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

Frequency up-conversion of quantum images

Open Access Open Access

Abstract

We develop the theory of frequency up-conversion of an arbitrary spatially-broadband quantum state (quantum image), deriving the analytical solutions for the cases of plane-wave pump and short crystal with arbitrary pump. By using an example of the quantum image imposed by the orbital angular momentum of the pump beam in spontaneous parametric down-conversion, we show that 99%-fidelity up-conversion of quantum images from an infrared wavelength to the visible wavelength can be obtained in periodically-poled lithium-niobate crystals at reasonable pump intensities well below the crystal damage threshold.

©2012 Optical Society of America

1. Introduction

Light with quantum correlations among various spatial modes (a quantum image) can be used in quantum communications, computing, and measurements. From a practical point of view, it is convenient to produce such quantum images via optical parametric processes that are pumped by easily available high-power laser sources; e.g., those with wavelengths at 532 nm, 780 nm, 1064 nm, or 1550 nm. The resulting quantum images, however, fall into the infrared region of the electromagnetic spectrum, where arrayed detectors have very poor quantum efficiencies, making it difficult to observe and utilize the spatial quantum features of light. One potential way to overcome these low detection efficiencies is to use phase-sensitive parametric amplifiers, which are uniquely capable of amplifying input signals without degrading their signal-to-noise ratios (noiseless amplification) [1]. This property, in combination with the wide temporal bandwidth of the parametric processes, has already been used to make noiseless amplifiers for fiber-optic communications [26]. The wide spatial bandwidth of these processes also enables their use for noiseless amplification of classical images [711]. It was also recently proposed to use a phase-sensitive parametric amplifier to noiselessly amplify some quantum images, in which multimode squeezed vacuum is used to suppress the loss of information at the periphery of a soft aperture of a spatially-broadband optical receiver [12]. While this approach allows one to extract the quantum features of the original input light even with low-efficiency detectors (e.g., those available in the infrared region of the electromagnetic spectrum), it can amplify only one quadrature of the input light, making this method unsuitable for quantum images in which the information is encoded in photon-number correlations.

An alternative method for overcoming the low infrared detection efficiency is the sum-frequency-generation (SFG) based up-conversion of images into the visible spectrum [13], where efficient arrayed detectors are readily available. This approach has already been demonstrated to provide high-fidelity up-conversion for the parametric twin-beam state [14], single-photon states [15, 16], and quantum memory [17], all of which were spatially single-mode states. A theory of frequency-degenerate image up-conversion via second-harmonic generation was recently introduced in [18], but its practical implementation requires high input signal powers that restrict the types of quantum states that can be up-converted. Another mechanism of frequency conversion relying on quantum teleportation protocol has also been proposed [19]. A quantum theory for simultaneous parametric amplification and up-conversion in the same crystal has recently been developed [20, 21], but it discussed only the signal-to-noise-ratio evolution of an input coherent-state (i.e., classical) image. In this paper, we discuss the up-conversion of a spatially-multimode quantum state (i.e., a quantum image). In Section 2 we derive the analytical expressions for the up-conversion efficiencies in the case of a plane-wave pump or a short crystal length. In Section 3 we calculate the fidelities of the up-converted images and the pump powers required for up-conversion. We do so by using the example of a quantum image produced via spontaneous parametric down-conversion in a χ(2) crystal pumped by a beam carrying orbital angular momentum [22, 23]. Our proposed up-conversion method is not limited only to that particular type of quantum image and can be equally well applied to other types, e.g., to the quantum images with sub-shot-noise spatial correlations [2426] and continuous-variable entanglement of many spatial modes [27] that have been recently generated. We conclude in Section 4 by summarizing the results.

2. Theory of quantum image conversion

We start with the SFG equations in the undepleted-pump and paraxial approximations with polarized (scalar) fields. Similarly to the derivations of the equations for the inverse process of parametric amplification [28] and using our previous notations [29, 30], we are looking for solutions in the form ei(r,t)=Ei(ρ,z)ei(kizωit)+c.c., where Ei(ρ,z) is the slowly-varying electric-field envelope of the ith wave propagating in z-direction, ρ is a transverse vector with coordinates (x, y), ωi is the angular frequency, ki = ωini/c is the wavevector, ni is the refractive index, and the intensity is given by Ii(ρ,z)=2ε0nic|Ei(ρ,z)|2. Subscript i can take values i = 1, 2, 3 so that ω1 + ω2 = ω3. In the presence of a strong pump E2(ρ,z) at angular frequency ω2, the input-signal field E1(ρ,z) at angular frequency ω1 is coupled to the up-converted field E3(ρ,z) at angular frequency ω3 through the following coupled-wave equations:

{E1(ρ,z)z=i2k1ρ2E1(ρ,z)+2iω1deffn1cE2*(ρ,z)E3(ρ,z)eiΔkFCz,E3(ρ,z)z=i2k3ρ2E3(ρ,z)+2iω3deffn3cE2(ρ,z)E1(ρ,z)eiΔkFCz,
where ΔkFC = k3k1k2 is the wavevector mismatch, deff is the effective nonlinear coefficient, and ρ2=(2/x2+2/y2). Equation (1) describes the traveling-wave up-converter in the paraxial approximation with a pump of arbitrary spatial profile. Since we are interested in the quantum properties of Eq. (1), it is helpful to use normalized input and up-converted fields a1 and a3, respectively, which satisfy commutation relations [a1,a1+]=δ(ρρ), [a3,a3+]=δ(ρρ). Since the normalized fields are related to the original fields by a1,3(ρ,z)n1,3/ω1,3E1,3(ρ,z) (where the omitted proportionality constants are not dependent on the frequency of the wave), the nonlinear coupling coefficients of Eq. (1) become symmetric when Eq. (1) is re-written for the normalized fields:
{a1(ρ,z)z=i2k1ρ2a1(ρ,z)+iκ*(ρ,z)a3(ρ,z)eiΔkFCz,a3(ρ,z)z=i2k3ρ2a3(ρ,z)+iκ(ρ,z)a1(ρ,z)eiΔkFCz,
where

κ(ρ,z)=4ω1ω3deff2n1n3c2E2(ρ,z).

Let us define the spatial-frequency (q) domain via the direct and inverse Fourier transforms

a˜1,3(q,z)=a1,3(ρ,z)eiqρdρ,a1,3(ρ,z)=a˜1,3(q,z)eiqρdq(2π)2.

In the absence of the pump (E2 = 0), Eq. (1) is reduced to the paraxial Helmholtz equation, whose solution for the normalized fields in the Fourier domain is given by

a˜1,3(q,z)=a˜1,3(q,0)exp(iq22k1,3z),
where q=|q|. In general, similarly to the case of optical parametric amplification, Eq. (2) has to be solved numerically in order to account for coupling among the various spatial modes due to the effects akin to gain-induced diffraction [31]. Some semi-analytical methods mirroring those used in solving spatially-broadband parametric-amplifier equations [3236] can be applied here as well. In this paper, however, we will restrict ourselves to the two particular cases that admit analytical solutions: the case of a plane-wave pump and the case of a short crystal with arbitrary pump.

Analytical solution 1: plane-wave pump

With the plane-wave pump E2(ρ,z)=|E2|eiθ2=const., Eq. (1) still retains the shift-invariance of the original paraxial Helmholtz equation and hence it is easily solved in the Fourier domain:

a˜1(q,z)=eiq22k1z[t(q)a˜1(q,0)+r(q)a˜3(q,0)],a˜3(q,z)=eiq22k3z[r*(q)a˜1(q,0)+t*(q)a˜3(q,0)],
where the coefficients of the above unitary input-output transformation are
t(q)=(cosγziΔkeff2γsinγz)×exp(iΔkeff2z),r(q)=iκ*γsinγz×exp(iΔkeff2z),
the effective wavevector mismatch is
Δkeff=ΔkFC+q22(1k11k3),
and the parametric gain coefficient γ is given by

γ=|κ|2+Δkeff2/4.

Equation (6) couples the input and up-converted fields of the same spatial frequency q. This transformation is equivalent to that of a beam-splitter with reflectance (up-conversion efficiency)

R=|r|2=|κ|2γ2sin2γz
and transmittance

T=|t|2=1|κ|2γ2sin2γz=1R.

The evolution of the quantum field operators also obeys Eq. (6), which leads to quantum correlations between the same spatial-frequency components of the input and the up-converted images. The correlations in the ρ-domain, on the other hand, are not ideal, because the limited spatial bandwidth of the up-conversion process, determined by the wavevector mismatch of Eq. (8), results in image distortion. The spatial frequencies, for which the mismatch in (8) is nearly zero, can be up-converted with near-unity efficiency R, whereas those outside the up-conversion spatial bandwidth experience very low conversion efficiency. For input images whose quantum features are contained within the spatial bandwidth of the up-conversion process, the output quantum image will reproduce all the information from the input with high fidelity. And vice versa, for images with broader spatial frequency content, the non-unity conversion efficiency R is equivalent to loss and results in reduction of the fidelity of the up-converted state. We also note that the up-conversion takes place equally for both quadratures of the image, and no underlying symmetry of the image is required (this is in contrast to the phase-sensitive amplification process, which requires symmetry between the signal’s positive- and negative-spatial-frequency components, leading to a flat phase front of the image [7,30,35]).

Analytical solution 2: short crystal with inhomogeneous pump

For sufficiently short crystals, the diffraction terms in Eqs. (1) and (2) can be neglected, and the coupled-wave equations take the following form:

{a1(ρ,z)z=iκ*(ρ)a3(ρ,z)eiΔkFCz,a3(ρ,z)z=iκ(ρ)a1(ρ,z)eiΔkFCz.

Here κ, defined in (3), is, in general, a complex parameter that depends on the coordinate ρ (if the pump is inhomogeneous) and, at the same time, due to the short crystal length, the z-dependence of the pump is neglected. One can then introduce parameters t and r as

t(ρ,z)=(cosγziΔkFC2γsinγz)×exp(iΔkFC2z),r(ρ,z)=iκ*γsinγz×exp(iΔkFC2z),
where
γ(ρ)=|κ|2(ρ)ΔkFC2/4,|κ|2(ρ)=2ω1ω3deff2I2(ρ)ε0n1n3n2c3,
so that the solution takes the form of point-by-point (pixel-by-pixel) up-conversion with efficiency R = |r|2:
a1(ρ,z)=t(ρ,z)a1(ρ,0)+r(ρ,z)a3(ρ,0),a3(ρ,z)=r*(ρ,z)a1(ρ,0)+t*(ρ,z)a3(ρ,0),
which is equivalent to the beam-splitter transformation.

3. Example of quantum image conversion

In this section, we will provide an example of frequency conversion of the quantum image [22, 23, 37] produced via the frequency-degenerate spontaneous parametric down-conversion (SPDC) process of type-I with spatially-modulated z-propagating pump Ep(x, y) of frequency ωp = 2ω1 (here we use subscript “p” to distinguish the SPDC pump from the up-conversion pump denoted by subscript “2”). Assuming that the SPDC crystal length lSPDC is much smaller than the Rayleigh distance of the pump, the SPDC state can be approximated for low pump intensities (i.e., neglecting multiple-pair generation) by an unnormalized wavefunction

|ψ=|0+dqsdqiF(qs,qi)|1qs|1qi,
where qs and qi are the transverse components (spatial frequencies) of the signal and idler wavevectors ks and ki, respectively, and |1qs|1qi denotes a state having one photon with spatial frequency qs and one photon with spatial frequency qi, with all other modes in the vacuum state. F is the joint two-photon wavefunction in the momentum space given by
F(qs,qi)drEp(r)eiΔkr=dzeiΔkzzdxdyEp(x,y)ei(Δkxx+Δkyy)=lSPDCeiΔkzlSPDC/2sinc(ΔkzlSPDC2)E˜p(qs+qi),
where the z-integral gives the phase-mismatch factor, and the (x,y)-integral gives the spatial Fourier transform E˜p(q) of the transverse spatial profile of the pump, with
Δk=kpkski=z^Δkzqsqiz^(kpkski+qs22ks+qi22ki)qsqiz^qs2+qi22q022ksqsqi,
qs,i2=(kx)s,i2+(ky)s,i2 and q02=ks(kp2ks). In the last line of Eq. (18) we assumed kiks.

The peculiar properties of the state of Eq. (16), obtained with orbital-momentum-carrying pump, are illustrated in the far-field (transverse momentum space) diagram shown in Fig. 1(a) , where pump profile of Laguerre-Gaussian mode LG02 with orbital momentum l = 2 is shown in purple as an example. For non-collinear phase matching (i.e., for kp ≠ 2ks), the SPDC photons are emitted in the pattern of a phase-matched cone [red ring of radius q0 in Fig. 1(a)]. The image obtained by direct detecting the SPDC simply reflects this ring shape [see insert 1 in Fig. 1(a)]. The coincidence count measurement, on the other hand, reveals the underlying quantum structure of this image. In particular, if an idler photon at spatial frequency qi0 is detected (heralded) by a small-area detector [yellow dot in Fig. 1(a)], then the probability amplitude of finding a signal photon with spatial frequency qs (or, in other words, the spatial profile of the resulting heralded single-photon mode) will be proportional to the function F(qs,qi0) from Eq. (17). The magnitude of this function is a product of the spatial profile of the phase-matching condition [red ring in Fig. 1(a)] and the far-field pattern of the SPDC pump centered at qi0 [checkered purple shape in Fig. 1(a), which has the profile of a ring if the pump field is given by LG02]. The resulting product is sketched by the two green dots in Fig. 1(a), while its actual calculated shape is shown by the insert 3 in the same Figure. If the pump is a simple fundamental Gaussian beam LG00, this product takes the form shown by the insert 2. Thus, the spatial mode of the heralded single-photon state of the signal depends on the pump profile. This property is only observable in coincidence counting and represents a quantum image.

 figure: Fig. 1

Fig. 1 (a) Illustration of the phase matching, as well as the quantum image generation and up-conversion in the far field (transverse momentum space) when both the SPDC pump and the up-conversion pump are propagating in the same direction. (b) A schematic of the corresponding experimental setup. Inserts 2 and 3 in (a) show the quantum image function |F(qs,qi0)|2 for the cases of SPDC pumping by LG00 and LG02 modes, respectively, and insert 1 is the image obtained by scanning a single photon-counting detector across the output plane in the far field. All three inserts are equally applicable to the configuration shown in Fig. 2. SPDC–spontaneous parametric down-conversion, FC–frequency up-conversion.

Download Full Size | PDF

Next, we study the up-conversion of the SPDC state of Eq. (16). Since the underlying quantum image is revealed in coincidence counting, it is convenient to use the same heralding procedure as that described above to check the fidelity of the up-converted quantum image (we use this procedure for convenience of calculation only; heralding is not required for up-conversion scheme itself to work). In order to up-convert the quantum image given by Eq. (16), we will assume that the output of the SPDC crystal is imaged into the input of the up-conversion crystal of length lFC by a 1:1 telescope and that the up-conversion pump is a plane wave (i.e., the conversion obeys the formulas for Analytical Solution 1 in Section 2), propagating in the same direction as the SPDC pump, as depicted in Fig. 1(b). The mode at frequency ω3 is in the vacuum state at the input of the up-converter. For a single-photon input at frequency ω1 and spatial frequency qs or qi, the input-output transformation (6) performs a beam-splitting procedure on signal and idler modes independently, which takes the following form in Schrödinger picture:

|Ψqs,i=t*(qs,i)|10qs,ir*(qs,i)|01qs,i,
wherein the transformation coefficients t(qs,i),r(qs,i) are given by Eq. (7) with z = lFC, and |10q (or |01q) refers to a state having one photon with spatial frequency q and frequency ω1 (or ω3, respectively), with all other modes in a vacuum state. Thus, the SPDC state (16) after up-conversion becomes:

|ψFC=|0+dqsdqiF(qs,qi)t*(qs)t*(qi)|10qs|10qi+dqsdqiF(qs,qi)r*(qs)r*(qi)|01qs|01qidqsdqiF(qs,qi)r*(qs)t*(qi)|01qs|10qidqsdqiF(qs,qi)t*(qs)r*(qi)|10qs|01qi.

In Eqs. (19) and (20) we have omitted the non-essential diffractive phase q2z/(2k3), which is equivalent to observing the far-field pattern in the Fourier plane of a lens.

Upon detection of the idler photon with transverse wavevector qi0, the up-converted state can be reduced to a heralded up-converted state

|ψHeralded,FC=dqsF(qs,qi0)[t*(qs)|10qsr*(qs)|01qs].

The fidelity η of the frequency up-converted quantum image can be quantified by the overlap of the heralded states obtained with ideal and non-ideal frequency conversion (the former corresponding to Eq. (21) with r = 1 and t = 0 and denoted simply “Heralded” below; it is also the same as the ideal up-conversion of the state heralded prior to frequency conversion):

η=|Heraldedψ|ψHeralded,FC|2Heraldedψ|ψHeraldedHeralded,FCψ|ψHeralded,FC=|dqs|F(qs,qi0)|2r*(qs)|2[dqs|F(qs,qi0)|2]2,
which is, in turn, determined by the overlap of the two-photon wavefunction F and the up-conversion coefficient r* in the transverse momentum space. Thus, to maximize the fidelity of the up-converted quantum image, one needs to ensure that r* is close to unity over the range of spatial frequencies where F has non-zero values. We note that the fidelity is the same regardless of whether the idler photon is detected before or after up-conversion (however, the probability of detection of the idler photon will be lower after imperfect up-conversion).

To achieve appreciable count rates, it makes sense to place the idler detector in the vicinity of qi0 = q0 (which corresponds to Δkz = 0 in the plane-wave pump case), so we will assume qi0=x^q0. Then, the up-conversion should be set up so that Δkeff = 0 takes place at q = qs = q0, i.e.,

ΔkFC=q022(1k11k3)<0,Δkeff=qs2q022(1k11k3),
and k1=ks. Thus, the phase-matched bands of both the up-conversion and the SPDC are centered at |qs|=q0, and to achieve the high fidelity we just need to ensure that the spatial bandwidth of the up-conversion [represented by the half-width of the blue ring in Fig. 1(a)] is greater than that of the SPDC [represented by the half-width of the red ring in Fig. 1(a)]. To estimate the former, we note that at |qs|=q0, r* = 1 for γlFC = |κ|lFC = π/2, i.e. lFC = π/(2|κ|). The first zero of r* in Eq. (7) takes place at |qs|=q0+ΔqcFC so that γlFC = π (i.e., Δkeff=π3/lFC), where the detuning from the center to the first zero quantifies the spatial bandwidth of the up-conversion:

ΔqcFC=π3q0lFCk1k3k3k1.

Similarly, the spatial bandwidth of the SPDC ΔqcSPDC can be estimated from the first zero of the sinc function in Eq. (17) by using Δkz from Eq. (18) with qi = q0 and |qs|=q0+ΔqcSPDC:

ΔqcSPDC=2πksq0lSPDC=ΔqcFC23lFCnslSPDCn1(1k1k3)=ΔqcFC2ω2lFCns3ω3lSPDCn1.

High fidelity is realized when ΔqcFC>ΔqcSPDC, i.e.,

lSPDClFC>2ω2ns3ω3n1,
where ns / n1 is the ratio of the signal-wave refractive indices in the SPDC and up-conversion crystals.

It is worth noting that one can also try an alternative configuration of the up-conversion pumping, where the up-conversion pump wavevector is collinear with the ki02qi0 direction, which corresponds to the direction of the signal SPDC photon in the plane-wave-pump approximation. The corresponding phase-matching diagram in the far field and the up-conversion setup are shown in Figs. 2(a) and 2(b), respectively. The SPDC state is still the same as that in Fig. 1, but at the up-converted stage r*(qs) in Eq. (22) should be replaced by r*(q), where q=qs+qi0. Equation (23) should be replaced by

 figure: Fig. 2

Fig. 2 (a) Illustration of the phase matching, as well as the quantum image generation and up-conversion in the far field (transverse momentum space) when the up-conversion pump is propagating along the ki02qi0 direction. (b) A schematic of the corresponding experimental setup. For SPDC pumping by LG02, the quantum image function |F(qs,qi0)|2 is proportional to the product of the red and checkered-purple rings in (a), which is sketched as two green dots. The single- and coincidence-count images for SPDC pumping by LG00 and LG02 modes are still represented by inserts 1–3 in Fig. 1. SPDC–spontaneous parametric down-conversion, FC–frequency up-conversion.

Download Full Size | PDF

ΔkFC=0,Δkeff=q22(1k11k3).

Then, the spatial bandwidth of the up-conversion in this configuration [radius of the blue circle in Fig. 2(a)] is still obtained from the condition Δkeff=π3/lFC and is given by

(ΔqcFC)=23πlFCk1k3(k3k1)=ΔqcFC2q0ΔqcFC>>ΔqcFC,
i.e., the condition (ΔqcFC)>ΔqcSPDC is easier to satisfy (it requires lower up-conversion pump power) than the condition ΔqcFC>ΔqcSPDC in the previous configuration. However, the alternative configuration is subject to an additional condition that (ΔqcFC)>Δqp, where Δqp is the spatial bandwidth of the SPDC pump beam [i.e., the blue circle in Fig. 2(a) should be larger than the checkered purple ring]. For the quantum images of interest, Δqp>ΔqcSPDC and, therefore, the requirement on the up-conversion spatial bandwidth to be greater than the pump spatial bandwidth becomes dominant for high fidelity, and may not be easy to satisfy. Note that in the original configuration of Fig. 1 (where the SPDC and up-conversion pumps are collinear) the pump spatial bandwidth does not significantly impact the fidelity as long as Eq. (26) is satisfied.

In the calculations of the spatial bandwidth and the fidelity for the configurations of both Fig. 1 (up-conversion pump co-propagating with the SPDC pump) and Fig. 2 (up-conversion pump propagating in the direction of the SPDC signal), two dimensionless parameters are of great importance. The first is the ratio of the center spatial frequency q0 to the SPDC spatial bandwidth ΔqcSPDC from Eq. (25):

N=q0ΔqcSPDC=q02lSPDC2πks.

The second is related to the square root of the ratio of the SPDC pump’s Rayleigh distance zR = kpa0p2 (where a0p is the 1/e intensity radius at the pump’s waist) to lSPDC:

M=2πzRlSPDC.

The SPDC spatial bandwidth given by Eq. (25) is the spatial bandwidth in the so-called “bandpass” configuration of the optical parametric process, where the signal and idler beams are centered at non-zero spatial frequency q0. It is also helpful to calculate the spatial bandwidth (at the first zero of the sinc function) of the signal for the “lowpass” configuration (i.e., assuming qi = q0 = 0):

ΔqcLP-SPDC=4πkslSPDC,
which allows us to express the dimensionless parameters M and N as
M=a0pΔqcLP-SPDC,N=2(q0ΔqcLP-SPDC)2,
and

a0pq0=MN2.

The dependence of the up-conversion fidelity upon the pump’s orbital angular momentum l, the parameter M, and the ratio of the up-conversion to the SPDC crystal lengths lFC / lSPDC is shown in Fig. 3 for ω2 / ω3 = 2/3 and N = 10. The results indicate that high fidelity can be achieved in all cases by having a sufficiently short lFC. However, because shortening lFC requires increasing the up-conversion pump power, the plots show that the configuration of Fig. 2 is clearly preferable as it achieves higher fidelity for a given up-conversion crystal length.

 figure: Fig. 3

Fig. 3 The fidelity of the up-converted quantum image as a function of the ratio between the lengths of the up-conversion and SPDC crystals. Black dash-dotted traces correspond to the configuration of Fig. 1, whereas colored solid and dashed traces correspond to the configuration of Fig. 2. The cases of orbital momenta l = 0, l = 2, and l = 4 correspond to the SPDC pumping by LG00, LG02, and LG04 modes, respectively. We do not show the black traces with M = 3, because they are almost indistinguishable from the black traces with M = 4.

Download Full Size | PDF

Let us consider a specific case of using λ2 = λp = 780 nm, λ1 = 2λ2 = 1560 nm, and the up-converted wavelength λ3 = (1/λ1 + 1/λ2)–1 = 2λ2/3 = λ1/3 = 520 nm. Both SPDC and wavelength conversion can be implemented, for example, in PPLN crystals (ns = n1n2 = npn3 ≈2.14). The (one-sided) angle at which the signal beam emerges in the free space is θs = q0ns / ks. Assuming the parameters θs = 0.04 rad, M = 4, and N = 10 — which are similar to those in [23] (θs = 0.07 rad, M = 4, and N = 8.3), but at a different signal wavelength — we obtain lSPDC = N λs ns / θs2 = 20.9 mm and a0p = (λp lSPDC / np)1/2M/(2π) = 55.5 μm. Then, in order to achieve the fidelity η = 99% with the configuration of Fig. 2, one needs to employ the up-conversion crystals with lengths lFC = 16.7 mm, 6.26 mm, and 4.17 mm for the cases of SPDC pumping by the LG00, LG02, and LG04 modes, respectively. This corresponds to the peak up-conversion pump intensities of 2.4, 17.2, and 38.7 MW/cm2, respectively, where all three intensity numbers are well below the ~250-MW/cm2 optical damage threshold of the lithium-niobate crystal. In contrast to the case of Fig. 2, the configuration of Fig. 1 requires significantly shorter lengths lFC of 2.5 mm, 1.67 mm, and 1.46 mm to achieve the same fidelity, taking the up-conversion pump requirements to the intensity levels of 108, 241, and 318 MW/cm2, respectively, where only the first number is safely below the crystal damage threshold. While the configuration of Fig. 2 requires lower pump intensities, it also makes it more difficult to spatially separate the up-converted image from the up-conversion pump. For our chosen example, where the up-conversion pump’s near-infrared wavelength lies within the spectral response of silicon detectors, a strong dichroic filtering might be needed to prevent the pump beam from affecting the signal detection.

4. Summary

We have developed the theory of frequency up-conversion of an arbitrary spatially-broadband quantum state (quantum image), with analytical solutions derived for the cases of a) plane-wave pump and b) short crystal with arbitrary pump. Using an example of the quantum image imposed by the orbital angular momentum of the pump beam in spontaneous parametric down-conversion, we have obtained the up-conversion fidelities in two up-converter configurations: one where the SPDC and up-conversion pumps co-propagate in the respective crystals and the other where the up-conversion pump co-propagates with the signal beam in the up-conversion crystal. We have shown that the former configuration requires unreasonably high levels of pump intensity, whereas in the latter configuration, 99%-fidelity quantum image up-conversion from 1560 nm to 520 nm wavelength can be obtained in PPLN crystals at moderate up-conversion pump intensities well below the PPLN damage threshold. This indicates the practicality of the quantum image up-conversion with readily available solid-state or fiber-based pump sources.

Acknowledgments

This work was supported in part by DARPA’s Quantum Sensor Program under AFRL Contract No. FA8750-09-C-0194. Any opinions, findings, conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of DARPA or the U.S. Air Force.

References and links

1. C. M. Caves, “Quantum limits on noise in linear amplifiers,” Phys. Rev. D Part. Fields 26(8), 1817–1839 (1982). [CrossRef]  

2. D. Levandovsky, M. Vasilyev, and P. Kumar, “Amplitude squeezing of light by means of a phase-sensitive fiber parametric amplifier,” Opt. Lett. 24(14), 984–986 (1999). [CrossRef]   [PubMed]  

3. D. Levandovsky, M. Vasilyev, and P. Kumar, ““Near-noiseless amplification of light by a phase-sensitive fibre amplifier,” PRAMANA–,” J. Phys. 56, 281 (2001).

4. W. Imajuku, A. Takada, and Y. Yamabayashi, “Low-noise amplification under the 3 dB noise figure in high-gain phase-sensitive fibre amplifier,” Electron. Lett. 35(22), 1954 (1999). [CrossRef]  

5. R. Slavík, F. Parmigiani, J. Kakande, C. Lundstrom, M. Sjodin, P. A. Andrekson, R. Weerasuriya, S. Sygletos, A. D. Ellis, L. Gruner-Nielsen, D. Jakobsen, S. Herstrom, R. Phelan, J. O'Gorman, A. Bogris, D. Syvridis, S. Dasgupta, P. Petropoulos, and D. J. Richardson, “All-optical phase and amplitude regenerator for next-generation telecommunications systems,” Nat. Photonics 4(10), 690–695 (2010). [CrossRef]  

6. Z. Tong, C. Lundstrom, P. A. Andrekson, C. J. McKinstrie, M. Karlsson, D. J. Blessing, E. Tipsuwannakul, B. J. Puttnam, H. Toda, and L. Gruner-Nielsen, “Towards ultrasensitive optical links enabled by low-noise phase-sensitive amplifiers,” Nat. Photonics 5(7), 430–436 (2011). [CrossRef]  

7. M. Kolobov, “The spatial behavior of nonclassical light,” Rev. Mod. Phys. 71(5), 1539–1589 (1999). [CrossRef]  

8. S.-K. Choi, M. Vasilyev, and P. Kumar, “Noiseless optical amplification of images,” Phys. Rev. Lett. 83(10), 1938–1941 (1999). [CrossRef]  

9. K. Wang, G. Yang, A. Gatti, and L. Lugiato, “Controlling the signal-to-noise ratio in optical parametric image amplification,” J. Opt. B Quantum Semiclassical Opt. 5(4), S535–S544 (2003). [CrossRef]  

10. A. Mosset, F. Devaux, and E. Lantz, “Spatially noiseless optical amplification of images,” Phys. Rev. Lett. 94(22), 223603 (2005). [CrossRef]   [PubMed]  

11. E. Lantz and F. Devaux, “Parametric amplification of images: from time gating to noiseless amplification,” J. Sel. Top. Quantum Electron. 14(3), 635–647 (2008). [CrossRef]  

12. P. Kumar, V. Grigoryan, and M. Vasilyev, “Noise-free amplification: towards quantum laser radar,” the 14th Coherent Laser Radar Conference, Snowmass, CO, July 2007. http://space.hsv.usra.edu/CLRC/presentations/Kumar.ppt

13. J. E. Midwinter, “Image conversion from 1.6 μm to the visible in lithium niobate,” Appl. Phys. Lett. 12(3), 68–70 (1968). [CrossRef]  

14. J. M. Huang and P. Kumar, “Observation of quantum frequency conversion,” Phys. Rev. Lett. 68(14), 2153–2156 (1992). [CrossRef]   [PubMed]  

15. H. J. McGuinness, M. G. Raymer, C. J. McKinstrie, and S. Radic, “Quantum frequency translation of single-photon states in a photonic crystal fiber,” Phys. Rev. Lett. 105(9), 093604 (2010). [CrossRef]   [PubMed]  

16. M. T. Rakher, L. Ma, O. T. Slattery, X. Tang, and K. Srinivasan, “Quantum transduction of telecommunications-band single photons from a quantum dot by frequency upconversion,” Nat. Photonics 4(11), 786–791 (2010). [CrossRef]  

17. A. G. Radnaev, Y. O. Dudin, R. Zhao, H. H. Jen, S. D. Jenkins, A. Kuzmich, and T. A. B. Kennedy, “A quantum memory with telecom-wavelength conversion,” Nat. Phys. 6(11), 894–899 (2010). [CrossRef]  

18. P. Scotto, P. Colet, A. Jacobo, and M. S. Miguel, “Optical image processing in second-harmonic generation,” chapter 8 in Quantum Imaging, ed. by M. Kolobov, Springer Verlag, New York, 2007.

19. L. V. Magdenko, I. V. Sokolov, and M. I. Kolobov, “Quantum teleportation of optical images with frequency conversion,” Opt. Spectrosc. 103(1), 62–66 (2007). [CrossRef]  

20. A. S. Chirkin and E. V. Makeev, “Simultaneous phase-sensitive parametric amplification and up-conversion of an optical image,” J. Opt. B Quantum Semiclassical Opt. 7(12), S500–S506 (2005). [CrossRef]  

21. E. V. Makeev and A. S. Chirkin, “Quantum fluctuations of parametrically amplified and up-converted optical images in consecutive wave interactions,” J. Russ. Laser Res. 27(5), 466–474 (2006). [CrossRef]  

22. G. A. Barbosa and H. H. Arnaut, “Twin photons with angular-momentum entanglement: phase matching,” Phys. Rev. A 65(5), 053801 (2002). [CrossRef]  

23. A. R. Altman, K. G. Köprülü, E. Corndorf, P. Kumar, and G. A. Barbosa, “Quantum imaging of nonlocal spatial correlations induced by orbital angular momentum,” Phys. Rev. Lett. 94(12), 123601 (2005). [CrossRef]   [PubMed]  

24. M. L. Marable, S.-K. Choi, and P. Kumar, “Measurement of quantum-noise correlations in parametric image amplification,” Opt. Express 2(3), 84–92 (1998). [CrossRef]   [PubMed]  

25. O. Jedrkiewicz, Y.-K. Jiang, E. Brambilla, A. Gatti, M. Bache, L. A. Lugiato, and P. Di Trapani, “Detection of sub-shot-noise spatial correlation in high-gain parametric down conversion,” Phys. Rev. Lett. 93(24), 243601 (2004). [CrossRef]   [PubMed]  

26. J. L. Blanchet, F. Devaux, L. Furfaro, and E. Lantz, “Measurement of sub-shot-noise correlations of spatial fluctuations in the photon-counting regime,” Phys. Rev. Lett. 101(23), 233604 (2008). [CrossRef]   [PubMed]  

27. V. Boyer, A. M. Marino, R. C. Pooser, and P. D. Lett, “Entangled images from four-wave mixing,” Science 321(5888), 544–547 (2008). [CrossRef]   [PubMed]  

28. A. Gavrielides, P. Peterson, and D. Cardimona, “Diffractive imaging in three-wave interactions,” J. Appl. Phys. 62(7), 2640–2645 (1987). [CrossRef]  

29. M. Vasilyev, N. Stelmakh, and P. Kumar, “Phase-sensitive image amplification with elliptical Gaussian pump,” Opt. Express 17(14), 11415–11425 (2009), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-17-14-11415. [CrossRef]   [PubMed]  

30. M. Vasilyev, N. Stelmakh, and P. Kumar, “Estimation of the spatial bandwidth of an optical parametric amplifier with plane-wave-pump,” J. Mod. Opt. 56(18-19), 2029–2033 (2009). [CrossRef]  

31. S.-K. Choi, R.-D. Li, C. Kim, and P. Kumar, “Traveling-wave optical parametric amplifier: investigation of its phase-sensitive and phase-insensitive gain response,” J. Opt. Soc. Am. B 14(7), 1564 (1997). [CrossRef]  

32. C. Schwob, P. F. Cohadon, C. Fabre, M. A. M. Marte, H. Ritsch, A. Gatti, and L. Lugiato, “Transverse effects and mode couplings in OPOs,” Appl. Phys. B 66(6), 685–699 (1998). [CrossRef]  

33. K. G. Köprülü and O. Aytür, “Analysis of Gaussian-beam degenerate optical parametric amplifiers for the generation of quadrature-squeezed states,” Phys. Rev. A 60(5), 4122–4134 (1999). [CrossRef]  

34. K. G. Köprülü and O. Aytür, “Analysis of the generation of amplitude-squeezed light with Gaussian-beam degenerate optical parametric amplifiers,” J. Opt. Soc. Am. B 18(6), 846 (2001). [CrossRef]  

35. M. Vasilyev, M. Annamalai, N. Stelmakh, and P. Kumar, “Quantum properties of a spatially-broadband traveling-wave phase-sensitive optical parametric amplifier,” J. Mod. Opt. 57(19), 1908–1915 (2010). [CrossRef]  

36. M. Annamalai, M. Vasilyev, N. Stelmakh, and P. Kumar, “Compact representation of spatial modes of phase-sensitive image amplifier,” the Conference on Lasers and Electro-Optics, Baltimore, MD, May 1–6, 2011, paper JThB77.

37. K. G. Köprülü, Y.-P. Huang, G. A. Barbosa, and P. Kumar, “Lossless single-photon shaping via heralding,” Opt. Lett. 36(9), 1674–1676 (2011). [CrossRef]   [PubMed]  

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 (3)

Fig. 1
Fig. 1 (a) Illustration of the phase matching, as well as the quantum image generation and up-conversion in the far field (transverse momentum space) when both the SPDC pump and the up-conversion pump are propagating in the same direction. (b) A schematic of the corresponding experimental setup. Inserts 2 and 3 in (a) show the quantum image function |F( q s , q i0 ) | 2 for the cases of SPDC pumping by LG00 and LG02 modes, respectively, and insert 1 is the image obtained by scanning a single photon-counting detector across the output plane in the far field. All three inserts are equally applicable to the configuration shown in Fig. 2. SPDC–spontaneous parametric down-conversion, FC–frequency up-conversion.
Fig. 2
Fig. 2 (a) Illustration of the phase matching, as well as the quantum image generation and up-conversion in the far field (transverse momentum space) when the up-conversion pump is propagating along the k i0 2 q i0 direction. (b) A schematic of the corresponding experimental setup. For SPDC pumping by LG02, the quantum image function |F( q s , q i0 ) | 2 is proportional to the product of the red and checkered-purple rings in (a), which is sketched as two green dots. The single- and coincidence-count images for SPDC pumping by LG00 and LG02 modes are still represented by inserts 1–3 in Fig. 1. SPDC–spontaneous parametric down-conversion, FC–frequency up-conversion.
Fig. 3
Fig. 3 The fidelity of the up-converted quantum image as a function of the ratio between the lengths of the up-conversion and SPDC crystals. Black dash-dotted traces correspond to the configuration of Fig. 1, whereas colored solid and dashed traces correspond to the configuration of Fig. 2. The cases of orbital momenta l = 0, l = 2, and l = 4 correspond to the SPDC pumping by LG00, LG02, and LG04 modes, respectively. We do not show the black traces with M = 3, because they are almost indistinguishable from the black traces with M = 4.

Equations (33)

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

{ E 1 ( ρ ,z) z = i 2 k 1 ρ 2 E 1 ( ρ ,z)+ 2i ω 1 d eff n 1 c E 2 * ( ρ ,z) E 3 ( ρ ,z) e iΔ k FC z , E 3 ( ρ ,z) z = i 2 k 3 ρ 2 E 3 ( ρ ,z)+ 2i ω 3 d eff n 3 c E 2 ( ρ ,z) E 1 ( ρ ,z) e iΔ k FC z ,
{ a 1 ( ρ ,z) z = i 2 k 1 ρ 2 a 1 ( ρ ,z)+i κ * ( ρ ,z) a 3 ( ρ ,z) e iΔ k FC z , a 3 ( ρ ,z) z = i 2 k 3 ρ 2 a 3 ( ρ ,z)+iκ( ρ ,z) a 1 ( ρ ,z) e iΔ k FC z ,
κ( ρ ,z)= 4 ω 1 ω 3 d eff 2 n 1 n 3 c 2 E 2 ( ρ ,z).
a ˜ 1,3 ( q ,z)= a 1,3 ( ρ ,z) e i q ρ d ρ , a 1,3 ( ρ ,z)= a ˜ 1,3 ( q ,z) e i q ρ d q (2π) 2 .
a ˜ 1,3 ( q ,z)= a ˜ 1,3 ( q ,0)exp( i q 2 2 k 1,3 z ),
a ˜ 1 ( q ,z)= e i q 2 2 k 1 z [t(q) a ˜ 1 ( q ,0)+r(q) a ˜ 3 ( q ,0)], a ˜ 3 ( q ,z)= e i q 2 2 k 3 z [ r * (q) a ˜ 1 ( q ,0)+ t * (q) a ˜ 3 ( q ,0)],
t(q)=( cosγz iΔ k eff 2γ sinγz )×exp( i Δ k eff 2 z ), r(q)=i κ * γ sinγz×exp( i Δ k eff 2 z ),
Δ k eff =Δ k FC + q 2 2 ( 1 k 1 1 k 3 ),
γ= |κ | 2 +Δ k eff 2 /4 .
R=|r | 2 = |κ | 2 γ 2 sin 2 γz
T=|t | 2 =1 |κ | 2 γ 2 sin 2 γz =1R.
{ a 1 ( ρ ,z) z =i κ * ( ρ ) a 3 ( ρ ,z) e iΔ k FC z , a 3 ( ρ ,z) z =iκ( ρ ) a 1 ( ρ ,z) e iΔ k FC z .
t( ρ ,z)=( cosγz iΔ k FC 2γ sinγz )×exp( i Δ k FC 2 z ), r( ρ ,z)=i κ * γ sinγz×exp( i Δ k FC 2 z ),
γ( ρ )= |κ | 2 ( ρ )Δ k FC 2 /4 ,|κ | 2 ( ρ )= 2 ω 1 ω 3 d eff 2 I 2 ( ρ ) ε 0 n 1 n 3 n 2 c 3 ,
a 1 ( ρ ,z)=t( ρ ,z) a 1 ( ρ ,0)+r( ρ ,z) a 3 ( ρ ,0), a 3 ( ρ ,z)= r * ( ρ ,z) a 1 ( ρ ,0)+ t * ( ρ ,z) a 3 ( ρ ,0),
|ψ=|0+ d q s d q i F( q s , q i ) |1 q s |1 q i ,
F( q s , q i ) d r E p ( r ) e iΔ k r = dz e iΔ k z z dxdy E p (x,y) e i(Δ k x x+Δ k y y) = l SPDC e iΔ k z l SPDC /2 sinc( Δ k z l SPDC 2 ) E ˜ p ( q s + q i ),
Δ k = k p k s k i = z ^ Δ k z q s q i z ^ ( k p k s k i + q s 2 2 k s + q i 2 2 k i ) q s q i z ^ q s 2 + q i 2 2 q 0 2 2 k s q s q i ,
|Ψ q s,i = t * ( q s,i ) | 10 q s,i r * ( q s,i ) | 01 q s,i ,
|ψ FC =|0+ d q s d q i F( q s , q i ) t * ( q s ) t * ( q i ) | 10 q s | 10 q i + d q s d q i F( q s , q i ) r * ( q s ) r * ( q i ) | 01 q s | 01 q i d q s d q i F( q s , q i ) r * ( q s ) t * ( q i ) | 01 q s | 10 q i d q s d q i F( q s , q i ) t * ( q s ) r * ( q i ) | 10 q s | 01 q i .
|ψ Heralded, FC = d q s F( q s , q i0 ) [ t * ( q s ) | 10 q s r * ( q s ) | 01 q s ].
η= | Heralded ψ|ψ Heralded,FC | 2 Heralded ψ|ψ Heralded Heralded,FC ψ|ψ Heralded,FC = | d q s |F( q s , q i0 ) | 2 r * ( q s ) | 2 [ d q s |F( q s , q i0 ) | 2 ] 2 ,
Δ k FC = q 0 2 2 ( 1 k 1 1 k 3 )<0, Δ k eff = q s 2 q 0 2 2 ( 1 k 1 1 k 3 ),
Δ q c FC = π 3 q 0 l FC k 1 k 3 k 3 k 1 .
Δ q c SPDC = 2π k s q 0 l SPDC =Δ q c FC 2 3 l FC n s l SPDC n 1 ( 1 k 1 k 3 ) =Δ q c FC 2 ω 2 l FC n s 3 ω 3 l SPDC n 1 .
l SPDC l FC > 2 ω 2 n s 3 ω 3 n 1 ,
Δ k FC =0, Δ k eff = q 2 2 ( 1 k 1 1 k 3 ).
( Δ q c FC ) = 2 3 π l FC k 1 k 3 ( k 3 k 1 ) =Δ q c FC 2 q 0 Δ q c FC >>Δ q c FC ,
N= q 0 Δ q c SPDC = q 0 2 l SPDC 2π k s .
M= 2π z R l SPDC .
Δ q c LP-SPDC = 4π k s l SPDC ,
M= a 0p Δ q c LP-SPDC , N=2 ( q 0 Δ q c LP-SPDC ) 2 ,
a 0p q 0 =M N 2 .
Select as filters


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