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

Effects of transmission on Gaussian optical states

Open Access Open Access

Abstract

The noise properties of phase-insensitive and phase-sensitive optical transmission links are described in detail, for Gaussian input signals. Formulas are derived for the quadrature covariance matrices of the output signals, which allow one to quantify the noise figures of the links and the fidelities of transmission. Another formula is derived, which relates the density operator of an output signal, in the number-state representation, to its covariance matrix. This density matrix allows one to quantify the decrease in coherence and changes in photon-number probabilities associated with transmission. Based on the aforementioned performance metrics, links with distributed phase-sensitive amplification perform significantly better than other links.

© 2015 Optical Society of America

1. Introduction

Gaussian optical states are states whose Wigner functions are Gaussian [14]. Such states include strong and weak coherent states, which are used in conventional communication systems [5, 6] and quantum information science [710], respectively, and the photon-pair-like states produced by parametric sources [11, 12], which are also used in quantum information science [13, 14]. Such functions are specified completely by their mean vectors and covariance matrices, whose elements are quadrature variances and correlations.

Conventional systems can be modeled as sequences of transmission fibers (attenuators) and (linear) amplifiers. As optical signals propagate through such systems (links), noise is imparted to them by the attenuators and amplifiers. The standard measure of signal quality is the signal-to-noise ratio (SNR), which for direct (incoherent) detection is the square of the mean photon-number divided by the number variance and for homodyne (coherent) detection is the square of the mean quadrature divided by the quadrature variance. The standard measure of link performance is the noise figure, which is the input SNR divided by the output SNR, for either mode of detection. Knowledge of the noise figure allows one to estimate the bit-error-ratio associated with a particular signal format [5, 6] and the information capacity (maximal information-transfer rate) of the link [15, 16]. Noise-figure formulas have been calculated for a wide variety of transmission links [1928]. For many links, the noise-noise contributions to the number variance can be neglected, in which case the direct and homodyne noise figures are equal [27], and are related to the input and output covariance matrices.

Many quantum information experiments involve single- or few-photon inputs. For such experiments, a better performance metric is the transmission fidelity [30], which is (roughly) the overlap of the input and output states. For Gaussian input states, the fidelity is also related to the input and output covariance matrices [31].

Although the noise figure and fidelity both characterize aspects of the output state, and are useful in certain contexts, neither metric specifies the state completely. It is difficult to determine the output state directly, because it is mixed and the number of basis states required to represent its density operator accurately increases rapidly with link length (distance). However, it is straightforward to determine the Wigner function of the output state (which is specified by the output covariance matrix), after which the associated density operator can be reconstructed by the evaluation of certain integrals.

This paper is organized as follows: In Sec. 2 the characteristic and Wigner functions used to describe optical states are introduced, and the basic properties of Gaussian states are mentioned briefly. The transmission fidelity is defined in terms of the density operators of the input and output states, and this definition is related to the associated Wigner functions. For Gaussian states, the fidelity formula involves only the input and output covariance matrices. In Sec. 3 the noise properties of transmission links are reviewed in detail for Gaussian input signals. Formulas are derived for the covariance matrices of the output signals, for links with lumped and distributed phase-insensitive (PI) and phase-sensitive (PS) amplification. A brief comparison is made of the noise figures of these links. In Sec. 4 the aforementioned covariance matrices are used to determine the transmission fidelities of one-mode links. The fidelities of two-mode links, which are related to the one-mode fidelities, are discussed briefly in Sec. 5. In Sec. 6 the covariance matrices are used to reconstruct the output density operators in the number-state representation. These density matrices are used to illustrate the decreases in coherence and changes in photon-number probabilities that are caused by transmission. Finally, in Sec. 7 the main results of this paper are summarized.

2. Representations and properties of Gaussian states

Suppose that W(p, q) is a classical probability density function (PDF) of the real variables p and q [29]. Then, if the PDF is normalized, the moments

pmqn=W(p,q)pmqndpdq.
If the PDF is known, the associated characteristic function [29] is defined as the Fourier transform
C(k,l)=W(p,q)exp[i(kp+lq)]dpdq.
The characteristic function is also called the moment generating function, because
pmqn=limk0liml0(ik)m(il)nC(k,l).
Conversely, if the moment generating function is known, the PDF is defined as the inverse transform
W(p,q)=C(k,l)exp[i(kp+lq)]dkdl/(2π)2.
Equation (1) is valid for arbitrary functions of p and q. Hence, a natural way to measure the overlap between two PDFs is to calculate the expectation value of one distribution relative to the other (reference) distribution. (Normalization might be required.)

In quantum mechanics, the characteristic function [24] is defined as the expectation value

C(k,l)=exp[i(kp^+lq^)]=tr{ρ^exp[i(kp^+lq^)]},
where and are quadrature operators and ρ̂ is the density operator of the system. The quadrature operators satisfy the boson commutation relation [, ] = i, and the density operator is hermitian and positive [tr(ρ̂) = 1]. This quantum function can also be used to calculate moments, but the analysis is more complicated because and do not commute. Characteristic function (5) is said to be symetrically ordered, because and appear symmetrically in its definition, and the symmetric moments
(kp^+lq^)n=limθ0(iθ)nC(θk,θl).
Other distribution functions and moment formulas exist, which are based on the expectation values of the asymmetric operators exp(−ikp̂) exp(−ilq̂) and exp(−ilq̂) exp(−ikp̂). The characteristic function (5) is a function of two real variables (k and l), so the definition of the associated distribution function, which is called a Wigner distribution function [14], is identical to the classical definition (4). Although definitions (4) and (5) ensure that the Wigner function is real, it is not a true PDF, because it can have negative values. Nonetheless, it is useful because Eqs. (2) and (6) imply that
(kp^+lq^)n=W(p,q)(kp+lq)ndpdq.
By comparing like powers of k and l on both sides of Eq. (7), one finds that
(p^mq^n)s=W(p,q)pmqndpdq,
where the subscript s denotes symmetric ordering. For example, the second-order moments 2 and 2 are intrinsically symmetric, whereas (p̂q̂)s = (p̂q̂ + q̂p̂)/2. The number operator = (2 + 2 − 1)/2 depends on the aforementioned second-order moments, and its square depends on the symmetric fourth-order moments 4, 4 and 22 + 22 = 2(22)s − 1. Thus, if the Wigner function is known, Eq. (8) allows one to calculate the most important moments in a straightforward manner.

Gaussian states are defined to be ones whose Wigner functions have the form

W(X)=(Δs1/2/2π)exp[(XV)tS(XV)/2],
where X = [p, q]t is the quadrature vector, S is a (symmetric) coefficient matrix and the determinant Δs = det(S). By using the formula
exp(XtAX/2)d2x=2π/Δa1/2,
where A is a positive matrix, one can verify that Wigner function (9) is normalized. One also can verify that the expectation values
X=V,(XV)(XV)t=S1=C.
(Tutorial derivations of these results are provided in [32].) Thus, V is the mean vector and the covariance matrix C is the inverse of the coefficient matrix S. (With the exceptions of this paragraph and the next, covariance matrices will appear with subscripts, so they should not be confused with characteristic functions.) C must have positive eigenvalues, because the principal widths of the Wigner function are positive. The second of Eqs. (11) implies that S also has positive eigenvalues and determinant, as required by Eq. (10). The Wigner function of a Gaussian state is positive (by definition) and can be regarded as a true PDF. Examples of Gaussian states include coherent and squeezed states, which are pure, and thermal states, which are mixed [24].

The Wigner functions of coherent [17] and squeezed [18] states are illustrated in Fig. 1. The Wigner function of the former state is symmetric, whereas the Wigner function of the latter is asymmetric. In Fig. 1(b), the p- and q-fluctuations are uncorrelated (C is diagonal), so the principal axes of the Wigner function are aligned with the quadrature axes. If the fluctuations were correlated (if C was full), the principal axes would be inclined at an angle with respect to the quadrature axes. The Wigner functions of some number states are illustrated in App. A.

 figure: Fig. 1

Fig. 1 Wigner distribution function of (a) a coherent state with a p-quadrature mean of 2.5, which corresponds to a number mean of 3.0, and (b) a state whose p- and q-quadratures have been stretched and squeezed, respectively. Both distributions are Gaussian and positive.

Download Full Size | PDF

Gaussian distributions play an important role in the theory of optical transmission, because they are specified completely by their first two moments (mean vectors and covariance matrices), and because sequences of (linear) attenuators and amplifiers preserve their Gaussian natures [9]. Furthermore, it is relatively easy to calculate the effects of transmission on the aforementioned moments.

In quantum mechanics, the fidelity of two arbitrary states (1 and 2) is defined to be [30]

F12=[tr(ρ^11/2ρ^2ρ^11/2)1/2]2=[tr(ρ^21/2ρ^1ρ^21/2)1/2]2.
If state 1 is pure (ρ̂1 = |ψ1〉 〈ψ1|, where |ψ1〉 is a state vector), then one can use the fact that tr(|a〉 〈b|) = 〈b|a〉 to rewrite the fidelity as 〈ψ1|ρ̂2|ψ1〉 = tr(ρ1ρ2). If state 2 is also pure, then the fidelity is |〈ψ1|ψ2〉|2. Thus, Eq. (12) is a plausible generalization of two natural measures of overlap, which are based on the inner product of two pure states and the expectation value of the density operator of one state relative to the other (pure-reference) state. In optical transmission the input state is usually pure and the output state is mixed, so the intermediate fidelity formula applies.

In most cases, the density operator of the output state is difficult to calculate directly. However, the trace formula [2] asserts that

tr(ρ^1ρ^2)=2πW1(p,q)W2(p,q)dpdq,
where W1 and W2 are the Wigner functions associated with states 1 and 2, respectively. Thus, the fidelity of two states is measured by the expectation value of one Wigner function relative to the other. [See the comment after Eq. (4).] If the states are identical, Eq. (13) implies that the purity tr(ρ̂2) = Δ1/2/2. Hence, a Gaussian state is pure [tr(ρ̂2) = 1] if and only if the determinant of its correlation matrix (D = 1/Δ) is 1/4. This value is associated with minimal uncertainty states, of which vacuum, coherent and squeezed states are examples. Formula (13) is useful because in optical transmission the input Wigner function is known and the output Wigner function is relatively easy to determine.

For two distinct states, Eqs. (9) and (13) imply that

F12=[(Δ1Δ2)1/2/2π]exp[(XV1)tS1(XV1)/2]×exp[(XV2)tS2(XV2)/2]d2x.
One can do the integral in Eq. (14) by completing the square and one can facilitate this process by measuring X relative to (V1 + V2)/2, in which case the exponent depends only on V = (V1V2)/2. Observe that
(XV)tS1(XV)+(X+V)tS2(X+V)=Xt(S1+S2)XXt(S1S2)VVt(S1S2)X+Vt(S1+S2)V=XtS+XXtSVVtSX+VtS+V=(XW)tS+(XW)WtS+W+VtS+V=(XW)tS+(XW)+Vt(S+SS+1S)V,
where S± = S1 ± S2 and W=S+1SV. By combining Eqs. (10), (14) and (15), one obtains the fidelity formula
F12=(Δ1Δ2/Δ+)1/2exp[Vt(S+SS+1S)V/2].

Formula (16) involves coefficient matrices and their determinants. In contrast, the Scutaru formula [31]

F12=D121/2exp[(V1V2)C121(V1V2)/2],
where C12 = C1 +C2 and D12 = det(C12), involves covariance matrices and their determinants. Since
det(S1+S2)/det(S1S2)=det(S1+S2)det(S11S21)=det(S11+S21),
the algebraic factors in Eqs. (16) and (17) are equal. In App. B it is shown that
S+SS+1S=4(S11+S21)1,
so the exponential factors also are equal. Thus, formulas (16) and (17) are equivalent.

The preceding results apply to one-mode states. To extend them to d-mode states (which involve 2d real quadratures), one increases the dimensionality of the relevant vectors, matrices and integrals, and makes the replacement 2π → (2π)d in Eqs. (4), (9), (10), (13) and (14). Factors of 2π remain absent from Eqs. (16) and (17).

If the density operator of a one-mode state is known, the characteristic and Wigner functions are determined by Eqs. (4) and (5). Conversely, if the Wigner function is known, the density operator can be reconstructed. It is often convenient to use the number-state representation, in which the density operator

ρ^=mn|mρmnn|,
where the matrix elements ρmn = 〈m|ρ|n〉. The non-diagonal elements provide information about the coherence of the state, whereas the diagonal elements provide information about the distribution of photon numbers within it. Notice that the density matrix is hermitian, like the operator it represents. Equation (13) implies that
ρmn=2πWnm(p,q)W(p,q)dpdq,
where Wnm is the Wigner function associated with the operator |n〉 〈m|. It is shown in Apps. A and C that
Wnm(p,q)={(1)n(2mnn!/m!)1/2(q+ip)mnexp(p2q2)Lnmn(2p2+2q2)/π,(1)m(2nmm!/n!)1/2(qip)nmexp(p2q2)Lmnm(2p2+2q2)/π,
for mn and m < n, respectively, where Lnm(x) is an associated Laguerre polynomial [33]. These formulas imply that Wnm=Wmn*, as required by Eq. (21). For d-mode states one can use a number-state basis for each mode, in which case Eqs. (20)(22) generalize in a straightforward manner: The density matrix involves 2d summations over number states, the matrix element is specified by an integral over 2d quadrature variables, which is preceded by the factor (2π)d, and the Wigner function associated with 2d number states is the product of d Wigner functions of the form (22).

3. Noise properties of transmission links

Optical transmission links are sequences of fibers (attenuators) and amplifiers, and their noise properties were studied in detail [1928]. In this section the properties of one-signal-mode phase-sensitive (PS) links, one-signal-mode phase-insensitive (PI) links and two-signal-mode PS links will be reviewed briefly, for convenience. (Introductions to quantum optics and parametric amplification are provided in [4, 34].)

Let â be the signal-mode amplitude operator, which satisfies the boson commutation relation [â, â] = 1, where square brackets denote a commutator and † denotes a hermitian conjugate. Then the (beam-splitter-like) input–output (IO) relation for attenuation is â′ = τâ + ρŵ, where denotes an output quantity, ŵ is the loss-mode operator, and the transmission coefficient τ and reflection coefficient ρ satisfy the auxiliary equation |τ|2 +|ρ|2 = 1. (Here ρ is a reflection coefficient, not a density operator.) In general, the transmission and reflection coefficients are complex. However, by defining the input and output reference phases judiciously, one can rewrite the IO relation in such a way that only the moduli of these coefficients appear. By defining the real and imaginary signal-mode quadrature operators = (â + â)/21/2 and = i(ââ)/21/2, which satisfy the commutation relation [, ] = i, one obtains the quadrature IO relations

p^=|τ|p^+|ρ|w^p,q^=|τ|q^+|ρ|w^q,
where ŵp = (ŵ + ŵ)/21/2 and ŵq = i(ŵŵ)/21/2 are the loss-mode quadrature operators. Both signal quadratures are diminished by the same amount, but inherit excess noise from the loss-mode quadratures.

The IO relation for one-mode parametric amplification is â′ = μâ + νâ, where μ and ν are gain coefficients that satisfy the auxiliary equation |μ|2 − |ν|2 = 1. In general, the gain coefficients are complex, but by defining the reference phases judiciously, one obtains the quadrature IO relations

p^=(|μ|+|ν|)p^,q^=(|μ||ν|)q^,
which depend on only the moduli of these coefficients. The real (in-phase) quadrature is stretched, whereas the imaginary (out-of-phase) quadrature is squeezed by the same amount, because |μ| − |ν| = 1/(|μ| + |ν|). Neither quadrature inherits excess noise.

The IO relations for two-mode parametric amplification are a^1=μa^1+νa^2 and a^2=μa^2+νa^1, where the subscripts 1 and 2 denote the signal and idler modes, respectively. If the input idler is absent, the output signal and idler operators are proportional to the input signal operator (or its conjugate), and the constants of proportionality are independent of the input signal phase, so the amplification process is PI. Conversely, if the input idler is present, the output operators depend on both input operators, so the amplification process is PS. By defining the reference phases judiciously, one obtains the quadrature IO relations

p^1=|μ|p^1+|ν|p^2,q^1=|μ|q^1|ν|q^2,
p^2=|μ|p^2+|ν|p^1,q^2=|μ|q^2|ν|q^1.
If the real signal and idler quadratures are in-phase (out-of-phase), they combine constructively (destructively), whereas if the imaginary signal and idler quadratures are in-phase (out-of-phase), they combine destructively (constructively). Each output quadrature inherits noise from two input quadratures.

3.1. One-mode phase-sensitive links

Consider the first stage of a one-mode PS link, in which loss is followed by PS gain. By combining Eqs. (23) and (24), one obtains the composite IO relations

p^=(λ+τ)p^+(λ+ρ)w^p,q^=(λτ)q^+(λρ)w^q,
where the dilation coefficients λ± = μ ± ν. Here and subsequently, the gain and loss coefficients are assumed to be real. Equations (27) imply that the output quadrature means 〈p̂′〉 = (λ+τ)〈〉 and 〈q̂′〉 = (λτ)〈〉. Define the quadrature variances vp = 〈2〉 − 〈2 and vq = 〈2〉 − 〈2, and the symmetrically-ordered correlation c = 〈p̂q̂s − 〈〉 〈〉, all of which can be measured by homodyne detection (App. D). Then, by squaring Eqs. (27) and using the fact that the signal and loss-mode fluctuations are uncorrelated, one finds that the output variances
vp=(λ+τ)2vp+(λ+ρ)2σ,vq=(λτ)2vq+(λρ)2σ,
where σ = 1/2 is the (common) vacuum-level variance of the loss-mode quadratures. By combining Eqs. (27), one finds also that c′ = τ2c: If the input quadratures are uncorrelated, so also are the output quadratures.

A balanced stage (v′p = vp) requires that λ+τ = 1, in which case λ = τ and

vp=vp+σ(l1),vq=vq/l2+σ(l1)/l2,
where l = 1/τ2 is the loss factor of the stage. In Eqs. (29) the σ terms represent noise added by the loss process only. One can derive the IO relations for a link with s stages by iterating Eqs. (28). For a balanced link, the results are
vp=vp+σs(l1),vq=vq/l2s+σ(11/l2s)/(l+1).
Equations (30) reduce to Eqs. (29) in the limit s → 1.

We are particularly interested in links with distributed amplification [35, 36], which can be modeled as concatenations of many short stages. For such a stage, λ±21±γδz, ρ2αδz and τ2 ≈ 1 − αδz, where α and γ are loss and gain coefficients, respectively, and δz is the stage length. By making these substitutions in Eqs. (28) and taking the limit as δz → 0, one obtains the differential equations

dzvp=(γα)vp+ασ,dzvq=(γ+α)vq+ασ.
Equations (31) have the solutions
vp(z)=vp(0)exp(λpz)+σα[exp(λpz)1]/λp,
vq(z)=vq(0)exp(λqz)+σα[1exp(λqz)]/λq,
respectively, where the dilation rates λp = γα and λq = γ + α. For a balanced link (γ = α),
vp(z)=vp(0)+σαz,
vq(z)=vq(0)exp(2αz)+σ[1exp(2αz)]/2.
Solutions (34) and (35) are consistent with the results of [24] and the appropriate limits of solutions (30). Notice that the in-phase variance increases only linearly with the loss distance αz, rather than exponentially, and the out-of-phase variance decreases rapidly to the sub-vacuum level σ/2.

3.2. Two-mode phase-insensitive links

Consider the first stage of a two-mode PI link. In the first part of the stage, a signal mode is attenuated. In the second part, the signal interacts with an idler (and a strong pump) and is amplified. The input idler is a vacuum state, so the amplification process is PI. At the end of the stage, the idler is discarded. Although PI amplification is a two-mode process, a two-mode PI link transmits only one signal mode. By combining Eqs. (23) and (25), one obtains the composite IO relations

p^1=(μτ)p^1+(μρ)w^p+νp^2,q^1=(μτ)q^1+(μρ)w^qνq^2.
Notice that the first and second of Eqs. (27) are related by the substitution ν → −ν, so it is sufficient to consider only the p-quadrature explicitly. By squaring the first of Eqs. (36), and using the fact that the idler and loss-mode fluctuations are not correlated with the signal or each other, one obtains the single-stage variance equation
vp=(μτ)2vp+(μ2ρ2+ν2)σ,
where σ is the (common) variance of the idler and loss-mode quadratures. A balanced stage requires that µτ = 1, in which case
vp=vp+2(l1)σ.
By comparing Eqs. (29) and (38), one finds that a PI stage adds twice as much noise to the signal as a PS link, because the idler and loss modes both contribute (equally). One can derive the IO relations for a multiple-stage link by iterating Eq. (37). For a balanced link, the result is
vp=vp+2s(l1)σ.
Equation (39) reduces to Eq. (38) in the limit s → 1.

For a short stage, τ2 ≈ 1 − αδz, ρ2αδz, μ2 ≈ 1 + γδz and ν2γδz. By making these substitutions in Eq. (37) and taking the limit as δz → 0, one obtains the differential equation

dzvp=(γα)vp+(γ+α)σ.
Equation (40) has the solution
vp(z)=vp(0)exp(λz)+σλ+[exp(λz)1]/λ,
where λ± = γ ± α. For a balanced link,
vp(z)=vp(0)+2σαz.
Solution (42) is consistent with the appropriate limit of Eq. (39). Because Eq. (37) depends only on ν2, equations similar to (38)(42) apply to the q-quadrature.

Notice that one can write Eq. (42) in the form of Eq. (38) by setting the loss factor l = 1+αz, which is equivalent to setting μ2 = 1 + αz, ν2 = αz, τ2 = 1/(1 + αz) and ρ2 = αz/(1 + αz). This result is consistent with a theorem in quantum information [37], which states that every PI Gaussian channel (link) is equivalent to a lumped attenuator followed by a lumped PI amplifier, with suitably chosen loss and gain parameters. It should be emphasized that in this example, the loss parameter depends only linearly on the loss distance αz (in contrast to the physical loss factor, which depends exponentially on distance).

3.3. Two-mode phase-sensitive links

Now consider the first stage of a two-mode PS link. In the first part of the stage the signal and idler modes both are attenuated. In the second part, the signal and idler interact (with a strong pump) and are amplified. By combining Eqs. (23), (25) and (26), one obtains for the p-quadratures the composite IO relations

p^1=(μτ)p^1+(ντ)p^2+(μρ)w^1+(νρ)w^2,
p^2=(μτ)p^2+(ντ)p^1+(μρ)w^2+(νρ)w^1,
where ŵ1 and ŵ2 are loss-mode quadratures. Similar relations apply to the q-quadratures (ν → −ν). Each output mean 〈p̂′1〉 = (µτ) 〈1〉 + (ντ) 〈2〉 and 〈p̂′2〉 = (µτ) 〈2〉 + (ντ) 〈1〉 depends on both input means. As stated earlier, the real quadratures interact constructively if they have the same sign (whereas the imaginary quadratures interact constructively if they have opposite signs).

Define the variances vj=p^j2p^j2 and correlation c12 = 〈12〉 − 〈1〉 〈2〉. Then, by combining Eqs. (43) and (44), and using the fact that the loss-mode quadratures are uncorrelated, one obtains the variance and correlation (moment) equations

v1=μ2τ2v1+2μντ2c12+ν2τ2v2+(μ2+ν2)ρ2σ,
c12=μντ2v1+(μ2+ν2)τ2c12+μντ2v2+2μνρ2σ,
v2=ν2τ2v1+2μντ2c12+μ2τ2v2+(μ2+ν2)ρ2σ,
where σ is the (common) variance of the loss modes.

If the input signal and idler means are equal, a balanced stage requires that (μ + ν)τ = 1. If the input fluctuations have the same variance (vj = v0), but are uncorrelated (c12 = 0), then the (common) output variance and correlation are

vj=v0(1+1/l2)/2+σ(l+1/l)(11/l)/2,
c12=v0(11/l2)/2+σ(l1/l)(11/l)/2,
respectively. If the stage loss (gain) is high (l ≫ 1), the output fluctuations are weaker after a PS stage than after the corresponding PI stage by a factor of 4, because the balanced-stage condition requires less pump power (so the gain process adds less excess noise). Furthermore, even though the input signal and idler fluctuations were uncorrelated, the output fluctuations are correlated strongly. One can derive the IO relations for a multiple-stage link by iterating Eqs. (45)(47). For a balanced link, the results are
vj=v0(1+1/l2s)/2+σ[s(l1)+(11/l2s)/(1+l)]/2,
c12=v0(11/l2s)/2+σ[s(l1)(11/l2s)/(1+l)]/2.
Once again, the output fluctuations are correlated strongly at a (common) level that is lower after a PS link than after the corresponding PI link by a factor of 4. Equations (50) and (51) reduce to Eqs. (48) and (49) in the limit s → 1.

For a short stage, τ2 ≈ 1 − αδz, ρ2αδz and (μ + ν)2 ≈ 1 + γδz. Hence, (μν)2 ≈ 1 − γδz and 2µνγδz. [These relations also imply that νγδz/2 and μ ≈ 1 + (γδz)2/8.] By making these substitutions in Eqs. (45)(47), one obtains the differential equations

dzv1=αv1+γc12+ασ,
dzc12=αc12+γ(v1+v2)/2,
dzv2=αv2+γc12+ασ.
Let d12 = (v1v2)/2 and s12 = (v1 + v2)/2. Then
dzs12=αs12+γc12+ασ,
dzc12=αc12+γs12,
dzd12=αd12.
Equation (57) implies that the variance difference decreases because of loss, independent of the variance sum and correlation. Equations (55) and (56) are linear, and can be solved for arbitrary gain and loss. The results are
s12(z)=[s12(0)cosh(γz)+c12(0)sinh(γz)]exp(αz)+σα{[αcosh(γz)+γsinh(γz)]exp(αz)α}/(γ2α2),
c12(z)=[s12(0)sinh(γz)+c12(0)cosh(γz)]exp(αz)+σα{[γcosh(γz)+αsinh(γz)]exp(αz)γ}/(γ2α2).
These solutions were checked numerically. For a balanced link with uncorrelated coherent-state-level input fluctuations (vj = σ), solutions (58) and (59) reduce to
s12(z)=α[2αz+3+exp(2αz)]/4,
c12(z)=σ[2αz+1exp(2αz)]/4,
respectively. Once again, in a link with distributed two-mode PS amplification, the variances and correlation increase only linearly with distance.

It is well known that a two-mode amplification process can be reformulated as two independent one-mode processes. Specifically, define the real sum- and difference-mode quadratures ± = (1 ± 2)/21/2, respectively, and the associated variances v±=p^±2p^±2. Then the superposition modes evolve independently, and their variances v± = s12 ± c12. By adding and subtracting Eqs. (58) and (59), one obtains Eqs. (32) and (33), respectively: The real sum-mode quadrature is stretched, whereas the real difference-mode quadrature is squeezed.

One can deduce results for the q-quadratures from the preceding results by making the substitution ν → −ν, which is equivalent to changing the sign of c12 in Eqs. (45)(47) and their solutions. In particular, the imaginary sum-mode quadrature is squeezed, whereas the imaginary difference-mode quadrature is stretched.

3.4. Noise figures

The standard metric for the performance of a link is its noise figure. For homodyne detection, the SNR of a signal is the square of its mean quadrature divided by its quadrature variance, and the noise figure of the link is the input SNR divided by the output SNR. Higher noise figures correspond to lower output SNRs, so the noise figure is a figure of demerit.

The SNR of a coherent state (CS) does not depend on its phase (the relative magnitude of p and q). It is well known that a CS subject to attenuation remains a CS: Its mean quadratures decrease, whereas its quadrature variances remain constant (because τ2 + ρ2 = 1). Hence, the noise figure of a CS subject to attenuation only (direct transmission) is the loss factor L = exp(αz), where α is the power-loss rate and z is the distance (link length).

By combining the preceding definition with the results of the preceding sections, one finds that the noise figures of one-mode PS links with in-phase CS inputs (p1 ≠ 0 and q1 = 0) are L and 1 + logL, for lumped and distributed amplification, respectively [Eqs. (29) and (34)]. For one-mode PI links with arbitrary CS inputs (p1 ≠ 0 and q1 ≠ 0), the noise figures are 1 + 2(L − 1) and 1 + 2logL, for lumped and distributed amplification, respectively [Eqs. (38) and (42)].

The noise figures of the aforementioned links are plotted as functions of loss in Fig. 2. (None of them depend on the power of the input signal.) Figure 2(a) shows that the noise figure of direct transmission increases linearly with the loss factor L (exponentially with the loss distance αz). Figure 2(b) shows that the noise figures of links with lumped amplification also increase linearly with loss, because attenuation preceeds amplification. The noise figure of the PS link equals the noise figure of direct transmission, because the PS amplifier adds no excess noise, whereas the noise figure of the PI link exceeds that of direct transmission by 3 dB, because the PI amplifier adds as much noise as the attenuator. Figure 2(b) also illustrates the benefits of distributed amplification: The noise figure of the distributed PS link is 3-dB lower than that of the distributed PI link, for the reason explained above, and the noise figures of both distributed links are lower than those of the lumped links, because the former noise figures increase only logarithmically with loss (linearly with distance). For high-loss links, this performance improvement is significant.

 figure: Fig. 2

Fig. 2 Noise figure plotted as a function of loss for links (a) without gain and (b) with gain. The dotted blue (green) curve represents a link with a single phase-sensitive (insensitive) amplifier, whereas the solid blue (green) curve represents a link with distributed phase-sensitive (phase-insensitive) amplification. Links with distributed amplification perform significantly better than links with lumped amplification.

Download Full Size | PDF

The noise figures of two-mode PS links are 6-dB lower than those of two-mode PI links. However, if one were to redefine the noise figures based on the total input power (signal and idler), they would be only 3-dB lower [23,24]. This result is consistent with the fact that a two-mode amplification process involving physical modes can be reformulated as two independent one-mode processes involving superposition modes (Sec. 3.3). Thus, the noise figures of two-mode PS links do not need to be illustrated separately.

4. One-mode fidelities and purities

In this section fidelity formulas will be derived for a single mode subject to attenuation and (PI and PS) amplification. However, to put these results in perspective, the fidelity formula for a single mode subject to attenuation only (direct transmission) is derived first. Consider an input coherent state (CS). It is well known that attenuation diminishes the mean quadratures, but maintains the quadrature variances at their vacuum level of 1/2 [Eqs. (32) and (33) with γ = 0.] Hence, in Eq. (17) the covariance matrices C1 = C2 = diag(1/2, 1/2) and C12 = diag(1, 1), so the determinant D12 = 1. The exponent is (p122+q122)/2, where p1 is an input quadrature mean, p2 = τp1 is an output quadrature mean and p12 = p1p2 is the quadrature difference. (Similar definitions apply to q1, q2 and q12.) Thus, for the attenuation of a coherent state, the fidelity

F12=exp[(p12+q12)(1τ)2/2].
Although fidelity (62) depends sensitively on the input photon number (p12+q12)/2=|α1|2, where α1 = (p1 + iq1)/21/2 is the input amplitude, it does not depend on the input phase arg(α1) = tan−1(q1/p1), because attenuation affects each quadrature in the same way. It is also well known that attenuation preserves the coherent nature of a state, so the fidelity should equal the overlap |〈α1|α2〉|2 = exp(−|α1α2|2), which it does.

Now consider an input single-photon state. In the beam-splitter model of attenuation, the input state is |1, 0〉 and the output state is τ|1, 0〉−ρ|0, 1〉, where the first and second elements of the ket-vectors represent the signal and loss mode, respectively, and the attenuation coefficients were assumed to be real, for simplicity. Hence, for the attenuation of a single-photon state, the fidelity

F12=τ2=1/L,
where L = exp(αz) is the loss factor. (In this example, one obtains the same result by using the state vectors of the signal and loss modes, or the reduced density matrices of the signal mode, which are |1〉 〈1| and τ2|1〉 〈1| + ρ2|0〉 〈0|.)

The fidelity of attenuation is plotted as a function of the loss factor in Fig. 3. The 1-photon CS curve decreases slowly towards the asymptote 0.37. This relatively-high fidelity reflects the fact that there is little difference between CS with the same quadrature fluctuations, and number means of 1 and 0. In contrast, the many-photon CS curves decrease rapidly to low asymptotic fidelitites and the 1-photon number-state curve also decreases rapidly towards the asymptote 0. These results show the limitations of direct transmission.

 figure: Fig. 3

Fig. 3 Fidelity of attenuation plotted as a function of the loss factor. The dotted, dashed, dot-dashed and solid orange curves represent input coherent states with photon-number means of 1, 3, 10 and 30, respectively, whereas the dotted red curve represents an input number state with 1 photon. For most input states, the fidelity decreases rapidly as the loss increases.

Download Full Size | PDF

The object of amplification is to compensate for attenuation, so the following discussions apply to balanced links, for which the output means of the quadratures that transport information (in principle) are equal the input means. First, consider a one-mode link with distributed PS amplification. For such a link the in-phase quadrature is preserved, so we choose p1 ≠ 0 and find that p2 = p1. However, the out-of-phase quadrature is squeezed, so we choose q1 = 0 and find that q2 = 0. In both cases (but for different reasons) the output mean equals the input mean, so the exponent in Eq. (17) is zero. For a CS both input variances are 1/2, so C1 = diag(1/2, 1/2). The output quadrature variances are specified by Eqs. (34) and (35), from which it follows that C2 = diag[(1 + logL)/2, (1 + 1/L2)/4]. By combining these results, one obtains the fidelity

F12={8/[(2+logL)(3+1/L2)]}1/2.
Notice that the fidelity decreases very slowly as the loss increases [F12 ∼ 1.6/(logL)1/2], because only the minimal amount of noise is added to the in-phase quadrature and the out-of-phase quadrature is squeezed slightly. One can derive a related formula for links with a finite number of discrete PS amplifiers by using Eqs. (30) instead of Eqs. (34) and (35).

Second, consider a one-mode link with distributed PI amplification. For such a link both quadratures are preserved, so we choose p1 ≠ 0 and q1 ≠ 0, and find that p2 = p1 and q2 = q1. In both cases the output mean equals the input mean, so the exponent in Eq. (17) is zero. For a CS both input variances are 1/2, so C1 = diag(1/2, 1/2). The output quadrature variances are specified by Eq. (42), from which it follows that C2 = diag(1/2 + logL, 1/2 + logL). By combining these results, one obtains the fidelity

F12=1/(1+logL).
For small losses, fidelity (65) decreases at the same rate as fidelity (63). Subsequently, fidelity (65) decreases slowly as the loss increases (F12 ∼ 1/logL), but it does not decrease as slowly as fidelity (64), because moderate amounts of excess noise are added to both quadratures. One can derive a related formula for links with a finite number of discrete PI amplifiers by using Eq. (39) instead of Eqs. (42). Notice that fidelities (64) and (65) are both independent of the input amplitude and phase.

The fidelity of transmission is plotted as a function of link loss in Fig. 4, for a variety of links with CS inputs. Also plotted are the attenuation fidelities of 1-photon-mean CS and 1-photon number-state inputs. Figure 4(a) shows that the transmission fidelity of a link with distributed PI amplification (which is the best PI link) is not significantly higher than the fidelity of a single photon subject to direct transmission. This result shows that fidelity is a more sensitive measure of link performance than noise figure (Sec. 3.4). In contrast, the transmission fidelity of a link with distributed PS amplification is higher than both of the direct-transmission fidelities for a wide range of loss factors. Consequently, we will not discuss PI links (with CS inputs) further, except to state that both types of link maintain the photon flux, which might be more important than the fidelity for some applications. Figure 4(b) shows the transmission fidelities of links with 1, 2 and 4 discrete PS amplifiers, which are placed at the ends of stages whose lengths equal the link length, and one-half and one-quarter of the link length, respectively. These results show that one can obtain good link performance with only a modest number of PS amplifiers

 figure: Fig. 4

Fig. 4 Fidelity of transmission plotted as a function of link loss. The solid blue and green curves represent links with distributed phase-sensitive and phase-insensitive amplification respectively, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. The inputs to these links are coherent states. The dotted orange and red curves are the attenuation fidelities of 1-photon-mean coherent-state and 1-photon number-state inputs, respectively. Phase-sensitive links with coherent-state inputs perform well.

Download Full Size | PDF

Since PS amplifiers stretch and squeeze the signal quadratures, it is natural to consider squeezed-state (SS) inputs. For such states C1 = diag(λ/2, 1/2λ), where λ is the (variance) squeezing factor. As explained before Eq. (64), for balanced links the fidelity does not depend on the mean (in-phase) quadrature. For a link with distributed PI amplification, C2 = diag(λ/2 + αz, 1/2λ + αz), from which it follows that

D12=1+(λ+1/λ)αz+(αz)2.
The right side of Eq. (66) depends symmetrically on λ and its inverse (because both quadratures evolve in the same way), so it is only necessary to consider λ ≥ 1. It is easy to show that D12 is an increasing function of λ, so there is no value of λ > 1 that allows a SS to have a higher transmission fidelity than a CS (λ = 1).

For a link with distributed PS amplification, C2 = diag[(λ + αz)/2, e−2αz/2λ + (1 − e−2αz)/4]. For short distances (αz ≪ 1), D12 ≈ 1 + (λ + 1/λ − 2)αz/2, which depends symmetrically on λ and its inverse. As αz increases, the second entry of C2 tends rapidly to 1/4, at which point

D121/2+λ/4+αz/4λ+αz/8.
If λ is much larger than 1 the coefficient of αz is 1/8, which is smaller than the CS coefficient 3/8 [Eq. (64)], whereas if λ is much smaller than 1 it is 1/4λ, which is larger than 3/8. Thus, one can increase the transmission fidelity of a high-loss link by using an input whose in-phase quadrature is stretched. Such an input has strong in-phase quadrature fluctuations, which make it less sensitive to the vacuum-level fluctuations contributed by the loss modes. For a fixed distance, the optimal value of λ is (αz)1/2 and the minimal value of D12 is [1 + (αz)1/2/2]2/2.

The fidelity of transmission is plotted as a function of loss in Fig. 5, for a link with distributed phase-sensitive amplification and a variety of SS inputs. The transmission fidelity of the same link with a CS input is also plotted for comparison. The results show that squeezing the in-phase quadrature always degrades the transmission fidelity. Stretching the in-phase quadrature decreases the fidelity for low losses, but increases the fidelity slightly for high losses. However, the loss value at which stretching improves the performance of the link is an increasing function of the stretching factor and high-loss links have low fidelities that decrease their usefulness.

 figure: Fig. 5

Fig. 5 Fidelity of transmission plotted as a function of loss for a link with distributed phase-sensitive amplification. The dotted, dashed, dot-dashed and solid purple curves represent inputs with stretching parameters of 0.1, 0.3, 3, and 10, respectively, whereas the solid blue curve represents an unstretched (coherent-state) input. Stretched inputs perform slightly better than the unstretched input for high losses The solid black curve represents the maximal fidelity (displaced upward by 0.01 for clarity).

Download Full Size | PDF

It is also instructive to consider the purities of the output states. As mentioned in Sec. 2, the purity is tr(ρ̂2). For a single-photon state subject to attenuation, the output density matrix is τ2|1〉 〈1| + ρ2|0〉 〈0|, so the output purity is τ4 + ρ4 = 1 − 2τ2ρ2. For a Gaussian state, the output purity is 1/2D21/2, where D2 is the determinant of the output covariance matrix. For all the examples considered in this section, the output covariance matrices were stated previously. It is easy to calculate their determinants (because they are diagonal), so there is no need to state the purity formulas explicitly.

The output purity is plotted as a function of loss in Fig. 6, for a variety of links without and with gain. For a single-photon state subject to attenuation only, the purity decreases from 1 to 1/2, then increases to 1: The state evolves from a pure state with 1 photon to a pure state with 0 photons. For a 1-photon-mean CS, the purity is constant. This result (which is independent of the number mean) reflects the fact that an attenuated CS is still coherent and pure. For a link with distributed PI amplification, the purity decreases rapidly: Not only is the output state mixed, rather than pure, it also bears little resemblance to the input state [Fig. 4(a)]. Thus, PI links do not transmit quantum signals faithfully. For links with PS amplification, the purity decreases slowly as the loss increases. For a 1-stage link, the purity is constant. This result (which is also independent of the number mean) reflects the fact that a stretched coherent state is pure. However, the output state bears little resemblance to the input state [Fig. 4(b)]. As the number of stages increases, the purity decreases. However, this decrease in purity corresponds to an increase in fidelity [Fig. 4(b)]. Thus, one can transmit a pure state that is nothing like the input state, or a mixed state that is a reasonable facsimile of the input state.

 figure: Fig. 6

Fig. 6 Purity of the output state plotted as a function of loss for links (a) without gain and (b) with gain. The dotted orange and red curves are the output purities of 1-photon-mean coherent-state and 1-photon number-state inputs, respectively. An attenuated coherent state is pure. The solid blue and green curves represent links with distributed phase-sensitive and phase-insensitive amplification respectively, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. The inputs to these links are coherent states. Phase-sensitive links degrade the purity less than phase-insensitive links, and a stretched coherent state is pure.

Download Full Size | PDF

The output purity is plotted as a function of loss in Fig. 7, for a link with distributed PS amplification and a variety of SS inputs. The output purity of a CS input is also plotted for comparison. For this example also, the output covariance matrix was stated previously. The associated determinant is the product of two factors, the variances of the in-phase and out-of-phase quadratures. The former variance increases monotonically as the loss increases, whereas the latter variance increases monotonically for λ > 2 and decreases monotonically for λ < 2. In both cases, the latter variance tends rapidly to its asymptotic value of 1/4. Consequently, if the in-phase quadrature is stretched, the purity tends to its asymptotic values [2/(λ + αz)]1/2 monotonically, whereas if the in-phase quadrature is squeezed, it tends to its asymptotic values nonmonotonically. For high losses, the purity depends only weakly on the stretching parameter.

 figure: Fig. 7

Fig. 7 Purity of the output state plotted as a function of loss for a link with distributed phase-sensitive amplification. The dotted, dashed, dot-dashed and solid purple curves represent inputs with stretching parameters of 0.1, 0.3, 3, and 10, respectively, whereas the solid blue curve represents an unstretched (coherent-state) input. For high losses, the output purity depends only weakly on the stretching parameter.

Download Full Size | PDF

5. Two-mode fidelities and purities

First, consider the fidelity of two-mode attenuation. If the input state is the product state |ψa〉 ⊗ |ψb〉, then the input density matrix ρ̂ = ρ̂aρ̂b, where ⊗ denotes a direct product. Such a density matrix obeys the trace rule tr(ρ̂) = tr(ρ̂a)tr(ρ̂b) and two such density matrices obey the product rule ρ̂1ρ̂2 = ρ̂1aρ̂2aρ̂1bρ̂2b. Each attenuation process is independent of the other. By combining these facts with Eq. (12), it is easy to show that the two-mode fidelity of attenuation is just the product of the one-mode fidelities. [The one-mode fidelities of attenuation for input coherent and single-photon states were specified by Eqs. (62) and (63), respectively.] Likewise, the two-mode output purity is the product of the one-mode output purities.

The fidelity of two-mode attenuation is plotted as a function of loss in Fig. 8(a), for a variety of input states. Although the two-mode fidelities are smaller than their one-mode counterparts (Fig. 3), their relative size and dependence on loss are similar.

 figure: Fig. 8

Fig. 8 (a) Fidelity of two-mode attenuation plotted as a function of loss. The dotted, dashed, dot-dashed and solid orange curves represent input coherent states with photon-number means of 1, 3, 10 and 30, respectively, whereas the dotted red curve represents an input number state with 1 photon. (b) Fidelity of transmission plotted as a function of loss. The solid blue curve represents a link with distributed phase-sensitive amplification, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. Two-mode phase-sensitive links with coherent-state inputs perform well.

Download Full Size | PDF

For two-mode PS links with CS inputs, the input and output covariance matrices

C1=σ[1000010000100001],C2=σ[s0c00s0cc0s00c0s],
where s and c represent variances and correlations, respectively, normalized to the vacuum variance σ = 1/2. For a link with distributed PS amplification, s and c are specified by Eqs. (60) and (61), respectively, whereas for a link with lumped PS amplification, they are specified by Eqs. (50) and (51). It follows from Eq. (17) and the preceding results that the two-mode fidelity
F12=4/[(1+s)2c2].

The fidelity of two-mode transmission is plotted as a function of loss in Fig. 8(b), for a variety of links with CS inputs. Although the two-mode fidelities are smaller than their one-mode counterparts [Fig. 4(b)], their relative size, and their dependence on loss and the number of amplifiers are similar. In every case considered, the two-mode fidelity (69) equals the square of the one-mode fidelity (64). This result reflects the fact that in the superposition-mode basis, two-mode transmission also is a separable process (Sec. 3.3). For this reason, there is no need to repeat the study of the fidelities associated with SS inputs, or the purities associated with arbitrary inputs.

Throughout this paper, the fidelity [Eqs. (12) and (17)] was used to quantify the similarities of the input and output states (of the same signal). In this context, high fidelities correspond to similar states. However, one could also use the fidelity to quantify the differences between two states (of different signals). In this context, low fidelities correspond to distinguishable states. For example, at the input, a vacuum state and a few-photon CS are easily distinguishable. During transmission through a balanced link, their mean quadratures stay the same, but their fluctuations increase (perhaps to the point of overlap). Hence, the output states are less distinguishable than the input ones. To quantify distinguishability, one must use the full fidelity formula of [31], because both output states are mixed.

6. Output-state reconstruction

In Secs. 3–5, two simple metrics (noise figure and fidelity) were used to characterize the performances of optical transmission systems. The first metric quantifies how much noise (in the form of quadrature fluctuations) a system adds to the signal, whereas the second quantifies the overlap between the input and output signal states. Although these metrics provide some information about the output signal (as does the purity), they do not describe its state completely. Remarkably, a complete description is available. All of the (linear) systems considered in Sec. 3 preserve the Gaussian nature of signal states. The Wigner functions of such states are specified completely by their covariance matrices, which were calculated in Sec. 3 and can be measured [44]. One can use these Wigner functions to calculate all of the quadrature (number) moments, which collectively specify the signal state, or reconstruct the associated density matrices [44]. In the number-state representation (20), the matrix elements ρmn are specified by integrals (21). In this section we take the latter approach, which we validated by comparing the integral results to the well-known formulas for coherent and squeezed states.

Array plots of the density matrices associated with different states are shown in Fig. 9. The input state [Fig. 9(a)] is coherent (pure), with a p-quadrature mean of 2.5 and a number mean of 3.0. Consequently, its density matrix has many large non-diagonal elements, which correspond to numbers (m or n) near the mean value. The output state produced by an attenuator with a loss factor of 3 dB [Fig. 9(b)] is also coherent. Its density matrix still has many large non-diagonal elements, which correspond to lower numbers. The output state produced by a balanced PI link with 3 dB of loss and lumped gain [Fig. 9(c)] has a density matrix with fewer large non-diagonal elements, which shows that the output state is less coherent (more mixed) than the input state. Although the output state produced by a balanced PS link [Fig. 9(d)] differs from the input state, its density matrix has many more large non-diagonal elements than the previous matrix, which shows that a PS link degrades the coherence of an input state less than the corresponding PI link, because it adds less noise to the in-phase quadrature and subtracts noise from the other (out-of-phase) quadrature.

 figure: Fig. 9

Fig. 9 Density matrices in the number-state representation. Darker squares denote higher values of |ρmn|, whereas lighter squares denote lower values. (a) input coherent state with a number mean of 3.0, (b) output coherent state produced by a 3-dB attenuator, (c) output state produced by a balanced phase-insensitive link with 3 dB of loss and lumped gain, and (d) output state produced by a balanced phase-sensitive link. The former link degrades the coherence of the state more than the latter.

Download Full Size | PDF

For the preceding cases, the theoretical purities (Sec. 4) are 1, 1, 1/3 and 1, respectively, and the numerical purities were 1.00, 1.00, 0.33 and 0.99. (An evaluation of the last purity based on a larger basis resulted in the value 1.00.) The first two values confirm that coherent and attenuated coherent states are pure. The third value shows that a PI link with only 3 dB of loss degrades the purity significantly. The fourth value shows that the state produced by a (one-stage) PS link is pure. This result reflects the fact that the input to the (first) PS amplifier is a coherent state, so the output from the amplifier is a squeezed coherent state [18], which is also pure.

Number distributions associated with the same states are shown in Fig. 10. The distribution associated with the input coherent state [Fig. 10(a)] is localized near the mean number 3.0, whereas the distribution associated with the attenuated coherent state [Fig. 10(b)] is localized near the mean number 1.5. In contrast, the distributions associated with the other output states are spread over broader ranges. This spreading is more significant for a PI link than a PS link, because the former link adds more noise to each quadrature.

 figure: Fig. 10

Fig. 10 Number distribution of (a) an input coherent state with a number mean of 3.0, (b) the output coherent state produced by a 3-dB attenuator, (c) the output state produced by a balanced phase-insensitive link with 3 dB of loss and lumped gain, and (d) the output state produced by a balanced phase-sensitive link. The former link distorts the number distribution more than the latter.

Download Full Size | PDF

A link with distributed PS gain (not shown) adds even less noise to the signal than the link with lumped PS gain, but for moderate loss factors, the performance improvement is also moderate. It is worth noting that the numerical purity associated with the distributed PS link is 0.97, which is slightly lower than the purity associated with the lumped PS link: Although less noise is added, the output state is not quite pure.

Loss and gain factors of 3 dB are enough to produce moderate changes in the signal state. However, typical factors are larger and can produce significant changes. To illustrate these changes, array plots of the density matrices produced by balanced links are shown in Fig. 11. For a link with 10 dB of loss and lumped PI gain [Fig. 11(a)], the matrix is almost completely diagonal, which means that the output state is almost completely incoherent. (The theoretical and numerical purities are 0.05.) In contrast, for a link with 10 dB of distributed loss and PS gain [Fig. 11(b)], the matrix still has many large non-diagonal elements, which means that the output state retains a high degree of coherence. (The theoretical and numerical purities are 0.77.)

 figure: Fig. 11

Fig. 11 Density matrices in the number-state representation. Darker squares denote higher values of |ρmn|, whereas lighter squares denote lower values. The output states are produced by balanced links with (a) 10 dB of loss and lumped gain, and (b) 10 dB of distributed loss and phase-sensitive gain. The state produced by the former link is nearly incoherent, whereas the state produced by the latter is nearly coherent.

Download Full Size | PDF

The number distributions produced by the same links are shown in Fig. 12. Both distributions differ noticeably from the input distribution [Fig. 10(a)]. For the lumped PI link [Fig. 12(a)], the probability of a substate with 20 photons is appreciable, despite the fact that the input state had only 3 photons! In contrast, for the distributed PS link [Fig. 12(b)], the probabilities of substates with 1–5 photons remain high, and the probabilities of substates with more than 12 photons are low, so the overlap of the input and output states is large (Fig. 4). For the stated loss factor, the distributed PS link performs significantly better than the corresponding lumped PS link (not shown).

 figure: Fig. 12

Fig. 12 Number distributions of the output states produced by balanced links with (a) 10 dB of loss and lumped phase-insensitive gain, and (b) 10 dB of distributed loss and phase-sensitive gain. The former link distorts the number distribution much more than the latter.

Download Full Size | PDF

7. Summary

In this paper, the effects of transmission on Gaussian optical states were studied in detail. Such states include strong and weak coherent states, which are used in conventional communication systems and quantum information experiments, respectively, and the photon-pair-like states produced by parametric sources, which also are used in quantum information experiments. Gaussian states are specified completely by their quadrature means, and the variances and correlations of their quadrature fluctuations.

In Sec. 3, equations were derived that model the spatial evolution of these quantities in transmission links with loss and gain. One-signal-mode links with phase-sensitive (PS) amplification were considered in Sec. 3.1, one-signal-mode links with (two-mode) phase-insensitive (PI) amplification were considered in Sec. 3.2 and two-mode links with PS amplification were considered in Sec. 3.3. In each section, formulas for the quadrature variances and correlations (elements of the covariance matrices) were derived for links with arbitrary numbers of discrete amplifiers (lumped links) and links with distributed amplification (distributed links). These formulas allow one to determine the noise figures of the links and their fidelities of transmission. The noise figures were compared in Sec. 3.4. In general, links with many weak amplifiers have lower noise figures than links with a few strong amplifiers. The noise figures of links with lumped amplification increase exponentially with distance (loss), whereas the noise figures of links with distributed amplification increase only linearly with distance. PI amplifiers add excess noise, whereas PS amplifiers do not. These facts are well known.

For a Gaussian state, there is a simple relation between its input and output covariance matrices, and the transmission fidelity [Eq. (17)]. In Sec. 4, this relation was used to determine the fidelities of one-signal-mode states transmitted through PI and PS links. For direct transmission (attenuation without amplification), the fidelity decreases rapidly with distance, because the quadrature means decrease. The fidelity associated with a distributed PI link is higher than that associated with direct transmission, but the margin is small. Although the link maintains the quadrature means, PI amplification adds excess noise to both quadratures, and even modest amounts of noise lower the fidelity appreciably. (For this link, fidelity is a more-sensitive performance metric than noise figure.) The fidelities associated with PS links are significantly higher. The distributed PS link has the highest fidelity, but even links with moderate numbers of discrete PS amplifiers have reasonable fidelities. One-mode PS links perform well because the noise in the (unused) out-of-phase quadrature is squeezed (clamped at half the vacuum level) and excess noise is not added to the stretched quadrature by the amplifiers (only by the attenuators). The fidelities of two-mode states transmitted through PS links were studied in Sec. 5. Because a two-mode state can be represented in terms of superposition modes, which are transmited independently, the two-mode fidelities are just the squares of the one-mode fidelitlies studied in the previous section. Thus, two-mode PS links also perform well.

It is possible to reconstruct the density operator of a Gaussian signal, in the number-state representation, from its covariance matrix [Eqs. (9), (11), (21) and (22)]. This density matrix allows one to determine the coherence (purity) of the output state and the probability distribution of photons within it (Sec. 6). For a balanced link, as distance increases the coherence decreases and the distribution spreads over a broader range of photon numbers. These changes develop much more rapidly for links with lumped PI amplification than for links with lumped or distributed PS amplification. In summary, based on the aforementioned considerations, links with distributed PS amplification perform significantly better than other links.

Appendix A: Wigner function

In this appendix an alternative formula for the Wigner function is derived and some of its properties are discussed. The trace in Eq. (5) is basis independent. In the q-basis, the characteristic function

C(k,l)=q|ρ^exp[i(kp^+lq^)]|qdq.
By using the Baker–Campbell–Hausdorff formula [24]
exp(x^+y^)=exp(x^)exp(y^)exp([x^,y^]/2),
for operators and ŷ that commute with their commutator, one finds that
exp[i(kp^+lq^)]=exp(ikp^)exp(ilq^)exp(ikl/2)=exp(ilq^)exp(ikp^)exp(ikl/2).
Further progress requires the identities
exp(ikp^)|p=exp(ikp)|p,exp(ikp^)|q=|q+k,
exp(ilq^)|p=|pl,exp(ilq^)|q=exp(ilq)|q.
The first and fourth of these identities follow from the definitions of quadrature ket-vectors, whereas the second and third follow from the generalized commutation relations
[q^,p^m]=imp^m1,[p^,q^n]=inq^n1,
and are consistent with the replacements = −iq and = ip, respectively. By combining definition (70) with the first of Eqs. (72) and second of Eqs. (73), one finds that
C(k,l)=qk/2|ρ^|q+k/2exp(ilq)dq.
By combining Eqs. (4) and (76), one obtains the Wigner formula [1]
W(p,q)=qk/2|ρ^|q+k/2exp(ikp)dk/2π,
which implies that the Wigner function is real. One also can derive a formula for the Wigner function by using the p-basis. The result is
W(p,q)=p+l/2|ρ^|pl/2exp(ilq)dl/2π.
The Wigner function is not a true PDF because it can have negative values (otherwise the overlap integrals of orthogonal states could not be zero). However, the marginal distributions
P(p)=W(p,q)dq=p|ρ^|p,Q(q)=W(p,q)dp=q|ρ^|q
are both non-negative and can be regarded as true PDFs. Formula (77) resembles a local Fourier transform, which indicates that the Wigner function contains information about the local wavenumber (instantaneous frequency) p at the position (time) q. Similar phase-space distribution functions have many applications in classical optics [38, 39].

Of particular importance is the Wigner function associated with the operator |n〉 〈m|. It follows from Eq. (77) that

Wnm(p,q)=ψn(qk/2)ψm*(q+k/2)exp(ikp)dk/2π,
where the number-state wavefunction [24]
ψn(q)=q|n=Hn(q)exp(q2/2)/(2nn!)1/2π1/4.
By combining Eqs. (80) and (81), completing the square and defining the variable l = k/2 − ip [3], one finds that
Wnm(p,q)=[exp(p2q2)/(2m+nm!n!)1/2π3/2]×Hm(q+ip+l)Hn(qipl)exp(l2)dl.
Further progress requires the identity Hn(−x) = (−1)nHn(x) [33] and formula 7.377 of [40], which states that
Hm(x+y)Hn(x+z)exp(x2)dx=2mπ1/2n!ymnLnmn(2yz)
for the case in which mn. By using these results to evaluate the integral in Eq. (82), one obtains the explicit formula
Wnm(p,q)=(1)n(n!/m!)1/2[21/2(q+ip)]mnexp(p2q2)Lnmn(2p2+2q2)/π.
Equation (84) is the first of Eqs. (22). For the complementary case in which m < n, one obtains the related formula
Wnm(p,q)=(1)m(m!/n!)1/2[21/2(qip)]nmexp(p2q2)Lmnm(2p2+2q2)/π,
which is the second of Eqs. (22).

The Wigner functions of the first four number states |n〉 (density operators |n〉 〈n|) are illustrated in Fig. 13. In contrast to the Wigner functions of Gaussian states, which are centered at arbitrary points in the pq-plane and decrease monotonically (Fig. 1), these Wigner functions are centered at the origin and all but one decrease in an oscillatory manner.

 figure: Fig. 13

Fig. 13 Wigner distribution functions of number states with 0, 1, 2 and 3 photons. The vacuum distribution (a) is Gaussian and positive, whereas the other distributions are non-Gaussian, and have both positive and negative values. As the photon number increases, so also does the number of oscillations.

Download Full Size | PDF

The right side of Eq. (13) involves the double integral of two Wigner functions, each of which is defined by Eq. (77). Observe that

qk1/2|ρ^1|q+k1/2qk2/2|ρ^2|q+k2/2exp[i(k1+k2)p]dk1dk2dpdq/2π=qk1/2|ρ^1|q+k1/2qk2/2|ρ^2|q+k2/2δ(k1+k2)dk1dk2dq=qk/2|ρ^1|q+k/2q+k/2|ρ^2|qk/2dkdq=q|ρ^1ρ^2|qdq,
from which follows the trace formula (13). This formula applies to hermitian and non-hermitian operators alike.

Appendix B: Mathematical identity

It is required to prove that

(A+B)(AB)S1(AB)=4(A1+B1)1,
where A and B are arbitrary square matrices and S = A + B is their sum. Suppose that the required matrix inverses exist. Then the relevant matrix product is
(A1+B1)[A+B(AB)S1(AB)]=(A1+B1)[(AAS1A)+(BBS1B)+AS1B+BS1A].
By expanding the first term in Eq. (88), one finds that
(A1+B1)(AAS1A)=IS1A+B1AB1(SB)S1A=I.
The expansion of the second term is similar. By expanding the third term, one finds that
(A1+B1)AS1B=S1B+B1(SB)S1B=I.
Alternatively, one can deduce Eq. (90) from Eq. (89) by using the fact that AS−1B = (SB)S−1B = BBS−1B. The expansion of the fourth term is similar. By combining Eqs. (89) and (90), one obtains the required result. Notice that this derivation does not depend on the dimension of the matrices involved (which is twice the number of participating modes).

Appendix C: Matrix elements

In this appendix, formulas are derived for the characteristic function Cnm(ξ) and Wigner function Wnm(α) associated with the density operator |n〉 〈m|. The former quantity is the matrix element 〈m|(ξ)|n〉, where (ξ) = exp(ξâξ*â) is the displacement operator, and the latter is the Fourier transform of the former. (Here is a displacement operator, not a determinant.) These formulas were derived by Cahill [41, 42], whose notations we follow loosely.

By using the Baker–Campbell–Hausdorff formula (71), one can rewrite the displacement operator in the alternative forms

D^(ξ)=exp(ξa^)exp(ξ*a^)exp(|ξ|2/2)=exp(ξ*a^)exp(ξa^)exp(|ξ|2/2),
which are described as normally and antinormally ordered, respectively. It is easy to verify that
m|exp(ξa^)=k=0m(mk)1/2ξkmk|(k!)1/2,
exp(ξ*a^)|n=l=0n(nl)1/2(ξ*)l|nl(l!)1/2.

Suppose that m = n + d, where d ≥ 0. Then, by combining the first of Eqs. (91) with Eqs. (92) and (93), and using the fact that 〈mk|nl〉 is only nonzero if k = l + d, one finds that

Cnm(ξ)=ξdexp(|ξ|2/2)l=0n[n!(n+d)!]1/2(|ξ|2)ll!(l+d)!(nl)!.
The associated Laguerre polynomials [33] are defined by the power series
Lnm(x)=l=0n(n+ml+m)(x)ll!.
By using this definition, one can rewrite the characteristic function (94) in the compact form
Cnm(ξ)=(n!/m!)1/2ξmnexp(|ξ|2/2)Lnmn(|ξ|2).

Conversely, suppose that n = m + d, where d > 0. Then, by repeating the preceding derivation, one obtains the related formula

Cnm(ξ)=(m!/n!)1/2(ξ*)nmexp(|ξ|2/2)Lmnm(|ξ|2).
If 〈m|(ξ)|n〉 is known for m > n, then its conjugate 〈n|(ξ)|m〉 = 〈n| (−ξ)|m〉, where n < m. Hence, one can deduce Eq. (97) from Eq. (96) by conjugating the characteristic function, replacing ξ by −ξ and interchanging the symbols m and n. For the characteristic function (96), conjugation and replacement are equivalent to the substitution ξ → −ξ*.

The characteristic and Wigner functions are related by the Fourier transforms [41, 42]

Wnm(α)=Cnm(ξ)exp(αξ*α*ξ)d2ξ/π,
Cnm(α)=Wnm(ξ)exp(ξα*ξ*α)d2α/π,
where d2ξ is an abbreviation for ri, and the subscripts r and i denote real and imaginary parts, respectively. Notice that in these definitions, the real and imaginary parts of the transform variables appear in the exponents with opposite signs, and the (common) denominator is π. In polar coordinates, α = re, ξ = ρe and d2ξ = ρdρdϕ. (Here ρ is a radius, not a density operator.) By combining Eq. (96), which is valid for mn, with Eq. (98), one finds that
Wnm(α)=(n!/m!)1/2ρd+1exp(ρ2/2)Lnd(ρ2)×exp[idϕ+i2rρsin(θϕ)]dρdϕ/π,
where d = mn ≥ 0. The angular part of this expression is
exp(idθ)exp[id(ϕθ)i2rρsin(ϕθ)]d(ϕθ).
By using the integral representation [33]
Jp(x)=ππexp[i(ptxsint)]dt/2π,
where p is an integer, one finds that
W(r,θ)=2(n!/m!)1/2exp(idθ)ρd+1exp(ρ2/2)Lnd(ρ2)Jd(ρy)dρ,
where y = 2r. Formula 7.421.4 of [40] implies that
xd+1exp(x2/2)Lnd(x2)Jd(xy)dx=(1)nydexp(y2/2)Lnd(y2),
from which it follows that
Wnm(r,θ)=2(1)n(n!/m!)1/2(2r)dexp(idθ2r2)Lnd(4r2).
This polar formula is equivalent to the cartesian formula [4143]
Wnm(α)=2(1)n(n!/m!)1/2(2α)mnexp(2|α|2)Lnmn(4|α|2).
By repeating the preceding derivation for m < n, one obtains the related formula
Wnm(α)=2(1)m(m!/n!)1/2(2α*)nmexp(2|α|2)Lmnm(4|α|2).
Notice that in contrast to Eq. (97), there is no factor of (−1)nm in Eq. (107). This result is a consequence of the fact that
Wnm(α)|m<n=Cmn(ξ*)|n>mexp(αξ*α*ξ)d2ξ/π=Cmn(ξ*)|n>mexp[α*(ξ*)*α(ξ*)]d2(ξ*)/π=Wmn(α*)|n>m.

By comparing the complex-argument transforms used in this appendix to the real-argument transforms used in the text, one finds that

Cr(k,l)=Cc(k/21/2,l/21/2),Wr(x,y)=Wc(y/21/2,x/21/2)/2π,
where the subscripts r and c denote real and complex, respectively, and the arguments of Cc and Wc are the real and imaginary parts of ξ and α, respectively. It follows from the real version of the trace formula [Eq. (13)] that
tr(ρ1ρ2)=W1(α)W2(α)d2α/π,
where the subscript c was omitted for simplicity.

Appendix D: Quadrature measurements

One can measure the photon number of a weak (quantum) signal by using a detector. One can measure the mode quadratures of the signal by combining it with a strong (classical) local oscillator at a beam splitter [24,23]. The difference between the output numbers is proportional to the input signal quadrature, whose phase is determined by the local-oscillator phase (which can be varied). Thus, homodyne detection allows one to measure the means and variances of the principal quadratures (in separate experiments). However, the quadrature correlation

p^q^=i(a^)2a^2+1/2,
which is not real. This result reflects the fact that and do not commute, so cannot be measured simultaneously [45].

If one were to pass the signal through a beam splitter prior to detection, the output mode-operators would be

a^1=τa^+ρw^,a^2=ρ*a^+τ*w^,
where ŵ is the input operator of the second beam-splitter mode (which is a vacuum state). For simplicity, we assume that τ and ρ are real, but our conclusions do not depend on this assumption. Equations (112) imply that the principal output quadratures are
p^1=[τ(a^+a^)+ρ(w^+w^)]/21/2,q^2=i[ρ(a^a^)+τ(w^+w^)]/21/2.
It is easy to verify that p̂′1 commutes with q̂′2, because the non-commuting contributions from â and ŵ cancel. Hence, p̂′1 and q̂′2 can be measured simultaneously, and their correlation
p^1q^2=iτρ(a^)2a^2/2=τρ(p^q^)s,
where the subscript s denotes symmetric ordering. Notice that the correlation measurement is not affected by the vacuum fluctuations of the second beam-splitter mode. However, the variances
(p^1)2=τ2p^2+ρ2/2,(q^2)2=ρ2q^2+τ2/2
do include contributions from the aforementioned fluctuations. Thus, by using a beam splitter prior to measurement, one can measure both signal quadratures, and their variances and correlation, simultaneously. The price one pays for this ability is the addition of noise to the variance measurements. Although this excess noise is comparable to the intrinsic noise of a coherent-state signal, it is much weaker than the excess noise added to that signal by a transmission link.

Acknowledgments

We acknowledge useful discussions with A. Agarwal, J. Dailey, D. James, N. Peters and P. Toliver. We also thank the reviewers for bringing to our attention [10] and [44]. This work was supported by the Defense Advanced Research Projects Agency under contract W31P4Q-13-C-0069. The views, opinions and findings contained in this article are those of the authors and should not be interpreted as representing the official views or policies of the Department of Defense or the U.S. Government.

References and links

1. E. Wigner, “On the quantum correction for thermodynamic equilibrium,” Phys. Rev. 40, 749–759 (1932). [CrossRef]  

2. U. Leonhardt, Measuring the Quantum State of Light (Cambridge University, 1997).

3. W. P. Schleich, Quantum Optics in Phase Space (Wiley, 2001), pp. 104–106.

4. C. C. Gerry and P. L. Knight, Introductory Quantum Optics (Cambridge University, 2005).

5. R. M. Gagliardi and S. Karp, Optical Communications, 2nd Ed. (Wiley, 1995).

6. G. P. Agrawal, Fiber-Optic Communication Systems, 4th Ed. (Wiley, 2010). [CrossRef]  

7. C. M. Caves and P. D. Drummond, “Quantum limits on bosonic communication rates,” Rev. Mod. Phys. 66, 481–537 (1994). [CrossRef]  

8. S. L. Braunstein and P. van Loock, “Quantum information with continuous variables,” Rev. Mod. Phys. 77, 513–577 (2005). [CrossRef]  

9. C. Weedbrook, S. Pirandola, R. Garcia-Patron, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaussian quantum information,” Rev. Mod. Phys. 84, 621–669 (2012). [CrossRef]  

10. F. Caruso, V. Giovannetti, C. Lupo, and S. Mancini, “Quantum channels and memory effects,” Rev. Mod. Phys. 86, 1203–1259 (2014). [CrossRef]  

11. D. C. Burnham and D. L. Weinberg, “Obervation of simultaneity in parametric production of optical photon pairs,” Phys. Rev. Lett. 25, 84–87 (1970). [CrossRef]  

12. M. Fiorentino, P. L. Voss, J. E. Sharping, and P. Kumar, “All-fiber photon-pair source for quantum communications,” IEEE Photon Technol. Lett. 14, 983–985 (2002). [CrossRef]  

13. H. A. Bachor and T. C. Ralph, A Guide to Experiments in Quantum Optics, 2nd Ed. (Wiley, 2004).

14. Y. Shih, An Introduction to Quantum Optics: Photon and Biphoton Physics (CRC Press, 2011).

15. C. E. Shannon, “A mathematical theory of communication,” Bell. Sys. Tech. J. 28, 379–423 and 623–656 (1948). [CrossRef]  

16. T. M. Cover and J. A. Thomas, Elements of Information Theory, 2nd Ed. (Wiley, 2006).

17. R. J. Glauber, “Coherent and incoherent states of the radiation field,” Phys. Rev. 131, 2766–2788 (1963). [CrossRef]  

18. H. P. Yuen, “Two-photon coherent states of the radiation field,” Phys. Rev. A 13, 2226–2243 (1976). [CrossRef]  

19. R. Loudon, “Theory of noise accumulation in linear optical-amplifier chains,” IEEE J. Quantum Electron. 21, 766–773 (1985). [CrossRef]  

20. R. E. Slusher and B. Yurke, “Squeezed light for coherent communications,” J. Lightwave Technol. 8, 466–477 (1990). [CrossRef]  

21. C. R. Giles and E. Desurvire, “Propagation of signal and noise in concatenated erbium-doped fiber optical amplifiers,” J. Lightwave Technol. 9, 147–154 (1991). [CrossRef]  

22. H. Yuen, “Reduction of quantum fluctuations and suppression of the Gordon–Haus effect with phase-sensitive linear amplifiers,” Opt. Lett. 12, 73–75 (1992). [CrossRef]  

23. C. J. McKinstrie, M. Yu, M. G. Raymer, and S. Radic, “Quantum noise properties of parametric processes,” Opt. Express 13, 4986–5012 (2005). [CrossRef]   [PubMed]  

24. M. Vasilyev, “Distributed phase-sensitive amplification,” Opt. Express 13, 7563–7571 (2005). [CrossRef]   [PubMed]  

25. C. J. McKinstrie, S. Radic, R. M. Jopson, and A. R. Chraplyvy, “Quantum noise limits on optical monitoring with parametric devices,” Opt. Commun. 259, 309–320 (2006). [CrossRef]  

26. M. Vasilyev and N. Stelmakh, “Squeezing and fiber-optic communications,” Int J. Mod. Phys. B 20, 1536–1542 (2006). [CrossRef]  

27. C. J. McKinstrie, M. Karlsson, and Z. Tong, “Field-quadrature and photon-number correlations produced by parametric processes,” Opt. Express 18, 19792–19823 (2010). [CrossRef]   [PubMed]  

28. C. J. McKinstrie, N. Alic, Z. Tong, and M. Karlsson, “Higher-capacity communication links based on two-mode phase-sensitive amplifiers,” Opt. Express 19, 11977–11991 (2011). [CrossRef]   [PubMed]  

29. A. Papoulis, Probability, Random Variables and Stochastic Processes, 3rd Ed. (McGraw-Hill, 1991).

30. R. Jozsa, “Fidelity for mixed quantum states,” J. Mod. Opt. 41, 2315–2323 (1994). [CrossRef]  

31. H. Scutaru, “Fidelity for displaced squeezed thermal states and the oscillator semigroup,” J. Phys. A: Math. Gen. 31, 3659–3663 (1998). [CrossRef]  

32. C. J. McKinstrie and N. Alic, “Information efficiencies of parametric devices,” IEEE J. Sel. Top. Quantum Electron. 18, 794–811 (2012). [CrossRef]  

33. N. N. Lebedev, Special Functions and their Applications (Dover, 1972).

34. M. E. Marhic, Fiber Optical Parametric Amplifiers, Oscillators and Related Devices (Cambridge University, 2007). [CrossRef]  

35. A. Yariv, “Signal-to-noise considerations in fiber links with periodic or distributed optical amplification,” Opt. Lett. 15, 1064–1066 (1990). [CrossRef]   [PubMed]  

36. J. P. Gordon and L. F. Mollenauer, “Effects of fiber nonlinearities and amplifier spacing on ultra-long distance transmission,” J. Lightwave Technol. 9, 170–173 (1991). [CrossRef]  

37. F. Caruso, V. Giovannetti, and A. S. Kholevo, “One-mode bosonic Gaussian channels: a full weak-degradability classification,” New J. Phys 8, 310 (2006). [CrossRef]  

38. M. J. Bastiaans, “Wigner distribution function and its application to first-order optics,” J. Opt. Soc. Am. 69, 1710–1716 (1979). [CrossRef]  

39. J. Paye, “The chroncyclic representation of ultrashort light pulses,” IEEE J. Quantum Electron. 28, 2262–2273 (1992). [CrossRef]  

40. I. S. Gradsteyn and I. M. Ryzhik, Tables of Integrals, Series and Products (Academic Press, 2007).

41. K. E. Cahill and R. J. Glauber, “Ordered expansions in boson amplitude operators,” Phys. Rev. 177, 1857–1881 (1969). [CrossRef]  

42. K. E. Cahill and R. J. Glauber, “Density operators and quasiprobability distributions,” Phys. Rev. 177, 1882–1902 (1969). [CrossRef]  

43. B. M. Garraway and P. L. Knight, “Quantum phase distributions and quasidistributions,” Phys. Rev. A 46, R5346–R5349 (1992). [CrossRef]   [PubMed]  

44. D. T. Smithey, M. Beck, M. G. Raymer, and A. Faridani, “Measurement of the Wigner distribution and the density matrix of a light mode using optical homodyne tomography: Application to squeezed states and the vacuum,” Phys. Rev. Lett. 70, 1244–1247 (1993). [CrossRef]   [PubMed]  

45. E. Arthurs and J. L. Kelley, “On the simultaneous measurement of a pair of conjugate variables,” Bell Sys. Tech. J. 44, 725–729 (1965). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (13)

Fig. 1
Fig. 1 Wigner distribution function of (a) a coherent state with a p-quadrature mean of 2.5, which corresponds to a number mean of 3.0, and (b) a state whose p- and q-quadratures have been stretched and squeezed, respectively. Both distributions are Gaussian and positive.
Fig. 2
Fig. 2 Noise figure plotted as a function of loss for links (a) without gain and (b) with gain. The dotted blue (green) curve represents a link with a single phase-sensitive (insensitive) amplifier, whereas the solid blue (green) curve represents a link with distributed phase-sensitive (phase-insensitive) amplification. Links with distributed amplification perform significantly better than links with lumped amplification.
Fig. 3
Fig. 3 Fidelity of attenuation plotted as a function of the loss factor. The dotted, dashed, dot-dashed and solid orange curves represent input coherent states with photon-number means of 1, 3, 10 and 30, respectively, whereas the dotted red curve represents an input number state with 1 photon. For most input states, the fidelity decreases rapidly as the loss increases.
Fig. 4
Fig. 4 Fidelity of transmission plotted as a function of link loss. The solid blue and green curves represent links with distributed phase-sensitive and phase-insensitive amplification respectively, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. The inputs to these links are coherent states. The dotted orange and red curves are the attenuation fidelities of 1-photon-mean coherent-state and 1-photon number-state inputs, respectively. Phase-sensitive links with coherent-state inputs perform well.
Fig. 5
Fig. 5 Fidelity of transmission plotted as a function of loss for a link with distributed phase-sensitive amplification. The dotted, dashed, dot-dashed and solid purple curves represent inputs with stretching parameters of 0.1, 0.3, 3, and 10, respectively, whereas the solid blue curve represents an unstretched (coherent-state) input. Stretched inputs perform slightly better than the unstretched input for high losses The solid black curve represents the maximal fidelity (displaced upward by 0.01 for clarity).
Fig. 6
Fig. 6 Purity of the output state plotted as a function of loss for links (a) without gain and (b) with gain. The dotted orange and red curves are the output purities of 1-photon-mean coherent-state and 1-photon number-state inputs, respectively. An attenuated coherent state is pure. The solid blue and green curves represent links with distributed phase-sensitive and phase-insensitive amplification respectively, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. The inputs to these links are coherent states. Phase-sensitive links degrade the purity less than phase-insensitive links, and a stretched coherent state is pure.
Fig. 7
Fig. 7 Purity of the output state plotted as a function of loss for a link with distributed phase-sensitive amplification. The dotted, dashed, dot-dashed and solid purple curves represent inputs with stretching parameters of 0.1, 0.3, 3, and 10, respectively, whereas the solid blue curve represents an unstretched (coherent-state) input. For high losses, the output purity depends only weakly on the stretching parameter.
Fig. 8
Fig. 8 (a) Fidelity of two-mode attenuation plotted as a function of loss. The dotted, dashed, dot-dashed and solid orange curves represent input coherent states with photon-number means of 1, 3, 10 and 30, respectively, whereas the dotted red curve represents an input number state with 1 photon. (b) Fidelity of transmission plotted as a function of loss. The solid blue curve represents a link with distributed phase-sensitive amplification, whereas the dotted, dashed and dot-dashed blue curves represent links with 1, 2 and 4 phase-sensitive amplifiers, respectively. Two-mode phase-sensitive links with coherent-state inputs perform well.
Fig. 9
Fig. 9 Density matrices in the number-state representation. Darker squares denote higher values of |ρmn|, whereas lighter squares denote lower values. (a) input coherent state with a number mean of 3.0, (b) output coherent state produced by a 3-dB attenuator, (c) output state produced by a balanced phase-insensitive link with 3 dB of loss and lumped gain, and (d) output state produced by a balanced phase-sensitive link. The former link degrades the coherence of the state more than the latter.
Fig. 10
Fig. 10 Number distribution of (a) an input coherent state with a number mean of 3.0, (b) the output coherent state produced by a 3-dB attenuator, (c) the output state produced by a balanced phase-insensitive link with 3 dB of loss and lumped gain, and (d) the output state produced by a balanced phase-sensitive link. The former link distorts the number distribution more than the latter.
Fig. 11
Fig. 11 Density matrices in the number-state representation. Darker squares denote higher values of |ρmn|, whereas lighter squares denote lower values. The output states are produced by balanced links with (a) 10 dB of loss and lumped gain, and (b) 10 dB of distributed loss and phase-sensitive gain. The state produced by the former link is nearly incoherent, whereas the state produced by the latter is nearly coherent.
Fig. 12
Fig. 12 Number distributions of the output states produced by balanced links with (a) 10 dB of loss and lumped phase-insensitive gain, and (b) 10 dB of distributed loss and phase-sensitive gain. The former link distorts the number distribution much more than the latter.
Fig. 13
Fig. 13 Wigner distribution functions of number states with 0, 1, 2 and 3 photons. The vacuum distribution (a) is Gaussian and positive, whereas the other distributions are non-Gaussian, and have both positive and negative values. As the photon number increases, so also does the number of oscillations.

Equations (115)

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

p m q n = W ( p , q ) p m q n d p d q .
C ( k , l ) = W ( p , q ) exp [ i ( k p + l q ) ] d p d q .
p m q n = lim k 0 lim l 0 ( i k ) m ( i l ) n C ( k , l ) .
W ( p , q ) = C ( k , l ) exp [ i ( k p + l q ) ] d k d l / ( 2 π ) 2 .
C ( k , l ) = exp [ i ( k p ^ + l q ^ ) ] = tr { ρ ^ exp [ i ( k p ^ + l q ^ ) ] } ,
( k p ^ + l q ^ ) n = lim θ 0 ( i θ ) n C ( θ k , θ l ) .
( k p ^ + l q ^ ) n = W ( p , q ) ( k p + l q ) n d p d q .
( p ^ m q ^ n ) s = W ( p , q ) p m q n d p d q ,
W ( X ) = ( Δ s 1 / 2 / 2 π ) exp [ ( X V ) t S ( X V ) / 2 ] ,
exp ( X t A X / 2 ) d 2 x = 2 π / Δ a 1 / 2 ,
X = V , ( X V ) ( X V ) t = S 1 = C .
F 12 = [ tr ( ρ ^ 1 1 / 2 ρ ^ 2 ρ ^ 1 1 / 2 ) 1 / 2 ] 2 = [ tr ( ρ ^ 2 1 / 2 ρ ^ 1 ρ ^ 2 1 / 2 ) 1 / 2 ] 2 .
tr ( ρ ^ 1 ρ ^ 2 ) = 2 π W 1 ( p , q ) W 2 ( p , q ) d p d q ,
F 12 = [ ( Δ 1 Δ 2 ) 1 / 2 / 2 π ] exp [ ( X V 1 ) t S 1 ( X V 1 ) / 2 ] × exp [ ( X V 2 ) t S 2 ( X V 2 ) / 2 ] d 2 x .
( X V ) t S 1 ( X V ) + ( X + V ) t S 2 ( X + V ) = X t ( S 1 + S 2 ) X X t ( S 1 S 2 ) V V t ( S 1 S 2 ) X + V t ( S 1 + S 2 ) V = X t S + X X t S V V t S X + V t S + V = ( X W ) t S + ( X W ) W t S + W + V t S + V = ( X W ) t S + ( X W ) + V t ( S + S S + 1 S ) V ,
F 12 = ( Δ 1 Δ 2 / Δ + ) 1 / 2 exp [ V t ( S + S S + 1 S ) V / 2 ] .
F 12 = D 12 1 / 2 exp [ ( V 1 V 2 ) C 12 1 ( V 1 V 2 ) / 2 ] ,
det ( S 1 + S 2 ) / det ( S 1 S 2 ) = det ( S 1 + S 2 ) det ( S 1 1 S 2 1 ) = det ( S 1 1 + S 2 1 ) ,
S + S S + 1 S = 4 ( S 1 1 + S 2 1 ) 1 ,
ρ ^ = m n | m ρ m n n | ,
ρ m n = 2 π W n m ( p , q ) W ( p , q ) d p d q ,
W n m ( p , q ) = { ( 1 ) n ( 2 m n n ! / m ! ) 1 / 2 ( q + i p ) m n exp ( p 2 q 2 ) L n m n ( 2 p 2 + 2 q 2 ) / π , ( 1 ) m ( 2 n m m ! / n ! ) 1 / 2 ( q i p ) n m exp ( p 2 q 2 ) L m n m ( 2 p 2 + 2 q 2 ) / π ,
p ^ = | τ | p ^ + | ρ | w ^ p , q ^ = | τ | q ^ + | ρ | w ^ q ,
p ^ = ( | μ | + | ν | ) p ^ , q ^ = ( | μ | | ν | ) q ^ ,
p ^ 1 = | μ | p ^ 1 + | ν | p ^ 2 , q ^ 1 = | μ | q ^ 1 | ν | q ^ 2 ,
p ^ 2 = | μ | p ^ 2 + | ν | p ^ 1 , q ^ 2 = | μ | q ^ 2 | ν | q ^ 1 .
p ^ = ( λ + τ ) p ^ + ( λ + ρ ) w ^ p , q ^ = ( λ τ ) q ^ + ( λ ρ ) w ^ q ,
v p = ( λ + τ ) 2 v p + ( λ + ρ ) 2 σ , v q = ( λ τ ) 2 v q + ( λ ρ ) 2 σ ,
v p = v p + σ ( l 1 ) , v q = v q / l 2 + σ ( l 1 ) / l 2 ,
v p = v p + σ s ( l 1 ) , v q = v q / l 2 s + σ ( 1 1 / l 2 s ) / ( l + 1 ) .
d z v p = ( γ α ) v p + α σ , d z v q = ( γ + α ) v q + α σ .
v p ( z ) = v p ( 0 ) exp ( λ p z ) + σ α [ exp ( λ p z ) 1 ] / λ p ,
v q ( z ) = v q ( 0 ) exp ( λ q z ) + σ α [ 1 exp ( λ q z ) ] / λ q ,
v p ( z ) = v p ( 0 ) + σ α z ,
v q ( z ) = v q ( 0 ) exp ( 2 α z ) + σ [ 1 exp ( 2 α z ) ] / 2 .
p ^ 1 = ( μ τ ) p ^ 1 + ( μ ρ ) w ^ p + ν p ^ 2 , q ^ 1 = ( μ τ ) q ^ 1 + ( μ ρ ) w ^ q ν q ^ 2 .
v p = ( μ τ ) 2 v p + ( μ 2 ρ 2 + ν 2 ) σ ,
v p = v p + 2 ( l 1 ) σ .
v p = v p + 2 s ( l 1 ) σ .
d z v p = ( γ α ) v p + ( γ + α ) σ .
v p ( z ) = v p ( 0 ) exp ( λ z ) + σ λ + [ exp ( λ z ) 1 ] / λ ,
v p ( z ) = v p ( 0 ) + 2 σ α z .
p ^ 1 = ( μ τ ) p ^ 1 + ( ν τ ) p ^ 2 + ( μ ρ ) w ^ 1 + ( ν ρ ) w ^ 2 ,
p ^ 2 = ( μ τ ) p ^ 2 + ( ν τ ) p ^ 1 + ( μ ρ ) w ^ 2 + ( ν ρ ) w ^ 1 ,
v 1 = μ 2 τ 2 v 1 + 2 μ ν τ 2 c 12 + ν 2 τ 2 v 2 + ( μ 2 + ν 2 ) ρ 2 σ ,
c 12 = μ ν τ 2 v 1 + ( μ 2 + ν 2 ) τ 2 c 12 + μ ν τ 2 v 2 + 2 μ ν ρ 2 σ ,
v 2 = ν 2 τ 2 v 1 + 2 μ ν τ 2 c 12 + μ 2 τ 2 v 2 + ( μ 2 + ν 2 ) ρ 2 σ ,
v j = v 0 ( 1 + 1 / l 2 ) / 2 + σ ( l + 1 / l ) ( 1 1 / l ) / 2 ,
c 12 = v 0 ( 1 1 / l 2 ) / 2 + σ ( l 1 / l ) ( 1 1 / l ) / 2 ,
v j = v 0 ( 1 + 1 / l 2 s ) / 2 + σ [ s ( l 1 ) + ( 1 1 / l 2 s ) / ( 1 + l ) ] / 2 ,
c 12 = v 0 ( 1 1 / l 2 s ) / 2 + σ [ s ( l 1 ) ( 1 1 / l 2 s ) / ( 1 + l ) ] / 2 .
d z v 1 = α v 1 + γ c 12 + α σ ,
d z c 12 = α c 12 + γ ( v 1 + v 2 ) / 2 ,
d z v 2 = α v 2 + γ c 12 + α σ .
d z s 12 = α s 12 + γ c 12 + α σ ,
d z c 12 = α c 12 + γ s 12 ,
d z d 12 = α d 12 .
s 12 ( z ) = [ s 12 ( 0 ) cosh ( γ z ) + c 12 ( 0 ) sinh ( γ z ) ] exp ( α z ) + σ α { [ α cosh ( γ z ) + γ sinh ( γ z ) ] exp ( α z ) α } / ( γ 2 α 2 ) ,
c 12 ( z ) = [ s 12 ( 0 ) sinh ( γ z ) + c 12 ( 0 ) cosh ( γ z ) ] exp ( α z ) + σ α { [ γ cosh ( γ z ) + α sinh ( γ z ) ] exp ( α z ) γ } / ( γ 2 α 2 ) .
s 12 ( z ) = α [ 2 α z + 3 + exp ( 2 α z ) ] / 4 ,
c 12 ( z ) = σ [ 2 α z + 1 exp ( 2 α z ) ] / 4 ,
F 12 = exp [ ( p 1 2 + q 1 2 ) ( 1 τ ) 2 / 2 ] .
F 12 = τ 2 = 1 / L ,
F 12 = { 8 / [ ( 2 + log L ) ( 3 + 1 / L 2 ) ] } 1 / 2 .
F 12 = 1 / ( 1 + log L ) .
D 12 = 1 + ( λ + 1 / λ ) α z + ( α z ) 2 .
D 12 1 / 2 + λ / 4 + α z / 4 λ + α z / 8 .
C 1 = σ [ 1 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 ] , C 2 = σ [ s 0 c 0 0 s 0 c c 0 s 0 0 c 0 s ] ,
F 12 = 4 / [ ( 1 + s ) 2 c 2 ] .
C ( k , l ) = q | ρ ^ exp [ i ( k p ^ + l q ^ ) ] | q d q .
exp ( x ^ + y ^ ) = exp ( x ^ ) exp ( y ^ ) exp ( [ x ^ , y ^ ] / 2 ) ,
exp [ i ( k p ^ + l q ^ ) ] = exp ( i k p ^ ) exp ( i l q ^ ) exp ( i k l / 2 ) = exp ( i l q ^ ) exp ( i k p ^ ) exp ( i k l / 2 ) .
exp ( i k p ^ ) | p = exp ( i k p ) | p , exp ( i k p ^ ) | q = | q + k ,
exp ( i l q ^ ) | p = | p l , exp ( i l q ^ ) | q = exp ( i l q ) | q .
[ q ^ , p ^ m ] = i m p ^ m 1 , [ p ^ , q ^ n ] = i n q ^ n 1 ,
C ( k , l ) = q k / 2 | ρ ^ | q + k / 2 exp ( i l q ) d q .
W ( p , q ) = q k / 2 | ρ ^ | q + k / 2 exp ( i k p ) d k / 2 π ,
W ( p , q ) = p + l / 2 | ρ ^ | p l / 2 exp ( i l q ) d l / 2 π .
P ( p ) = W ( p , q ) d q = p | ρ ^ | p , Q ( q ) = W ( p , q ) d p = q | ρ ^ | q
W n m ( p , q ) = ψ n ( q k / 2 ) ψ m * ( q + k / 2 ) exp ( i k p ) d k / 2 π ,
ψ n ( q ) = q | n = H n ( q ) exp ( q 2 / 2 ) / ( 2 n n ! ) 1 / 2 π 1 / 4 .
W n m ( p , q ) = [ exp ( p 2 q 2 ) / ( 2 m + n m ! n ! ) 1 / 2 π 3 / 2 ] × H m ( q + i p + l ) H n ( q i p l ) exp ( l 2 ) d l .
H m ( x + y ) H n ( x + z ) exp ( x 2 ) d x = 2 m π 1 / 2 n ! y m n L n m n ( 2 y z )
W n m ( p , q ) = ( 1 ) n ( n ! / m ! ) 1 / 2 [ 2 1 / 2 ( q + i p ) ] m n exp ( p 2 q 2 ) L n m n ( 2 p 2 + 2 q 2 ) / π .
W n m ( p , q ) = ( 1 ) m ( m ! / n ! ) 1 / 2 [ 2 1 / 2 ( q i p ) ] n m exp ( p 2 q 2 ) L m n m ( 2 p 2 + 2 q 2 ) / π ,
q k 1 / 2 | ρ ^ 1 | q + k 1 / 2 q k 2 / 2 | ρ ^ 2 | q + k 2 / 2 exp [ i ( k 1 + k 2 ) p ] d k 1 d k 2 d p d q / 2 π = q k 1 / 2 | ρ ^ 1 | q + k 1 / 2 q k 2 / 2 | ρ ^ 2 | q + k 2 / 2 δ ( k 1 + k 2 ) d k 1 d k 2 d q = q k / 2 | ρ ^ 1 | q + k / 2 q + k / 2 | ρ ^ 2 | q k / 2 d k d q = q | ρ ^ 1 ρ ^ 2 | q d q ,
( A + B ) ( A B ) S 1 ( A B ) = 4 ( A 1 + B 1 ) 1 ,
( A 1 + B 1 ) [ A + B ( A B ) S 1 ( A B ) ] = ( A 1 + B 1 ) [ ( A A S 1 A ) + ( B B S 1 B ) + A S 1 B + B S 1 A ] .
( A 1 + B 1 ) ( A A S 1 A ) = I S 1 A + B 1 A B 1 ( S B ) S 1 A = I .
( A 1 + B 1 ) A S 1 B = S 1 B + B 1 ( S B ) S 1 B = I .
D ^ ( ξ ) = exp ( ξ a ^ ) exp ( ξ * a ^ ) exp ( | ξ | 2 / 2 ) = exp ( ξ * a ^ ) exp ( ξ a ^ ) exp ( | ξ | 2 / 2 ) ,
m | exp ( ξ a ^ ) = k = 0 m ( m k ) 1 / 2 ξ k m k | ( k ! ) 1 / 2 ,
exp ( ξ * a ^ ) | n = l = 0 n ( n l ) 1 / 2 ( ξ * ) l | n l ( l ! ) 1 / 2 .
C n m ( ξ ) = ξ d exp ( | ξ | 2 / 2 ) l = 0 n [ n ! ( n + d ) ! ] 1 / 2 ( | ξ | 2 ) l l ! ( l + d ) ! ( n l ) ! .
L n m ( x ) = l = 0 n ( n + m l + m ) ( x ) l l ! .
C n m ( ξ ) = ( n ! / m ! ) 1 / 2 ξ m n exp ( | ξ | 2 / 2 ) L n m n ( | ξ | 2 ) .
C n m ( ξ ) = ( m ! / n ! ) 1 / 2 ( ξ * ) n m exp ( | ξ | 2 / 2 ) L m n m ( | ξ | 2 ) .
W n m ( α ) = C n m ( ξ ) exp ( α ξ * α * ξ ) d 2 ξ / π ,
C n m ( α ) = W n m ( ξ ) exp ( ξ α * ξ * α ) d 2 α / π ,
W n m ( α ) = ( n ! / m ! ) 1 / 2 ρ d + 1 exp ( ρ 2 / 2 ) L n d ( ρ 2 ) × exp [ i d ϕ + i 2 r ρ sin ( θ ϕ ) ] d ρ d ϕ / π ,
exp ( i d θ ) exp [ i d ( ϕ θ ) i 2 r ρ sin ( ϕ θ ) ] d ( ϕ θ ) .
J p ( x ) = π π exp [ i ( p t x sin t ) ] d t / 2 π ,
W ( r , θ ) = 2 ( n ! / m ! ) 1 / 2 exp ( i d θ ) ρ d + 1 exp ( ρ 2 / 2 ) L n d ( ρ 2 ) J d ( ρ y ) d ρ ,
x d + 1 exp ( x 2 / 2 ) L n d ( x 2 ) J d ( x y ) d x = ( 1 ) n y d exp ( y 2 / 2 ) L n d ( y 2 ) ,
W n m ( r , θ ) = 2 ( 1 ) n ( n ! / m ! ) 1 / 2 ( 2 r ) d exp ( i d θ 2 r 2 ) L n d ( 4 r 2 ) .
W n m ( α ) = 2 ( 1 ) n ( n ! / m ! ) 1 / 2 ( 2 α ) m n exp ( 2 | α | 2 ) L n m n ( 4 | α | 2 ) .
W n m ( α ) = 2 ( 1 ) m ( m ! / n ! ) 1 / 2 ( 2 α * ) n m exp ( 2 | α | 2 ) L m n m ( 4 | α | 2 ) .
W n m ( α ) | m < n = C m n ( ξ * ) | n > m exp ( α ξ * α * ξ ) d 2 ξ / π = C m n ( ξ * ) | n > m exp [ α * ( ξ * ) * α ( ξ * ) ] d 2 ( ξ * ) / π = W m n ( α * ) | n > m .
C r ( k , l ) = C c ( k / 2 1 / 2 , l / 2 1 / 2 ) , W r ( x , y ) = W c ( y / 2 1 / 2 , x / 2 1 / 2 ) / 2 π ,
tr ( ρ 1 ρ 2 ) = W 1 ( α ) W 2 ( α ) d 2 α / π ,
p ^ q ^ = i ( a ^ ) 2 a ^ 2 + 1 / 2 ,
a ^ 1 = τ a ^ + ρ w ^ , a ^ 2 = ρ * a ^ + τ * w ^ ,
p ^ 1 = [ τ ( a ^ + a ^ ) + ρ ( w ^ + w ^ ) ] / 2 1 / 2 , q ^ 2 = i [ ρ ( a ^ a ^ ) + τ ( w ^ + w ^ ) ] / 2 1 / 2 .
p ^ 1 q ^ 2 = i τ ρ ( a ^ ) 2 a ^ 2 / 2 = τ ρ ( p ^ q ^ ) s ,
( p ^ 1 ) 2 = τ 2 p ^ 2 + ρ 2 / 2 , ( q ^ 2 ) 2 = ρ 2 q ^ 2 + τ 2 / 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.