach, 1988).
`O. Minaeva, C. Bonato, B.E.A. Saleh, D.S. Simon, A.V. Sergienko, ”Odd- and Even-Order Dispersion Cancella-
`tion in Quantum Interferometry,” Phys. Rev. Lett. 102, 100504 (2009).
`
`Introduction: Dispersion Measurement - Classical versus Quantum
`1.
`As optical communication networks migrate towards higher 40 Gbps and 100 Gbps data rates,
`system impairments due to dispersion, especially polarization mode dispersion (PMD), become
`a primary issue. This includes not only fiber PMD, but also contributions from switches, ampli-
`fiers, and all other components in the optical path. The fiber PMD and component PMD tend to
`accumulate in different manners as the size of the network grows. In the long length regime, the
`√
`differential group delay (DGD) due to fiber PMD has a known dependance on length, growing
`L [1]. In a similar manner, contributions from chromatic dispersion increase linearly in L.
`as
`This known length dependence makes the dispersion of the optical fibers themselves relatively
`straightforward to measure and to take into account.
`In contrast, component PMD was until recently considered to be too small in comparison to
`fiber PMD to affect significant penalties at the system level. Since the introduction of recon-
`figurable add-drop multiplexers (ROADMs), the number of components that could potentially
`contribute to the PMD in a given system has increased significantly. Although the dispersive
`contribution of each separate component is relatively small, together they are capable of accu-
`mulating and of thereby making a significant contribution to the total system impairment. It is
`therefore important to be able to precisely and efficiently measure small values of DGD. How-
`ever, since only fiber PMD was important in the past, no measuring techniques were developed
`for efficient evaluation of small DGD values. With component PMD starting to play a signifi-
`cant role, developing high-resolution evaluation of small PMD values in a single optical switch
`or other small discrete optical component represents a new challenge to optical researchers that
`must be addressed by modern optical metrology. In this paper, we address the measurement of
`dispersive effects in such discrete elements.
`Polarization mode dispersion is the difference between wavenumbers of two orthogonal
`states of light at fixed wavelength, or equivalently, a polarization-dependent variation of a
`material’s index of refraction. A number of methods have been developed for measuring it
`[2, 3, 4, 5, 6, 7, 8, 9]. Many traditional techniques for measuring PMD rely on an interfero-
`metric approach for high-resolution measurements of absolute values of optical delays. This
`approach requires one to use a monochromatic laser source and to keep track of the number
`of interference fringes. Therefore, the accuracy of the approach is limited by the stability of
`the interferometer, by the signal-to-noise level of the detector, and by the wavelength of the
`monochromatic radiation, leading to significant limitations. For example, the use of monochro-
`matic classical polarized light does not allow one to measure the relative delay between two
`orthogonally polarized waves in a single measurement, so several measurements at different
`frequencies must be used to reconstruct the polarization dispersion properties of materials. The
`use of highly monochromatic laser sources creates the additional problem of multiple reflec-
`tions and strong irregular interference that may have detrimental effect on measuring polariza-
`tion dispersion.
`White-light or low-coherence interferometry [8] is another widely used approach. The ulti-
`mate resolution of such interferometric measurements will depend on the spectral bandwidth
`of the light source. Achieving sub-fs resolution in PMD measurement dictates the use of light
`sources with bandwidth in excess of 200 nm. Generating light of such a bandwidth with a
`smooth spectral profile is not an easy task in itself. Spectral modulations from existing sources
`with bumpy spectra produce ’ghost’ features during measurement, leading to complications in
`dispersion evaluation. In addition, the visibility of interference with such super-broadband light
`
`0002
`
`

`
`is diminished due to dispersion effects.
`Overall, while classical techniques can provide high-resolution measurement of polariza-
`tion mode dispersion they still have limitations in many areas that quantum-based techniques
`can address. For example, entangled photon states intrinsically provide an absolute value for
`polarization optical delay, in contrast to the conventional (classical) case, which is limited to
`determination of delay modulo an integer number of cycles of the light. This is mainly due to
`the fact that quantum interferometry exploits both phase and group velocity effects in the same
`measurement [10, 11], a feat not possible in classical optics.
`The current practical resolution of conventional dispersion evaluation techniques is limited
`to a few femtoseconds (fs). The primary goal here is to use an interferometric setup with an
`entangled photon source to measure the component PMD of a small, discrete optical element
`to sub-femtosecond precision. Ideally, it would be desirable to measure chromatic dispersion
`with the same device, while allowing for the polarization and chromatic effects to be easily
`separable. We will show that this is indeed possible. Due to the frequency-anticorrelation in
`the entangled downconversion source used as illumination, we may independently determine
`the even-order and odd-order parts of the PMD’s frequency dependence. Due to the reliance on
`the frequency anticorrelations within pairs of photons, the separation method is intrinsically a
`two-photon quantum effect, and is not present in the classical interferometer.
`Classical attempts to simulate this even-odd separation effect by symmetrical chirping and
`anti-chirping of femtosecond laser pulses are constrained to a very narrow size of wavepacket
`thus making it not very practical. The availability of such a separation is useful in a number
`of circumstances. One example is when there is enough pulse broadening (second-order dis-
`persion) to make accurate measurement of group velocity (first order dispersion) difficult. In a
`fiber, group velocity and broadening effects can be separated to some extent by simply taking
`a sufficiently long length of fiber as sample; the longer the fiber, the more accurately each can
`be measured. When dealing with switching elements or other small discrete optical elements,
`this option is not available. Another means must be found to prevent accurate measurement of
`the first-order group delay from being obscured by second-order broadening effects. That is
`what is accomplished here: the location of a dip in the coincidence rate may be used to find
`the group velocity, and this location is unaffected by the second-order broadening as a result
`of the even-order dispersion cancellation. Conversely, although the amount of broadening in
`a single small component may seem negligible, the total broadening from many such compo-
`nents present in a large network may be significant; thus high-accuracy measurements of these
`very small second-order dispersive contributions is important. Separating them off from the
`generally larger first-order contributions makes accurate measurements much easier.
`We note that since the component being analyzed is assumed to be relatively small, the prin-
`cipal polarization axes may be assumed to remain constant over the longitudinal length of the
`object and to be independent of frequency, with the dispersive contributions of the two polar-
`ization components remaining independent of each other. This greatly simplifies the analysis.
`After a review of background and notation in section 2, three measurement methods will be
`discussed in sections 3-5. The apparatus of section 3 uses a single detector to make a classical
`measurement; the system is illuminated with a broadband classical light source. In contrast,
`quantum measurements are made using two detectors connected in coincidence with illumina-
`tion provided by a source of entangled photon pairs (spontaneous parametric downconversion,
`(SPDC)). We will examine two quantum measurement setups in sections 4 and 5. In addition,
`in section 5 we give a qualitative analysis that allows the positions of dips (or peaks) indepen-
`dently of the mathematical formalism.
`The two quantum configurations will be distinguished from each other by referring to them
`as type A or type B. They differ only in the presence or absence of a final beam splitter before
`
`0003
`
`

`
`detection, so they may both be implemented in a single apparatus by allowing a beam splitter
`to be switched in or out of the optical path. Similarly, by adding an additional polarizer and
`counting the singles rate at one detector instead of coincidence events, the classical setup may
`also be implemented in the same device. Thus, a single apparatus could be made which is
`capable of performing any of the three types of measurements to be discussed.
`This paper builds on two previous lines of work. The apparatus used for the type A setup was
`introduced previously [10, 11], where it was shown that quantum interferometry can achieve
`higher resolution than classical methods in measurements of PMD. Separately, the segregation
`of even- and odd-order chromatic dispersion effects was demonstrated in [14]. Here, we bring
`the two strands together in a single device (type B), showing that we can separate even- and
`odd-order effects in PMD, as well as in chromatic dispersion, and that we can do so with the
`resolution available to the type A device.
`As a further benefit of the quantum devices over classical methods, note that for the quantum
`cases there is no need to know in advance the principal axis directions of the device or ob-
`ject being measured. Although the incoming photons are aligned along particular axes that are
`linked to a birefringent crystal orientation, their projections onto any rotated pair of orthogonal
`axes (including the principle axes of the sample) will remain equally entangled, allowing the
`method to work without any need to align the axes of the source and the device under test.
`
`2. Chromatic Dispersion and Polarization Mode Dispersion
`First consider a material for which the index of refraction is independent of polarization. The
`frequency dependence of the wavenumber k = 2πn(λ)
`is given by a dispersion relation, which

`can be written near some central frequency Ω0 as
`k(Ω0 ± ω) = k0 ± αω+ βω2 ± γω3 + . . .
`(1)
`for |ω| << Ω0. The coefficients α, β, . . . characterize the chromatic dispersion or variation of
`the refractive index with frequency. Explicitly,
`
`k0 = k(Ω0),
`d2k(ω(cid:4))
`dω(cid:4)2
`
`β = 1
`2!
`
`(cid:2)(cid:2)(cid:2)(cid:2)
`
`α= dk(ω(cid:4))
`dω(cid:4)
`γ= 1
`3!
`
`,
`
`,
`ω(cid:4)=Ω0
`d3k(ω(cid:4))
`dω(cid:4)3
`
`(cid:2)(cid:2)(cid:2)(cid:2)
`
`ω(cid:4)=Ω0
`ω(cid:4)=Ω0
`Rather than looking at the individual terms in the expansion (1), we may also collect together
`all terms containing even powers of ω and all terms containing odd powers to arrive at an
`expansion containing only two terms:
`k(Ω0 + ω) = keven(ω) + kodd(ω),
`
`(cid:2)(cid:2)(cid:2)(cid:2)
`
`, . . .
`
`(2)
`
`(3)
`
`(4)
`
`where
`
`keven(ω) = k0 + βω2 + O(ω4),
`
`(5)
`
`and
`
`kodd(ω) = αω+ γω3 + O(ω5).
`(6)
`In the case of nonzero polarization mode dispersion (PMD), the index of refraction varies
`with polarization. We now have two copies of the dispersion relation, one for each independent
`polarization state:
`
`kH (Ω0 ± ω) = kH0 ± αHω+ βHω2 + . . .
`
`(7)
`
`0004
`
`

`
`(8)
`(9)
`(10)
`
`= kH,even(ω) + kH,odd(ω)
`kV (Ω0 ± ω) = kV0 ± αV ω+ βVω2 + . . .
`= kV,even(ω) + kV,odd(ω),
`where H,V denote horizontal and vertical polarization.
`To describe the PMD, we must define quantities that measure the differences between the
`two polarization states:
`Δβ = βV − βH.
`Δα= αV − αH,
`Δk0 = kV0 − kH0,
`These parameters are defined per unit length. For the case of primary interest to us, discrete
`fixed-size objects, the formulas should really be written in terms of the relevant lumped quan-
`tities
`ΔA ≡ lΔα,
`ΔB ≡ lΔβ,
`Δφ≡ lΔk0,
`(12)
`where l is the axial thickness of the device under study. However, we will continue to use the
`α, β, and Δk0 parameters of eq. 11, both because they are more commonly used, and because
`they allow easy comparison to the formulas used in fiber optics.
`Note that Δk0 = Ω0Δn(Ω0)
`is a measure of the difference in phase velocity between the two
`c
`polarization modes, while Δα and Δβ are related to the difference in group velocity. Also, it
`should be pointed out that the PMD and the chromatic dispersion are not entirely independent
`effects; in particular, the PMD coefficients themselves (Δk0, Δα, Δβ) are frequency dependent.
`In the quantum cases, it is convenient to also define τ− = DL, where L is the thickness of the
`0 − u−1
`nonlinear downconversion crystal and D = u−1
`is the difference of the group velocities
`e
`of the two polarizations inside the crystal. We will restrict ourself to the simplest case of a
`bulk crystal, so the spectral distribution of the downconverted pairs is described by the function
`(cid:4)
`(cid:3)
`[12, 13]
`τ−ω
`
`(11)
`
`(13)
`
`,
`
`12
`
`Φ(ω) = sinc
`
`where the sinc function is defined by sinc(x) = sin(x)
`. Photons are emitted from the downconver-
`sion process in frequency- anticorrelated pairs: the frequencies Ω0 ± ωin each pair are shifted
`x
`equally, but in opposite directions, from the central frequency Ω0 = ωpump/2, with the distribu-
`tion of frequency shifts ω being given by Φ(ω) of eq. 13. The downconversion time scale,
`τ−, is inversely proportional to the spectral width of the source, and therefore determines the
`precision of the resulting measurements. The spectrum may be made wider by using a thinner
`nonlinear crystal, but this occurs at the expense of reducing the intensity of the downconverted
`light. High intensity and large bandwidth may be obtained simultaneously by use of a chirped
`crystal, although some of the details of the following analysis will then be changed.
`
`3. Classical PMD Measurement
`An apparatus equivalent to that shown schematically in fig. 1 [8] is commonly used to measure
`polarization mode dispersion. The illumination may be provided by any sufficiently broadband
`light source. For easier comparison with the later sections, we will assume the illumination is
`provided by type II parametric downconversion, but this is not necessary; since we use a single
`detector, the entanglement of the downconverted photons will play no role.
`Assume an arbitrary amount of H and V polarization out of the downconversion crystal, so
`(cid:5) (cid:3)
`(cid:4)
`that the incident field in Jones vector notation is proportional to
`AH (ω)
`AV (ω)
`
`dω,
`
`(14)
`
`0005
`
`

`
`Crystal
`
`Half-wave
`plate (45°)
`
`NPBS
`
`Pump
`
`Horizontal
`Polarizer
`
`d
`
`1
`
`d
`
`2
`

`
`Non-
`birefringent
`delay
`
`Polarizer
`at 45°
`
`Object
`
`D
`
`NPBS
`
`l
`
`Fig. 1. Classical (single-detector) white-light setup for finding total PMD.
`
`where AH and AV are the incoming amplitudes of the horizontal and vertical components. After
`a horizontal polarizer, we destroy the quantum state and just pick off one component. We can
`(cid:4)
`(cid:5) (cid:3)
`think of it as a classical broadband source of horizontally polarized light,
`AH(ω)
`0
`
`dω.
`
`(15)
`
`dω=
`
`dω.
`
`(16)
`
`For path 1 (lower), the horizontally polarized light accumulates a phase corresponding to the
`path length d1. For path 2 (upper), the horizontally polarized light passes through a λ
`2 wave
`plate with fast axis 45◦ from the horizontal, converting it into vertically polarized light,
`(cid:4)(cid:5) (cid:3)
`(cid:5) (cid:3)
`(cid:4)
`(cid:3)
`(cid:4)
`0 −1
`AH(ω)
`0
`AH (ω)
`1
`0
`0
`In addition, the vertically polarized light in path 2 experiences a phase corresponding to the
`path length d2 and an adjustable delay δ = cτ2.
`(cid:4)
`(cid:3)
`At the second beam splitter, the two components form a superposition of the form
`(cid:5)
`eik(ω)d1
`eik(ω)(d2+δ)
`with k(ω) = ω
`c (assuming the paths are in free space). In the absence of any sample after the
`second beam splitter, this superposition will pass through a linear polarizer at 45◦, resulting in
`(cid:4)
`(cid:3)
`(cid:5)
`AH (ω)√
`eik(ω)d1 + eik(ω)(d2+δ)
`J(cid:4)
`0 =
`eik(ω)d1 + eik(ω)(d2+δ)
`2
`The intensity at the detector is then given by
`(cid:5)
`I = |J(cid:4)
`0|2 =
`|AH (ω)|2 [1 + cos(k(ω) (Δd − δ))] dω.
`Here, Δd = d1 − d2 is the path length difference between the two arms.
`If a birefringent sample of length l is introduced between the last beam splitter and the final
`(cid:4)
`(cid:3)
`polarizer, an additional polarization-dependent phase shift is added to the vector in eq. 17:
`(cid:5)
`ei[k(ω)d1+kH (ω)l]
`e[ik(ω)(d2+δ)+kV (ω)l]
`
`(17)
`
`(18)
`
`(19)
`
`(20)
`
`J0 =
`
`AH(ω)
`
`dω,
`
`dω,
`
`J0 =
`
`AH (ω)
`
`dω.
`
`0006
`
`

`
`Fig. 2. Interferograms produced by apparatus of fig. 1 for samples of different thicknesses.
`For a fixed thickness, the size of the shift may be used as a measure of the difference in
`phase velocities of the two polarizations.
`
`(cid:7)ω
`
`(cid:8)(cid:9)
`
`dω.
`
`(21)
`
`(cid:6)
`The resulting intensity at the detector is then:
`(cid:5)
`I = |J(cid:4)
`(Δd − δ)− Δk(ω)l
`0|2 =
`|AH (ω)|2
`1 + cos
`c
`(cid:11)
`(cid:10)
`For Type II downconversion, the AH (ω) and AV (ω) are both proportional to Φ(ω) =
`τ−ω
`. Plotting eq. 21 as a function of birefringent delay δ leads to interferograms such
`sinc
`as those shown in fig. 2. Each interferogram will be phase shifted (moving the positions of
`the peaks and troughs within the envelope) due to the zeroth order difference in dispersion
`Δk0, while the envelope as a whole will be shifted horizontally due to the first order difference
`in dispersion Δα and broadened due to the second order difference Δβ. The interferograms
`shown in fig. 2 are shifted by different amounts due to the use of different sample thicknesses.
`In this plot, a 200 nm bandwidth centered at 1550 nm was assumed, with a coherence length of
`λ2
`0Δλ = 12 μm.
`xc =
`4. Type A Quantum Measurement
`The goal now is to extract the polarization mode dispersion of an object with a higher precision
`than is possible with the classical apparatus of the previous section. In addition, we would like
`to be able to measure the even and odd orders of chromatic dispersion for each polarization.
`The setup [10, 11] is shown in fig. 3. The downconversion is type II so that the two photons
`have opposite polarization (H and V). The photons have frequencies Ω0 ± ω, where 2Ω0 is
`the pump frequency. Controllable birefringent time delays τ1 and τ2 are inserted before and
`after the beam splitter. Objects of lengths l1 and l2 may be placed before and after the beam
`splitter, respectively. Polarizers at angles θ1 and θ2 from the horizontal are placed before the
`two detectors. In the following, we will take θ1 = θ2 = π
`4 and assume that the beam splitter is
`50/50. Information about which polarization state travels in which branch of the apparatus will
`therefore be erased, allowing interference to occur with maximum visibility.
`Rather than the Jones matrix formalism used in the previous section, it will be more conve-
`nient here to use creation and annihilation operators for horizontally and vertically polarized
`photons. The portion of the output from the downconversion process that is relevant to our
`purposes is the biphoton state
`(cid:5)
`|Ψ(cid:7) =
`
`
`
`dωΦ(ω) ˆa†H (Ω0 + ω) ˆa†V (Ω0 − ω)|0(cid:7),
`
`
`
`(22)
`
`0007
`
`12
`
`

`
`ˆE (+)
`1
`ˆE (+)
`2
`
`(cid:2)(cid:2)(cid:2)2
`
`G(2)(t1,t2) =
`
`Rc(τ1,τ2) =
`
`dt1dt2G(2)(t1,t2).
`
`=
`
`e−iω1t1
`
`(23)
`
`(24)
`
`(25)
`
`(26)
`
`(27)
`dω|Φ(ω)|2 = 2π
`τ− .
`
`(cid:13)
`
`(28)
`
`which will serve as the incident state of our setup. The positive-frequency parts of the fields at
`detectors D1 and D2, respectively, can be written in the forms
`(cid:12)
`(cid:13)
`(cid:5)
`(t1) = 1
`ˆaH (ω1)eikH (ω1)l1 + ˆaV (ω1)ei[kV (ω1)l1+ω1τ1]
`dω
`(cid:12)
`(cid:5)
`2
`(t2) = 1
`ˆaH (ω2)eikH (ω2)(l1+l2)
`dω
`(cid:13)
`2
`e−iω2(t2+τ).
`+ ˆaV (ω2)ei[kV (ω2)(l1+l2)+ω2(τ1+τ2)]
`The coincidence rate is then computed by integrating the correlation function
`(t2)|Ψ(cid:7)|2
`
`(cid:2)(cid:2)(cid:2)(cid:8)0|E (+)
`(t1)E (+)
`1
`2
`over the characteristic time scale T of the detectors:
`(cid:5) T /2
`−T /2
`Since T is generally much larger than the downconversion time τ−, we may safely simplify by
`taking T → ∞.
`Using eqs. 22-26, the coincidence rate may be written in the general form ([12])
`Rc(τ1,τ2) = R0{1 +CM(τ1,τ2)} ,
`(cid:14)
`where R0 is a constant (delay-independent) background term and C−1 =
`The dependence on the time delays is contained in the modulation term
`(cid:5)
`M(τ1,τ2) = 1
`dω|Φ(ω)|2 e−i[Δk(ω)−Δk(−ω)]l1−2iωτ1
`(cid:12)
`2

`eiΔk(−ω)l2+i(Ω0−ω)τ2 + e−iΔk(ω)l2−i(Ω0+ω)τ2
`(cid:5)
`dω|Φ(ω)|2
`(29)
`× cos{[Δk(ω)− Δk(−ω)]l1 + Δk(ω)l2 + 2ωτ1 + (Ω0 + ω)τ2} ,
`where the second form follows by changing the sign of the integration variable in the first term
`of the previous line. It can be seen that even-order PMD terms arising from the pre-beam splitter
`object cancel. Thus, measurements made with the object before the beam splitter will give us
`the odd-order PMD, and measurements made with the object after the beams splitter give the
`total PMD; making both measurements and then taking the difference will provide the even-
`order PMD. We can see the roles of the even and odd parts more clearly by splitting Δk into its
`even and odd parts, then using the identity cos(A + B) = cosAcosB− sinAsinB. The result is:
`(cid:5)
`
`M(τ1,τ2) =
`
`dω|Φ(ω)|2{cos [Δkodd(ω)(2l1 + l2) + ω(2τ1 + τ2)]cos [Δkeven(ω)l2 + Ω0τ2]
`−sin [Δkodd(ω)(2l1 + l2) + ω(τ1 + τ2)]sin [Δkeven(ω)l2 + Ω0τ2]} .
`(30)
`Note that the integrand in the second term is odd in ω, so the integral over that term vanishes.
`Therefore, this simplifies to
`(cid:5)
`
`M(τ1,τ2) =
`
`dω|Φ(ω)|2 cos [Δkodd(ω)(2l1 + l2) + ω(2τ1 + τ2)]
`× cos [Δkeven(ω)l2 + Ω0τ2] .
`
`(31)
`
`0008
`
`

`
`Controllable birefringent delays
`
`Crystal
`
`Object
`

`1
`
`1l
`
`Object
`
`l2
`
`Pump
`
`Polarizer
`
`D
`2
`

`
`2
`
`Polarizer
`
`D
`1
`
`Coincidence
`counter
`
`Fig. 3. Type A setup for measuring PMD parameters Δα≡ αV − αH and Δβ≡ βV − βH.
`
`We see that the even- and odd-order terms have separated into different cosine terms.
`In the special case that Δβ and all higher order terms vanish, the integral of the previous line
`can be done explicitly:
`(cid:16)
`(cid:15)
`M(τ1,τ2) = 2π
`τ− cos [Δk0l2 + Ω0τ2]Λ
`In the last line we have used the result
`(cid:5)
`
`Δα(2l1 + l2) + (2τ1 + τ2)
`τ−
`(cid:7) τ
`(cid:8)
`
`,
`
`2a
`
`π a
`

`
`dωsinc2(aω)cos (ωτ) =
`(cid:17)
`
`.
`
`(32)
`
`(33)
`
`(34)
`
`(35)
`
`where
`
`Λ(x) =
`
`1−|x|,
`0,
`
`|x| ≤ 1
`|x| > 1
`
`is the unit triangle function.
`(cid:17)
`The coincidence rate is then
`
`Rc(τ1,τ2) = R0
`
`1 + cos[Δk0l2 + Ω0τ2]Λ
`
`(cid:15)
`
`Δα(2l1 + l2) + (2τ1 + τ2)
`τ−
`
`(cid:16)(cid:18)
`
`.
`
`This result is consistent with equation A31 of [10], with the caveat that an extra time delay
`τ1 has been added here. We now have two possibilities: we can scan over τ1 while holding τ2
`fixed, or vice-versa. If we scan over τ1 with τ2 = 0, we find a triangular dip similar to the HOM
`dip, as shown in fig. 4. The first order term in the PMD, Δα shifts the triangular envelope left
`or right, so that the bottom of the dip is at τ1 = − Δα
`(2l1 + l2); thus Δαmay be determined by
`2
`measuring the location of the minimum. The absolute value of the factor cos(Δk0l2) in front of
`the triangle function gives the visibility of the dip; so measuring the depth of the dip allows Δk0
`to be determined. Note that (depending on the sign of cos(Δk0l2)) the ”dip” may actually be a
`peak.
`Alternatively, we may scan over τ2 while holding τ1 = 0. This leads to an oscillating inter-
`ference fringe pattern within the triangular envelope, similar to those of fig. 2. The shift of the
`triangular envelope allows Δα, the first order term in the PMD, to be determined as before. In
`this case, rather than determining visibility, the zeroth order term Δk0 horizontally shifts the
`fringe pattern by distance τ2 = Δk0l2Ω0
`within the envelope, allowing determination of Δk0 from
`the size of this shift. To see clearly the effects of each order of dispersion, fig. 5 shows examples
`
`0009
`
`

`
`Coincidence rate
`
`
`
`_.'A(I(L.)ll
`2
`
`0
`
`T1
`
`Fig. 4. Scanning over 1:; while keqring 1:2 = 0. The horizontal shifi‘ ofthe minimum away
`fiom the origin determines Aa, while the depth of the dip determines
`Ihe triangle
`function mayleadeithertoadip (asshown)ortoapealr, dqrendingon thesign ofthe
`cosine.
`
`(8)
`
`(D)
`
`Fig. 5. Scanning over 1:2 while beping 1:1 = 0. In (a), a nonzem Au shifm the envelope
`from itspositionfor At! = 0 in part
`The size ofthe shift can be measured with accuracy
`on the order of0.1
`In part (c), a nonzero Ah) shifts the locations of the peak within
`the unshified envelope. The size ofthe shift can be measured with accuracy on the order of
`.001 fs = l as.
`
`0010
`
`

`
`of such scans in the presence of zeroth-order and first-order dispersion separately. The fringes
`within the envelope as τ2 is scanned allow evaluation of the phase delays (the Δk0 term) to an
`accuracy on the order of attoseconds (10−18 s) [10]. Group delays from the Δα term down to
`the order of 0.1 fs.
`Note that only the differences of the horizontal and vertical polarization parameters (Δα, Δβ,
`etc.) appear in the formulas above. The resulting interferogram is independent of the values of
`the parameters for fixed polarization (αH, αV , etc.) and so are insensitive to non-polarization-
`dependent dispersive effects.
`In principle, Fourier transforming experimental data for the coincidence rate and then fit-
`ting parameterized curves to it will allow the determination of higher order PMD parameters.
`However, this requires a large quantity of data to be obtained at high precision. By adding an
`additional beam splitter to the apparatus in the next section, we will arrive at a better method,
`which allows us to extract additional information; namely, it will also give us information about
`the H and V polarizations separately, not just their difference.
`
`5. Type B Quantum Measurement
`The goal here is to see if additional information may be obtained with a variant of the previous
`apparatus that mixes the final beams via an additional beam splitter. This variation is inspired
`by the setup of ref. [14], in which even and odd portions of the chromatic dispersion were
`separated into different parts of an interferogram, allowing them to be studied independently of
`each other.
`Consider the setup in fig. 6. This differs from the setup of the previous section (fig. 3) only
`by the addition of an extra beam splitter before the detectors and an additional nonbirefringent
`delay τin one arm, after the first beam splitter. Two birefringent samples of lengths l1 and l2 are
`placed before and after the first beam splitter. Birefringent delays τ1 and τ2 are present before
`and after the beam splitter as well, and a nonbirefringent delay τ is added to one of the two
`arms after. For the sake of definiteness, assume that τ1 and τ2 delay the vertical (V) polarization
`and leave the horizontal (H) unaffected. The system is illuminated with type II downconversion
`beams. The pump frequency is at 2Ω0, while the signal and idler frequencies will be written as
`Ω0 ± ω. We will make use of the fact that the downconversion spectral function is symmetric,
`Φ(ω) = Φ(−ω).
`(36)
`
`We will identify the e and o polarizations with V and H respectively.
`It should be emphasized that in the notation used here, τ is an absolute delay, so it must be
`positive. However, τ1 and τ2 are relative delays of the vertically polarized photon compared to
`the horizontal, and so τ1 and τ2 may be positive or negative.
`We will find below that the effects of the even and odd orders separate and play different
`roles: the location of each dip in the interferogram (represented mathematically by a triangle
`function in the coincidence rate) is determined by the odd part, while the relative depths of
`the dips are controlled by the even part. We may predict the number and location of each of
`these dips by identifying the ways in which it becomes impossible from the relative timing of
`detection events in the two detectors to identify which photon took which path. To do so, first
`note that the delay between the V and H photons arising before the first beam splitter is
`Δτpre ≡ τV − τH = Δαl1 + τ1.
`(37)
`(This is the delay due to the object and τ1 alone; it is assumed that the intrinsic delay introduced
`by the known birefringence of the crystal itself has been compensated.) There are four possible
`ways in which the delay after the first beam splitter may compensate this pre-beam splitter
`
`0011
`
`

`
`Controllable birefringent delays
`
`Crystal
`
`Object
`

`1
`
`1l
`
`Object
`
`l2
`
`Pump
`
`Non birefringent
`delay
`

`
`2
`

`
`D 2
`
`Polarizers
`
`D
`1
`
`Coincidence counter
`
`Fig. 6. Type B setup for finding even- and odd-order PMD
`
`delay, leaving a total delay of zero between the two photons. These are enumerated in the
`table of figure 7, which gives the total post-beam splitter delay Δτpost for each case in the final
`column. Setting
`Δτpre + Δτpost = 0
`(38)
`for these four possibilities predicts four dips in the coincidence rate at delay values for which
`the difference in the final column vanish; at these values, there is no path information available
`because the two photons arrive at the detector simultaneously, allowing for complete destructive
`interference between paths.
`One additional dip (represented by the second triangle function in eq. 46 below) arises in a
`slightly different fashion. Here the time delay between the two photons is nonzero, but has a
`value tha

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket