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

Enhanced secret-key generation from atmospheric optical channels with the use of random modulation

Open Access Open Access

Abstract

Secret-key extraction from atmospheric optical channels can generate common keys used by two communicating parties to encrypt their transmitted confidential information. The correlation time of turbulence-induced optical fluctuations imposes a restriction on the number of extracted uncorrelated key bits per second. To break this restriction, we propose a novel randomness sharing scheme between two communicating parties using an atmospheric optical channel equipped with random modulation and develop mathematical models for the common randomness source created by such a randomness-sharing scheme. Our randomness-sharing scheme provides the legitimate parties with the ability to decrease the temporal autocorrelation of the said common randomness source, which is called controllable common randomness source with memory (CCRSM), thereby enabling an enhanced secret-key extraction that can break the aforementioned restriction. Both the autocorrelation of the legitimate parties’ observations of the CCRSM and the cross-correlation of the two legitimate parties’ observations of the CCRSM are formulated and examined theoretically. It is found that the performing random modulation can decorrelate consecutive observations of the CCRSM obtained by the legitimate parties using a sampling interval smaller than the correlation time of turbulence-induced optical fluctuations. The cross-correlation coefficient of the eavesdropper’s and legitimate parties’ observations of the CCRSM is dealt with theoretically and the eavesdropping risk in the CCRSM-based key extraction is inspected for the fully-disclosed-single-modulation worst eavesdropping (FDSM-WE) scenario, partially-disclosed-single-modulation worst eavesdropping (PDSM-WE) scenario and double-modulation worst eavesdropping (DM-WE) scenario. It is shown that the FDSM-WE scenario has the highest degree of eavesdropping risk. Finally, the lowest limit of the secret-key capacity in consideration of using the CCRSM is theoretically formulated. The effects of random modulation on such the lowest limit are quantitatively analyzed from an information-theoretic perspective. It is manifested that random modulation does not harm the potential of extracting secret keys from the CCRSM’s randomness component stemming from turbulence-induced optical fluctuations.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

Corrections

Chunyi Chen and Qiong Li, "Enhanced secret-key generation from atmospheric optical channels with the use of random modulation: erratum," Opt. Express 31, 513-513 (2023)
https://opg.optica.org/oe/abstract.cfm?uri=oe-31-1-513

1. Introduction

Secure communications are currently an issue that concerns numerous scientists and engineers over the world because the demand for confidential information transmission between distantly separated communicating parties spawns every day [1,2]. Secret keys shared by the communicating parties are necessary when symmetric encryption is adopted to protect the confidential information transmission against malicious eavesdropping. Hence, secret-key sharing is important for secure communications based on symmetric encryption. Secret-key agreement has been initially suggested in [3,4] to share common keys between two communicating parties with information-theoretic security. To date, different methods for implementing the secret-key agreement have been reported in the literature [525], e.g., the quantum key distribution (QKD) [57], classical optical key distribution (OKD) [813] and channel-based key extraction (CKE) [1425]. These methods have been actively studied in recent years. Generally speaking, the QKD is currently still technically demanding and costly, especially when it operates in an atmospheric wireless environment. In the classical OKD, initial random bit sequences are first transmitted from the sender to the receiver by using classical optical signals; subsequently, the sender and receiver agree on the final keys based on the transmitted random bit sequences. The CKE, which is another distinctive paradigm of secret-key agreement, treats a reciprocal random channel as a common randomness source and generates shared keys from it. If we take into account different application scenarios, the QKD, OKD and CKE should be regarded as three complementary technologies for secret-key sharing with information-theoretic security.

Until now, much research work [1925] on the CKE has utilized the common randomness source provided by reciprocal radio-frequency (RF) wireless channels. Besides, generation of secret keys from optical wireless channels through atmospheric turbulence also attracts much attention in recent years [1418]. The turbulence-induced perturbation can cause randomness in optical wireless channels. It has been proved both theoretically and experimentally [26,27] that single-mode atmospheric optical channels can maintain fading reciprocity. Indeed, extracting secret keys from atmospheric optical channels is a promising solution to long-distance secret-key agreement between two terminals operating in the earth’s atmosphere in certain cases, such as those in which other form of secret-key agreement is unavailable. It provides us with a paradigm of utilizing turbulence-induced optical perturbation to benefit confidential information transmission.

The majority of reported investigation concerning the CKE [15,16,19,20,23] has utilized scalar quantization to change channel measurements to discrete key bits because implementing it is relatively straightforward. Statistical correlation between consecutive channel measurements generally results in statistical correlation between bits in a key generated by using a scalar quantization scheme to quantize each channel measurement separately [24,25]. Although a key bit sequence can be randomized by performing privacy amplification [1], making the randomness of the key bit sequence as high as possible before carrying out privacy amplification is still paramount. From this point of view, statistical uncorrelation between consecutive channel measurements is really preferred [24,25]. Generally, the measurement rate of the channel state is required to be decreased to a proper value depending on the channel’s correlation time for ensuring the uncorrelation between consecutive channel measurements; this requirement imposes a restriction on the number of extracted key bits per second. To break this restriction, researchers [24,25] first performed a preprocessing operation to decorrelate the channel measurements and then carried out scalar quantization on every decorrelated entry separately; although such method is effective in decorrelating the channel measurements, the covariance matrix of channel estimation vectors and its singular value decomposition (SVD) are needed to be evaluated and the obtained eigenvectors are required to be transmitted from one legitimate party to the other legitimate party over a public channel, thus leading to additional computational complexity and transmission cost.

Physically, correlation between consecutive channel measurements can be attributed to the lack of random variations in these measurements. Note that, atmospheric optical channel fading seldom changes within the correlation time of turbulence-induced optical fluctuations [28]. If the sampling interval employed in the measurement of the channel fading is smaller than such correlation time, consecutive channel measurements are inevitably statistically correlated. Intuitively, if we can make the optical signal intensity output from the transmitter of an optical channel fluctuate randomly beforehand, with a correlation time much smaller than the correlation time of turbulence-induced optical fluctuations, consecutive channel measurements obtained with a sampling interval much smaller than the correlation time of turbulence-induced optical fluctuations are possible to be statistically uncorrelated. Random fluctuations can be introduced into the initial transmitted optical-signal intensity by use of random modulation. With these observations in mind, we become aware of the possibility of decorrelating consecutive channel measurements and hence increasing the number of extracted uncorrelated key bits per second by utilizing random modulation. Here it should be pointed out that use of random modulation at the two legitimate communicating terminals gives rise to the problem that the cross-correlation between the optical-signal fluctuations observed by the two legitimate communicating parties may decrease significantly, which will become apparent later, hence making the secret-key extraction unavailable. To address this problem, we will propose, in Sec. 2, a novel randomness sharing scheme for enhanced secret-key extraction and thereby construct a new type of common randomness source, which is compatible with the random modulation.

To comprehend how random modulation impacts enhanced secret-key extraction from our new common randomness source that can break the aforesaid restriction on the number of extracted uncorrelated key bits per second, it is necessary to develop theoretical models for both the temporal autocorrelation of our randomness source’s fluctuations observed by the legitimate parties and the cross-correlation between our randomness source’s fluctuations observed by the two legitimate parties and between our randomness source’s fluctuations observed by the legitimate party and eavesdropper. Furthermore, it is worthwhile to examine the effects of random modulation on the enhanced secret-key extraction from an information-theoretic perspective; thereby one can clarify how random modulation affects the fundamental limits on the secret-key capacity. The above consideration motivates the work presented here. The remainder of this paper is organized as follows. First, in Sec. 2, we elaborate our novel key-extraction oriented randomness sharing scheme and construct the relevant new type of common randomness source theoretically. Second, we develop, in Sec. 3, the correlation models for the legitimate parties’ observations of such common randomness source. Subsequently, we deal with the eavesdropping risk and secret-key capacity in consideration of using our randomness sharing scheme in Secs. 4 and 5, respectively. Finally, conclusions are drawn in Sec. 6.

2. Key-extraction oriented randomness sharing using an atmospheric optical channel with random modulation

The objective of the CKE is to let two legitimate communicating parties agree on a key by exploiting the channel’s inherent randomness shared by them, while concealing the key from an eavesdropper. Randomness sharing is the first step in an actual implementation of secret-key agreement. After carrying out the randomness sharing step, the legitimate parties obtain a given number of their own observations of the shared common randomness. In this section, we propose a novel randomness sharing scheme that uses atmospheric optical channels equipped with random modulation to produce a common randomness source whose temporal autocorrelation can be decreased by the legitimate parties via performing random modulation; later we refer to the source of common randomness generated by our randomness sharing scheme as controllable common randomness source with memory (CCRSM).

To proceed further, let us consider a reciprocal single-mode bidirectional optical channel through the atmosphere illustrated by Fig. 1, where the two legitimate terminals, referred to as Alice and Bob, respectively, can transmit and receive optical signals simultaneously, and a passive eavesdropper, called Eve, can also collect the optical signal waves, coming from Alice and Bob, outside the receiving pupils of the legitimate terminals. In Fig. 1, Eve is located in proximity to Bob. Apparently, Eve can also be located close to Alice. However, in our subsequent theoretical development, we assume that Eve is situated close to the right terminal (i.e., Bob). We emphasize here that this assumption does not impose any restrictions on the generality of our theoretical formulations, because the formation of the left and right terminals is identical and they are designated arbitrarily; i.e., we can also refer to the left terminal as Bob and the right terminal as Alice. With respect to Eve’s eavesdropping action, we comment that, from a practical standpoint, it is extremely difficult for Eve to intercept the main lobe of a propagated laser beam without revelation of her malicious action because of the line-of-sight nature of free-space optical propagation [9]. Hence, in this work, we take into account the practically feasible eavesdropping case shown by Fig. 1, where Eve is situated at a position of the divergence region of propagated optical beams, and collects the optical signal waves outside the legitimate terminals’ receiving pupils.

 figure: Fig. 1.

Fig. 1. Schematic diagram of a reciprocal single-mode bidirectional atmospheric optical channel equipped with random modulation.

Download Full Size | PDF

The optical signal received by Alice at time t can be expressed by

$${y_A}(t) = {s_T}{x_B}(t){h_{BA}}(t), $$
where sT is the pristine optical power output by the legitimate transceivers, xB(t) denotes a random modulation signal, and hBA(t) is the nonnegative fluctuating transmission coefficient at time t of the atmospheric optical channel from Bob to Alice. Similarly, the optical signal received by Bob at time t can be expressed by
$${y_B}(t) = {s_T}{x_A}(t){h_{AB}}(t), $$
where xA(t) denotes a random modulation signal, and hAB(t) is the nonnegative fluctuating transmission coefficient at time t of the atmospheric optical channel from Alice to Bob. Here we assume that sT does not vary with time. We consider that intensity modulation is employed in Fig. 1; thus it is required that 0 ≤ xA(t) ≤ 1 and 0 ≤ xB(t) ≤ 1. Moreover, hereinafter we assume hAB(t) ≡ hBA(t) ≡ hM(t) due to the fact that the single-mode optical channel is reciprocal; the role of detrimental effects that practical hardware imperfection has on the identity of hAB(t) and hBA(t) will be formally incorporated into noises later. One can find from Eqs. (1) and (2) that yA(t) is not inevitably identical to yB(t) because xA(t) and xB(t) are two independent signals and are not necessarily identical at time t.

With Eqs. (1) and (2) in hand, we can further formulate the photocurrent output by the legitimate party’s photodetector as

$${\tilde{y}_X}(t) = R \cdot {y_X}(t )+ {n_X}(t ), $$
where the subscript X = “A” for Alice and the subscript X = “B” for Bob; R denotes the photodetector’s responsibility; nX(t) is a random process of additive white Gaussian noise (AWGN) with zero mean. To construct a common randomness source shared by Alice and Bob, we let
$${\phi _X}(t )= {x_X}(t ){\tilde{y}_X}(t), $$
where the subscript X = “A” for Alice and the subscript X = “B” for Bob. We comment that ${\phi _A}(t)$ and ${\phi _B}(t)$ can be obtained straightforwardly by Alice and Bob, respectively, because xA(t) and xB(t) are modulation signals generated by Alice and Bob, respectively. In this work, we treat the sampling values of ${\phi _A}(t)$ and ${\phi _B}(t)$ as Alice’s and Bob’s observations of the CCRSM generated by the single-mode atmospheric optical channel shown by Fig. 1, respectively. It can be found by examination of Eq. (4) that ${\phi _A}(t)$ and ${\phi _B}(t)$ are statistically correlated; if nA(t) ≡ nB(t) ≡ 0, the equality ${\phi _A}(t)$${\phi _B}(t)$ holds true. As a result, use of the reciprocal single-mode bidirectional atmospheric optical channel equipped with random modulation displayed by Fig. 1 can implement randomness sharing between two legitimate parties, and meanwhile the legitimate parties are capable of controlling the shared randomness.

We note that Eve in Fig. 1 can simultaneously collect optical signals coming from Alice and Bob, respectively, at her own position. Such eavesdropping action of Eve may compromise the secrecy of the legitimate parties’ key extraction. Theoretical development and analysis from the eavesdropper’s perspective will be presented in Sec. 4.

3. Correlation models for legitimate parties’ observations of CCRSM

To clarify the distinctive nature of the CCRSM, in this section, we theoretically characterize the autocorrelation and cross-correlation behaviors of ${\phi _A}(t)$ and ${\phi _B}(t)$. By theoretically formulating the normalized temporal autocovariance function of ${\phi _A}(t)$ and ${\phi _B}(t)$, we can examine how random modulation impacts the correlation between Alice’s (and Bob’s) consecutive observations of the CCRSM quantitatively. The expression for the normalized temporal autocovariance function of ${\phi _A}(t)$ can be developed to give

$${b_{\phi ,A}}(\tau )= \frac{{[{{b_{x,A}}(\tau )\sigma_{x,A}^2 + 1} ][{{b_{x,B}}(\tau )\sigma_{x,B}^2 + 1} ][{{b_h}(\tau )\sigma_h^2 + 1} ]- 1 + {b_d}(\tau )\gamma _A^{ - 1}({\sigma_{x,A}^2 + 1} )}}{{({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )({\sigma_h^2 + 1} )- 1 + \gamma _A^{ - 1}({\sigma_{x,A}^2 + 1} )}}, $$
where
$$\sigma _{x,A}^2 = {{\left\langle {x_A^2(t )} \right\rangle } / {{{\left\langle {{x_A}(t )} \right\rangle }^2} - 1}}, $$
$$\sigma _{x,B}^2 = {{\left\langle {x_B^2(t )} \right\rangle } / {{{\left\langle {{x_B}(t )} \right\rangle }^2} - 1}}$$
$$\sigma _h^2 = {{\left\langle {h_M^2(t )} \right\rangle } / {{{\left\langle {{h_M}(t )} \right\rangle }^2} - 1}}, $$
$${\gamma _A} = {{{R^2}s_T^2{{\left\langle {{h_M}(t )} \right\rangle }^2}{{\left\langle {{x_B}(t )} \right\rangle }^2}} / {\sigma _{n,A}^2}}, $$
$${b_{x,A}}(\tau )= \frac{{\left\langle {{x_A}(t ){x_A}({t + \tau } )} \right\rangle - {{\left\langle {{x_A}(t )} \right\rangle }^2}}}{{\left\langle {x_A^2(t )} \right\rangle - {{\left\langle {{x_A}(t )} \right\rangle }^2}}}, $$
$${b_{x,B}}(\tau )= \frac{{\left\langle {{x_B}(t ){x_B}({t + \tau } )} \right\rangle - {{\left\langle {{x_B}(t )} \right\rangle }^2}}}{{\left\langle {x_B^2(t )} \right\rangle - {{\left\langle {{x_B}(t )} \right\rangle }^2}}}, $$
$${b_h}(\tau )= \frac{{\left\langle {{h_M}(t ){h_M}({t + \tau } )} \right\rangle - {{\left\langle {{h_M}(t )} \right\rangle }^2}}}{{\left\langle {h_M^2(t )} \right\rangle - {{\left\langle {{h_M}(t )} \right\rangle }^2}}}, $$
bd(τ) = δ(τ)/δ(0), δ(τ) denotes Dirac delta function, the angle brackets represent ensemble average, and $\sigma _{n,A}^2$ is the variance of random process nA(t). Similarly, we can derive the expression for the normalized temporal autocovariance function of ${\phi _B}(t)$ and obtain
$${b_{\phi ,B}}(\tau )= \frac{{[{{b_{x,A}}(\tau )\sigma_{x,A}^2 + 1} ][{{b_{x,B}}(\tau )\sigma_{x,B}^2 + 1} ][{{b_h}(\tau )\sigma_h^2 + 1} ]- 1 + {b_d}(\tau )\gamma _B^{ - 1}({\sigma_{x,B}^2 + 1} )}}{{({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )({\sigma_h^2 + 1} )- 1 + \gamma _B^{ - 1}({\sigma_{x,B}^2 + 1} )}}$$
with
$${\gamma _B} = {{{R^2}s_T^2{{\left\langle {{h_M}(t )} \right\rangle }^2}{{\left\langle {{x_A}(t )} \right\rangle }^2}} / {\sigma _{n,B}^2}}, $$
where $\sigma _{n,B}^2$ is the variance of random process nB(t). To clarify the physical meaning of the quantities shown by Eqs. (6)–(12) and (14), we point out here that $\sigma _{x,A}^2$ and $\sigma _{x,B}^2$ are the normalized variances of modulation signals xA(t) and xB(t), respectively; $\sigma _h^2$ is the normalized variance of random process hM(t); γA and γB are the average signal-to-noise ratios (SNR) at the outputs of Alice’s and Bob’s photodetectors, respectively. In arriving at Eqs. (5)–(14), we have assumed that the random processes xA(t), xB(t), hM(t), nA(t) and nB(t) are stationary processes, at least in the wide sense. Hence, the time argument t can be omitted for $\sigma _{x,A}^2$, $\sigma _{x,B}^2$, $\sigma _h^2$, γA, and γB. Moreover, we point out that bx,A(τ), bx,B(τ) and bh(τ) are the normalized temporal autocovariance function of xA(t), xB(t) and hM(t), respectively. Furthermore, there exist $\sigma _{x,A}^2$ = 0 and $\sigma _{x,B}^2$ = 0 if xA(t) and xB(t) are not varying with time; under this condition, ${b_{\phi ,A}}(\tau )$ = ${b_{\phi ,B}}(\tau )$bh(τ) in the limit of γA → ∞ and γB → ∞.

If both bx,A(τ)$\sigma _{x,A}^2$ and bx,B(τ)$\sigma _{x,B}^2$ become much smaller than 1 when τ increases beyond a value τth, Eqs. (5) and (13) can be reduced to

$${b_{\phi ,A}}(\tau )\simeq \frac{{{b_h}(\tau )}}{{\zeta ({{\gamma_A},\sigma_{x,A}^2,\sigma_{x,B}^2} )}}$$
and
$${b_{\phi ,B}}(\tau )\simeq \frac{{{b_h}(\tau )}}{{\zeta ({{\gamma_B},\sigma_{x,B}^2,\sigma_{x,A}^2} )}}$$
for τ > τth with
$$\zeta ({\gamma ,\sigma_1^2,\sigma_2^2} )= ({\sigma_1^2 + 1} )({\sigma_2^2 + 1} )+ [{({\sigma_1^2 + 1} )({\sigma_2^2 + 1} )+ {\gamma^{ - 1}}({\sigma_1^2 + 1} )- 1} ]\sigma _h^{ - 2}. $$

We point out that ζ(⋅,⋅,⋅) in Eqs. (15) and (16) indeed functions as a scaling factor of bh(τ). It is evident that ζ(γ, $\sigma _1^2,\sigma _2^2$) ≥ 1 with γ > 0, $\sigma _1^2$ ≥ 0, $\sigma _2^2$ ≥ 0 and $\sigma _h^2$ > 0. It can be found that, with other parameters fixed, ζ(γ, $\sigma _1^2,\sigma _2^2$) enlarges with increasing $\sigma _1^2$ or $\sigma _2^2$. Hence, we can make the temporal autocorrelation of both ${\phi _A}(t)$ and ${\phi _B}(t)$ much smaller than the temporal autocorrelation of hM(t) by using modulation signals whose normalized variance is relatively large and meanwhile whose correlation time is small enough that bx,A(τ)$\sigma _{x,A}^2$≪ 1 and bx,B(τ)$\sigma _{x,B}^2$≪ 1 hold true at a value of τ, viz., the said τth, which is much smaller than the correlation time of hM(t). With the above finding in mind, a conclusion can be made that consecutive observations of both ${\phi _A}(t)$ and ${\phi _B}(t)$ obtained by Alice and Bob, respectively, with a given sampling rate may be decorrelated by increasing the normalized variances of modulation signals xA(t) and xB(t) under the condition that the correlation time of both xA(t) and xB(t) is short enough. Moreover, in practice, the time interval between two consecutive observations of ${\phi _A}(t)$ or ${\phi _B}(t)$ is equal to the temporal sampling interval τs associated with the sampling operation of ${\phi _A}(t)$ or ${\phi _B}(t)$. Hence, decorrelating consecutive observations of ${\phi _X}(t)$ is equivalent to making ${b_{\phi ,X}}({\tau _s})$ tend to a negligible value, with the subscript X = “A” or “B”. From a practical perspective, in the case that the correlation time of both xA(t) and xB(t) is small enough that bx,A(τ)$\sigma _{x,A}^2$≪ 1 and bx,B(τ)$\sigma _{x,B}^2$≪ 1 hold true at a value of τ smaller than τs, the statistical correlation between two consecutive observations of ${\phi _X}(t)$ can be dropped to a negligible level by properly increasing ζ(⋅,⋅,⋅), even though the correlation time of hM(t) is actually a good deal longer than τs. This opens a promising pathway for decorrelating consecutive observations of the CCRSM. Here, we emphasize the following three points: 1) the CCRSM is compatible with off-the-shelf wireless optical communication systems using the intensity modulation and direct detection, and hence is easy to construct; 2) xA(t) and xB(t) are two independent random modulation signals and thus there is no need for synchronization between them; 3) even though the correlation time of hM(t) is larger than τs, the statistical correlation between two consecutive observations of the CCRSM can be reduced to being negligible by using appropriate random modulation, which does not suffer from the additional computational complexity and transmission cost as in the SVD-based decorrelation operation [24,25].

Below we give a concrete example for demonstration of decorrelating two consecutive observations of ${\phi _A}(t)$ by increasing the normalized variance of the modulation signal. To evaluate the correlation coefficient of two consecutive observations, the normalized temporal autocovariance function bh(τ) needs to be specified explicitly. For the example given here, we assume that bh(τ) takes a form of Gaussian-type function, i.e., bh(τ) = exp(−τ2/τh,02), where τh,0 is the correlation time of hM(t), which is defined as the time delay τ required for bh(τ) to reach e−1 [29]. Figure 2 illustrates how the growth in $\sigma _{x,A}^2$ affects the correlation coefficient of two consecutive observations of ${\phi _A}(t)$ under the condition that the correlation time of the random modulation signal is rather smaller than the temporal sampling interval τs; Fig. 2(a) corresponds to the single-modulation case in which xA(t) changes randomly with time whereas xB(t) remains constant; Fig. 2(b) corresponds to the double-modulation case in which both xA(t) and xB(t) change randomly with time. It is seen from Fig. 2 that the correlation coefficient of two consecutive observations of ${\phi _A}(t)$ diminishes considerably with increasing $\sigma _{x,A}^2$ for both the single- and double-modulation cases. Comparison of Fig. 2(a) with Fig. 2(b) shows that use of double modulation makes the correlation coefficient decrease more rapidly with increasing $\sigma _{x,A}^2$ than use of single modulation. This phenomenon can be readily justified by examination of Eqs. (15) and (17); because of ζ(γA, $\sigma _{x,A}^2,\sigma _{x,B}^2$≡0) < ζ(γA, $\sigma _{x,A}^2,\sigma _{x,B}^2 \equiv \sigma _{x,A}^2$), ${b_{\phi ,A}}({\tau _s})$ with $\sigma _{x,B}^2$≡ 0 is greater than that with $\sigma _{x,B}^2$$\sigma _{x,A}^2$. Moreover, one can find from Fig. 2 that the decrease in ${b_{\phi ,A}}({\tau _s})$ with $\sigma _h^2$ = 0.3 is less substantial than that with $\sigma _h^2$ = 0.1; specifically, for the single-modulation case, ${b_{\phi ,A}}({\tau _s})$ with $\sigma _h^2$ = 0.3 becomes smaller than 0.2 when $\sigma _{x,A}^2$ increases to 1, whereas ${b_{\phi ,A}}({\tau _s})$ with $\sigma _h^2$ = 0.1 decreases below 0.1 only when $\sigma _{x,A}^2$ grows beyond 0.81; for the double-modulation case, ${b_{\phi ,A}}({\tau _s})$ with $\sigma _h^2$ = 0.3 begins to be smaller than 0.2 when $\sigma _{x,A}^2$ grows beyond 0.39, whereas ${b_{\phi ,A}}({\tau _s})$ with $\sigma _h^2$ = 0.1 is below 0.1 only when $\sigma _{x,A}^2$ grows beyond 0.35. This finding reveals that the normalized variance $\sigma _h^2$ of hM(t) has an important effect on the reduction in the correlation coefficient caused by increasing $\sigma _{x,A}^2$; smaller $\sigma _h^2$ results in greater reduction in the correlation coefficient with other parameters fixed. The underlying physical reason for this fact is elucidated as follows. Smaller $\sigma _h^2$ implies weaker random fluctuations in hM(t) and hence hM(t) plays a less important role in determining the temporal autocorrelation of ${\phi _A}(t)$; in this case, the temporal autocorrelation behavior of xA(t) becomes more prominent in determining the temporal autocorrelation behavior of ${\phi _A}(t)$.

 figure: Fig. 2.

Fig. 2. Correlation coefficient of two consecutive observations of ${\phi _A}(t)$, i.e., ${b_{\phi ,A}}({\tau _s})$, in terms of $\sigma _{x,A}^2$ with γA ≡ ∞ and different $\sigma _h^2$; (a) the single-modulation case with $\sigma _{x,A}^2$ > 0 and $\sigma _{x,B}^2$ ≡ 0; (b) the double-modulation case with $\sigma _{x,A}^2$$\sigma _{x,B}^2$> 0. The correlation time τh,0 and the temporal sampling interval τs are specified as 1 ms and 2.5×10−5 ms, respectively; the correlation time of the random modulation signal is small enough that bx,A(τ)$\sigma _{x,A}^2$≪ 1 and bx,B(τ)$\sigma _{x,B}^2$≪ 1 hold true at a value of τ smaller than τs.

Download Full Size | PDF

Now we turn our attention to the role of random modulation in determining the cross-correlation between the observations of ${\phi _A}(t)$ and ${\phi _B}(t)$ obtained by Alice and Bob, respectively. The cross-correlation coefficient of ${\phi _A}(t)$ and ${\phi _B}(t)$ can be expressed by

$${\rho _{AB}} = \frac{{\left\langle {\left[ {{\phi_A}(t )- \left\langle {{\phi_A}(t )} \right\rangle } \right]\left[ {{\phi_B}(t )- \left\langle {{\phi_B}(t )} \right\rangle } \right]} \right\rangle }}{{\sqrt {\left\langle {{{\left[ {{\phi_A}(t )- \left\langle {{\phi_A}(t )} \right\rangle } \right]}^2}} \right\rangle \left\langle {{{\left[ {{\phi_B}(t )- \left\langle {{\phi_B}(t )} \right\rangle } \right]}^2}} \right\rangle } }}.$$

The time argument t is omitted for ρAB because of the aforementioned assumption of stationary processes. Introducing Eq. (4) into Eq. (18) and evaluating the ensemble averages therein lead us to

$$\begin{array}{c} {\rho _{AB}} = \frac{1}{{\sqrt {1 + \gamma _A^{ - 1}{{[{({\sigma_{x,B}^2 + 1} )({\sigma_h^2 + 1} )- {1 / {({\sigma_{x,A}^2 + 1} )}}} ]}^{ - 1}}} }}\\ \times \frac{1}{{\sqrt {1 + \gamma _B^{ - 1}{{[{({\sigma_{x,A}^2 + 1} )({\sigma_h^2 + 1} )- {1 / {({\sigma_{x,B}^2 + 1} )}}} ]}^{ - 1}}} }} \end{array}.$$

It is apparent that ρAB depends on γA and γB. If both γA and γB tend to infinity, it follows that ρAB → 1. In fact, the normalized variances $\sigma _{x,A}^2$, $\sigma _{x,B}^2$ and $\sigma _h^2$ are combined together in Eq. (19) to act as a scaling factor of γA and γB. For fixed γA and γB, enlarging either $\sigma _{x,A}^2$ or $\sigma _{x,B}^2$ causes an increase in ρAB, i.e., random modulation increases the cross-correlation coefficient ρAB.

4. Analysis of eavesdropping risk in CCRSM-based secret-key extraction

As shown by Fig. 1, besides Alice and Bob, Eve can also obtain measurements of her collected optical signals. Similar to the mathematical development of Eq. (3), we can formulate the photocurrent produced by Eve’s terminal due to her collected optical signal as follows:

$${\tilde{y}_{XE}}(t )= {R_E}{s_T}{x_X}(t ){h_{XE}}(t )+ {n_{XE}}(t ), $$
where X = “A” and “B” for the cases of Eve collecting optical signal coming from Alice and Bob, respectively; RE is the responsibility of Eve’s photodetector, hXE(t) denotes the nonnegative fluctuating transmission coefficient at time t of the atmospheric optical channel from Alice (X = “A”) or Bob (X = “B”) to Eve, and nXE(t) is a random process of AWGN with zero mean. It has been shown that the spatial coherence length of the atmosphere is generally on the order of several centimeters [28]. In practice, it is really hard for Eve to stay at a position separated from a legitimate party by a distance on the order of several centimeters while keeping the legitimate party unaware of her presence. With such practical restriction in mind, it is reasonable to assume that both hAE(t) and hBE(t) are statistically independent of hM(t). Hence, we think, in this work, that hAE(t) and hBE(t) do not carry any useful information about the instantaneous state of the CCRSM.

The secrecy of keys extracted from observations of the CCRSM counts on the lack of cross-correlation between the legitimate parties’ and eavesdropper’s observations. To assess the eavesdropping risk in secret-key extraction from observations of the CCRSM, below we inspect whether the random modulation makes ${\tilde{y}_{AE}}(t)$ and ${\tilde{y}_{BE}}(t)$ become partially correlated with ${\phi _A}(t)$ and ${\phi _B}(t)$. For this purpose, we first formulate the statistical cross-correlation between ${\tilde{y}_{XE}}(t)$ and ${\phi _A}(t)$ and between ${\tilde{y}_{XE}}(t)$ and ${\phi _B}(t)$, with the subscript X = “A” or “B”. The cross-correlation coefficient of ${\tilde{y}_{XE}}(t)$ and ${\phi _A}(t)$ can be formulated by

$${\rho _{XE,A}} = {{\sigma _{x,X}^2} / {{\chi _A}}}$$
with
$${\chi _A} = \sqrt {({\sigma_{x,X}^2 + 1} )({\sigma_{h,XE}^2 + 1} )- 1 + \gamma _{XE}^{ - 1}} \cdot \sqrt {({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )({\sigma_h^2 + 1} )- 1 + ({\sigma_{x,A}^2 + 1} )\gamma _A^{ - 1}}, $$
$$\sigma _{h,XE}^2 = {{\left\langle {h_{XE}^2(t )} \right\rangle } / {{{\left\langle {{h_{XE}}(t )} \right\rangle }^2} - 1}}, $$
$${\gamma _{XE}} = {{R_E^2s_T^2{{\left\langle {{h_{XE}}(t )} \right\rangle }^2}{{\left\langle {{x_X}(t )} \right\rangle }^2}} / {\sigma _{n,XE}^2}}, $$
where $\sigma _{n,XE}^2$ is the variance of random process nXE(t). We note that $\sigma _{h,XE}^2$ is the normalized variance of random process hXE(t), and ${\gamma _{XE}}$ is the average SNR at the output of Eve’s photodetector. Similarly, the cross-correlation coefficient of ${\tilde{y}_{XE}}(t)$ and ${\phi _B}(t)$ can be formulated as follows:
$${\rho _{XE,B}} = {{\sigma _{x,X}^2} / {{\chi _B}}}$$
with
$${\chi _B} = \sqrt {({\sigma_{x,X}^2 + 1} )({\sigma_{h,XE}^2 + 1} )- 1 + \gamma _{XE}^{ - 1}} \cdot \sqrt {({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )({\sigma_h^2 + 1} )- 1 + ({\sigma_{x,B}^2 + 1} )\gamma _B^{ - 1}}.$$
We point out once again that the subscript X = “A” or “B” in Eqs. (21)–(26); this notation will be used later without explicit explication. One can find from Eqs. (21)–(26) that ρXE,A = ρXE,B if $\sigma _{x,A}^2$ =$\sigma _{x,B}^2$ and ${\gamma _A}$=${\gamma _B}$.

It can be found from Eqs. (21) and (25) that both ρXE,A and ρXE,B are equal to zero if the modulation signal xX(t) does not vary with time, and are larger than 0 otherwise. This fact reveals that, if xA(t) does not remain constant with time, Eve can acquire some information about both ${\phi _A}(t)$ and ${\phi _B}(t)$ by observing, at her own position, the fluctuations in the optical signal coming from Alice; if xB(t) varies with time, Eve can gain some information about both ${\phi _A}(t)$ and ${\phi _B}(t)$ by observing, at her own position, the fluctuations in the optical signal coming from Bob. The underlying reason for this is that the changes in ${\phi _A}(t)$, ${\phi _B}(t)$ and ${\tilde{y}_{XE}}(t)$ are dependent on the variation of the modulation signal xX(t); in other words, the variation of xX(t) is shared by ${\phi _A}(t)$, ${\phi _B}(t)$ and ${\tilde{y}_{XE}}(t)$. On the other hand, if xX(t) is a constant signal, i.e., $\sigma _{x,X}^2$ ≡ 0, it follows that ρXE,A = ρXE,B ≡ 0. Hence, a conclusion can be made that the random modulation with use of a time-varying signal does result in some information leakage of the CCRSM to the eavesdropper. Incidentally, the detrimental effects of such information leakage on the secrecy of extracted keys can be eliminated by carrying out privacy amplification after the information reconciliation phase.

The preceding analysis reveals that it is Eve’s acquisition of information about the modulation signal that renders ${\tilde{y}_{AE}}(t)$ and (or) ${\tilde{y}_{BE}}(t)$ partially correlated with ${\phi _A}(t)$ and ${\phi _B}(t)$. Intuitively, if Eve knows the modulation signal exactly, she is able to gain the largest amount of information about ${\phi _A}(t)$ and ${\phi _B}(t)$. In recognition of this, we can conceive the worst eavesdropping scenario in which Eve, located in the vicinity of Bob, simultaneously collects the optical signals coming from Alice and Bob, respectively, and, however, the propagation distance from Bob to Eve is not large enough to produce turbulence-induced optical fluctuations, implying that fluctuations in the optical signal observed by Eve coming from Bob can be completely attributed to xB(t), i.e., fluctuations in the modulation signal xB(t) are exactly known by Eve. We think that Eve has perfect photodetectors without noises in the worst eavesdropping scenario. In what follows, we concentrate our attention only on the said worst eavesdropping scenario. Notice that, as mentioned earlier, the requirement of keeping legitimate parties unaware of Eve’s existence makes it reasonable to assume that both hAE(t) and hBE(t) are statistically independent of hM(t). In the worst eavesdropping scenario, the legitimate parties can adopt either the single modulation scheme or the double modulation scheme. In the single modulation scheme, only one of the two legitimate parties performs the random modulation, whereas, in the double modulation scheme, both legitimate parties perform the random modulation. Below, we deal with the two schemes separately.

For the single modulation scheme with xA(t) ≡ 1 and randomly time-varying xB(t), it follows that ρAE,A = ρAE,B ≡ 0, ρBE,A > 0 and ρBE,B > 0. In the worst eavesdropping scenario with xA(t) ≡ 1 and randomly time-varying xB(t), the random modulation is actually adopted only by Bob, so one can find that ${\tilde{y}_{AE}}(t)$ is independent of both ${\phi _A}(t)$ and ${\phi _B}(t)$, whereas ${\tilde{y}_{BE}}(t)$, equal to ${R_E}{s_T}{\tilde{h}_{BE}}{x_B}(t)$, is partially correlated with both ${\phi _A}(t)$ and ${\phi _B}(t)$ even though hM(t), xB(t), nA(t) and nB(t) are independent random quantities, where ${\tilde{h}_{BE}}$ is the deterministic transmission coefficient of the optical channel from Bob to Eve. Later we refer to the worst eavesdropping scenario with xA(t) ≡ 1 and randomly time-varying xB(t) as “fully-disclosed-single-modulation worst eavesdropping (FDSM-WE) scenario” for the reason that Eve knows the variation of xB(t) exactly. On the other hand, for the worst eavesdropping scenario with xB(t) ≡ 1 and randomly time-varying xA(t), it can be found that the random modulation is carried out only by Alice, and ${\tilde{y}_{AE}}(t)$ is partially correlated with both ${\phi _A}(t)$ and ${\phi _B}(t)$, whereas ${\tilde{y}_{BE}}(t)$ remains constant and is independent of both ${\phi _A}(t)$ and ${\phi _B}(t)$, implying that ρBE,A = ρBE,B ≡ 0, ρAE,A > 0 and ρAE,B > 0. Thereinafter, the worst eavesdropping scenario with xB(t) ≡ 1 and randomly time-varying xA(t) is called “partially-disclosed-single-modulation worst eavesdropping (PDSM-WE) scenario” for the reason that ${\tilde{y}_{AE}}(t)$ is only partially correlated with xA(t).

Figure 3 illustrates the cross-correlation coefficients of ${\tilde{y}_{BE}}(t)$ and ${\phi _X}(t)$ and of ${\tilde{y}_{AE}}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,B}^2$ and $\sigma _{x,A}^2$ in the FDSM-WE and PDSM-WE scenarios, respectively, where the subscript X = “A” or “B”. It can be seen from Fig. 3 that the cross-correlation coefficients enlarge with increasing $\sigma _{x,B}^2$ or $\sigma _{x,A}^2$, implying that enlarging the normalized variance of the randomly time-varying modulation signal, with other parameters fixed, leads to an increase in the said cross-correlation coefficients. This fact demonstrates that Eve can gain a larger proportion of information about both ${\phi _A}(t)$ and ${\phi _B}(t)$ when $\sigma _{x,B}^2$ (or $\sigma _{x,A}^2$) becomes greater in the FDSM-WE (or PDSM-WE) scenario. The physical reason for this phenomenon is that the role of the modulation signal’s variation, which can be perceived somewhat by Eve, becomes more prominent in determining the aforementioned cross-correlation coefficients when $\sigma _{x,B}^2$ (or $\sigma _{x,A}^2$) turns greater. One can see from Fig. 3 that each curve therein, with increasing $\sigma _{x,B}^2$ or $\sigma _{x,A}^2$, saturates to a certain level depending on the relevant specific parameter combinations. In fact, it is easy to find from Eqs. (21) and (25) that ρBE,A and ρBE,B tend to ($\sigma _h^2$+1)−1/2 and ($\sigma _h^2$+1+$\gamma _B^{ - 1}$)−1/2, respectively, when $\sigma _{x,B}^2$ → ∞ in the FDSM-WE scenario, and ρAE,A and ρAE,B tend to [($\sigma _{h,AE}^2$+1)($\sigma _h^2$+1+$\gamma _A^{ - 1}$)]−1/2 and [($\sigma _{h,AE}^2$+1)($\sigma _h^2$+1)]−1/2, respectively, when $\sigma _{x,A}^2$ → ∞ in the PDSM-WE scenario. Hence, the saturated level of the curves in Figs. 3(a) and 3(b) depends only on $\sigma _h^2$ and γB; the saturated level of the curves in Figs. 3(c) and 3(d) is determined only by $\sigma _h^2$, $\sigma _{h,AE}^2$ and γA. For instance, if $\sigma _{h,AE}^2$=$\sigma _h^2$=0.3, ρAE,B in the PDSM-WE scenario is always not larger than 1.3−1 no matter what value $\sigma _{x,A}^2$ takes on. It is straightforward to infer that greater $\sigma _h^2$ leads to smaller cross-correlation coefficients with other parameters fixed. From this observation, one can roughly predict that greater $\sigma _h^2$ results in a smaller leakage proportion of information about ${\phi _A}(t)$ and ${\phi _B}(t)$. Moreover, comparison of Figs. 3(a) and 3(b) with Figs. 3(c) and 3(d) reveals that, with the same normalized variance of the fluctuating channel transmission coefficient and normalized variance of the randomly time-varying modulation signal, ρBE,X in the FDSM-WE scenario is larger than ρAE,X in the PDSM-WE scenario. This roughly means that, with identical normalized variance of the random modulation signals and fixed $\sigma _h^2$, the information leakage to Eve in the PDSM-WE scenario is smaller than that in the FDSM-WE scenario.

 figure: Fig. 3.

Fig. 3. Cross-correlation coefficients of ${\tilde{y}_{BE}}(t)$ and ${\phi _X}(t)$ and of ${\tilde{y}_{AE}}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,B}^2$ and $\sigma _{x,A}^2$ in the FDSM-WE and PDSM-WE scenarios, respectively, where the subscript X = “A” or “B”; ${\gamma _A}$=${\gamma _B}$=20, ${\gamma _{AE}}$= ∞, and ${\gamma _{BE}}$= ∞; (a) and (b) correspond to the FDSM-WE scenario where ${\tilde{y}_{BE}}(t)$=${R_E}{s_T}{\tilde{h}_{BE}}{x_B}(t)$ and ${\tilde{y}_{AE}}(t)$ is independent of ${\phi _X}(t)$; (c) and (d) correspond to the PDSM-WE scenario where ${\tilde{y}_{AE}}(t)$ is partially correlated with ${\phi _X}(t)$ and ${\tilde{y}_{BE}}(t)$ is independent of ${\phi _X}(t)$. (a) $\sigma _h^2 = 0.3$, $\sigma _{h,BE}^2$=0, and $\sigma _{x,A}^2$=0; (b) $\sigma _h^2 = 0.1$, $\sigma _{h,BE}^2$= 0, and $\sigma _{x,A}^2$= 0; (c) $\sigma _h^2 = 0.3$, $\sigma _{h,AE}^2$=$\sigma _h^2$, and $\sigma _{x,B}^2$=0; (d) $\sigma _h^2 = 0.1$, $\sigma _{h,AE}^2$=$\sigma _h^2$, and $\sigma _{x,B}^2$= 0.

Download Full Size | PDF

Now we turn to dealing with the double modulation scheme. For the worst eavesdropping scenario with both xA(t) and xB(t) being randomly time-varying, it is evident that ρAE,A > 0, ρAE,B > 0, ρBE,A > 0 and ρBE,B > 0. Hereinafter, we refer to the worst eavesdropping scenario in which both Alice and Bob perform random modulation independently as “double-modulation worst eavesdropping (DM-WE) scenario”. For the DM-WE scenario, besides ${\tilde{y}_{AE}}(t)$ and ${\tilde{y}_{BE}}(t)$, Eve can construct a combined version of her observations of the CCRSM by multiplying ${\tilde{y}_{AE}}(t)$ and ${\tilde{y}_{BE}}(t)$ together, and thereby obtaining ${\phi _E}(t)$=$R_E^2s_T^2{\tilde{h}_{BE}}$xA(t)xB(t)hAE(t). It should be emphasized here that the output of Eve’s photodetectors is assumed to be noiseless in the worst eavesdropping scenario. In analogy with the derivation of Eqs. (21) and (25), we can formulate the cross-correlation coefficient of ${\phi _E}(t)$ and ${\phi _X}(t)$ as follows:

$${\rho _{CE,X}} = {{[{({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )- 1} ]} / {{\chi _{CE,X}}}}, $$
where
$$\begin{array}{c} {\chi _{CE,X}} = {[{({\sigma_{x,B}^2 + 1} )({\sigma_{x,A}^2 + 1} )({\sigma_h^2 + 1} )- 1 + ({\sigma_{x,X}^2 + 1} )\gamma_X^{ - 1}} ]^{ - 1/2}}\\ \times {[{({\sigma_{x,A}^2 + 1} )({\sigma_{x,B}^2 + 1} )({\sigma_{h,AE}^2 + 1} )- 1} ]^{ - 1/2}} \end{array}$$
with the subscript X = “A” or “B”.

Figure 4 illustrates the cross-correlation coefficients of ${\tilde{y}_{XE}}(t)$ and ${\phi _A}(t)$, of ${\tilde{y}_{XE}}(t)$ and ${\phi _B}(t)$, and of ${\phi _E}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,A}^2$ ($\sigma _{x,B}^2$), with the subscript X = “A” or “B”, where the said DM-WE scenario is considered; for comparison purposes, the curves related to the FDSM-WE and PDSM-WE scenarios are reproduced in Fig. 4. One finds from Fig. 4 that, in the DM-WE scenario with $\sigma _{x,A}^2$=$\sigma _{x,B}^2$, ρBE,A and ρBE,B are identical, and the same relationship holds both for ρAE,A and ρAE,B and for ρCE,A and ρCE,B. This fact can be readily verified by examination of Eqs. (21)–(26), with the use of γA = γB in mind. It can be seen from Fig. 4 that the cross-correlation coefficients ${\rho _{BE,A}}$, ${\rho _{BE,B}}$, ${\rho _{AE,A}}$ and ${\rho _{AE,B}}$ in the DM-WE scenario first enlarge with increasing $\sigma _{x,A}^2$ and $\sigma _{x,B}^2$ until they reach their peak values and after that they begin to decrease when $\sigma _{x,A}^2$ and $\sigma _{x,B}^2$ increase further; interestingly, these cross-correlation coefficients actually approach zero when $\sigma _{x,A}^2$ and $\sigma _{x,B}^2$ tend to infinity, which can be readily justified by examination of Eqs. (21), (22), (25) and (26). One can see from Fig. 4 that the cross-correlation coefficients ρCE,A and ρCE,B in the DM-WE scenario behave in terms of $\sigma _{x,A}^2$ (or $\sigma _{x,B}^2$) in a way similar to ρBE,A in the FDSM-WE scenario and ρAE,B in the PDSM-WE scenario. Indeed, we can infer that ρCE,A = ρCE,B → [($\sigma _{h,AE}^2$+1)($\sigma _h^2$+1)]−1/2 when both $\sigma _{x,A}^2$ and $\sigma _{x,B}^2$ tend to infinity. Because of $\sigma _{h,BE}^2$= 0 and $\sigma _{h,AE}^2$ > 0 in the DM-WE scenario, it can be observed from Fig. 4 that ρAE,A and ρAE,B are smaller than ρBE,A and ρBE,B. Comparison of the curves associated with the DM-WE scenario and those related to the FDSM-WE and PDSM-WE scenarios reveals that, with the same normalized variance $\sigma _h^2$ of the fluctuating channel transmission coefficient and normalized variance of the randomly time-varying modulation signals, the cross-correlation coefficients related to the DM-WE scenario are all smaller than ρBE,A in the FDSM-WE scenario. By comparing the curves in Fig. 4, one finds that, among the various scenarios involved therein, Eve in the FDSM-WE scenario may acquire observations of the CCRSM that have highest cross-correlation with those obtained by the legitimate parties. From this point of view, we can regard the FDSM-WE scenario as the one in which the degree of eavesdropping risk is highest.

 figure: Fig. 4.

Fig. 4. Cross-correlation coefficients of ${\tilde{y}_{XE}}(t)$ and ${\phi _A}(t)$, of ${\tilde{y}_{XE}}(t)$ and ${\phi _B}(t)$, and of ${\phi _E}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,B}^2$ ($\sigma _{x,A}^2$), with the subscript X = “A” or “B”. γA = γB = 20, γAE = γBE = ∞, $\sigma _{h,BE}^2$= 0, and $\sigma _{h,AE}^2$ = $\sigma _h^2$. (a) $\sigma _h^2$ = 0.3; (b) $\sigma _h^2$ = 0.1. The worst eavesdropping scenario is considered in producing the curves. Both Alice and Bob perform the random modulation independently with $\sigma _{x,B}^2$ = $\sigma _{x,A}^2$.

Download Full Size | PDF

5. Information-theoretic analysis

By examination of Eqs. (1)–(4), one finds that the terms hM(t), xA(t) and xB(t) are three common components of ${\phi _A}(t)$ and ${\phi _B}(t)$. Obviously, it is these three common components that enable high level of cross-correlation between the two legitimate parties’ observations of the CCRSM. Note that, hM(t) is indeed an uncontrollable external source of randomness due to turbulence-induced optical fluctuations, and, on the other hand, xA(t) and xB(t) are two controllable internal sources of randomness due to random modulation performed by the legitimate parties. Although the previous analysis of the eavesdropping risk in Sec. 4 does give us an intuitive understanding of the problem as to what role the random modulation plays in information leakage of the CCRSM to Eve, it is still unclear how the random modulation impacts the common-randomness component of the CCRSM stemming from the turbulence-induced optical fluctuations and what effects the random modulation has on the secret-key capacity. To address these issues, in this section, we examine the secret-key capacity of shared key extraction from the CCRSM.

When xA(t) = xB(t) ≡ 1, there is actually no random modulation at both Alice’s and Bob’s terminals. Below we call this case “no-modulation-with-independent-eavesdropping (NMIE) scenario”. For convenience of later comparison, here we first derive expressions for the secret-key capacity in the NMIE scenario; because both ${\tilde{y}_{AE}}(t)$ and ${\tilde{y}_{BE}}(t)$ are independent of ${\phi _A}(t)$ and ${\phi _B}(t)$ in the NMIE scenario, the corresponding secret-key capacity can be written by [1]

$${C_{NMIE}} = {I_{NMIE}}({{\phi_A};{\phi_B}} ), $$
where ${I_{NMIE}}({\phi _A};{\phi _B})$ is the mutual information of ${\phi _A}(t)$ and ${\phi _B}(t)$ for the NMIE scenario. By letting $\sigma _{n,A}^2$=$\sigma _{n,B}^2$=$\sigma _{n,0}^2$, ${I_{NMIE}}({\phi _A};{\phi _B})$ can be formulated by (see Appendix A)
$${I_{NMIE}}({{\phi_A};{\phi_B}} )= \frac{{{\gamma _0}}}{{2\pi }}\int_{ - \infty }^\infty {\textrm{d}{\varphi _A}\int_{ - \infty }^\infty {\textrm{d}{\varphi _B}\left\{ {{Q_0}({{\varphi_A},{\varphi_B}} ){{\log }_2}\left[ {\frac{{{Q_0}({{\varphi_A},{\varphi_B}} )}}{{{Q_1}({{\varphi_A}} ){Q_1}({{\varphi_B}} )}}} \right]} \right\}} } , $$
where γ0 = R2$s_T^2\bar{h}_M^2{/}\sigma _{n,0}^2$ denotes the average SNR at the legitimate terminals without random modulation, ${\bar{h}_M}$ is the mean of channel-transmission-coefficient random process hM(t),
$${Q_0}({{\varphi_A},{\varphi_B}} )= \int_0^\infty {\exp \left[ { - \frac{{{{({{\varphi_A} - \hat{h}} )}^2} + {{({{\varphi_B} - \hat{h}} )}^2}}}{{2/{\gamma_0}}}} \right]} {p_{\hat{h}}}(\hat{h})\textrm{d}\hat{h},$$
$${Q_1}({{\varphi_X}} )= \int_0^\infty {\exp \left[ { - \frac{{{{({{\varphi_X} - \hat{h}} )}^2}}}{{2/{\gamma_0}}}} \right]} {p_{\hat{h}}}(\hat{h})\textrm{d}\hat{h},$$
${p_{\hat{h}}}(\hat{h})$ represents the probability density function (PDF) of the normalized channel transmission coefficient $\hat{h}$, and X = “A” or “B”. In Eqs. (29) and (30), the argument t of ${\phi _A}(t)$ and ${\phi _B}(t)$ is omitted for succinctness. We emphasize that the normalized channel transmission coefficient $\hat{h}$ is defined as the ratio of the channel transmission coefficient to its mean. Moreover, we point out that the mutual information given by Eq. (30) is a functional of the statistical distributions of ${\phi _A}(t)$ and ${\phi _B}(t)$; hence the right-hand side of Eq. (30) does not include ${\phi _A}(t)$ and ${\phi _B}(t)$. It is seen from Eqs. (30)–(32) that, with given ${p_{\hat{h}}}(\hat{h})$, ${I_{NMIE}}({\phi _A};{\phi _B})$ depends only on γ0. From another point of view, when ${p_{\hat{h}}}(\hat{h})$ takes on a single-peak shape and the normalized variance $\sigma _h^2$ of hM(t) is much smaller than 1/γ0, roughly speaking, the peak of ${p_{\hat{h}}}(\hat{h})$ will be much sharper than that of the exponential functions appearing in the integrands of Eqs. (31) and (32). Under such condition, it follows that Q0(φA, φB) ≈ exp{−γ0[(φA${\hat{h}_m}$)2 + (φB${\hat{h}_m}$)2]/2} and Q1(φX) ≈ $\exp [-\gamma_0(\varphi_X-\hat{h}_m)^2 / 2]$, with ${\hat{h}_m}$ being the value of $\hat{h}$ at which ${p_{\hat{h}}}(\hat{h})$ achieves its maximum, thus implying that the logarithm in Eq. (30) is approximately zero and in turn ${I_{NMIE}}({\phi _A};{\phi _B})$ becomes approximately zero. As a result, ${C_{NMIE}}$ ≈ 0 if $\sigma _s^2{/}\sigma _{n,0}^2\ll1$, where $\sigma _s^2$=R2$s_T^2\bar{h}_M^2\sigma _h^2$. Note that, $\sigma _s^2$ actually characterizes the variance of the output component of a legitimate terminal’s photodetector due to the desired optical signal coming from its legitimate partner. The above observation indeed discloses that, besides γ0, the ratio $\sigma _s^2/\sigma _{n,0}^2$ should play a role in determining ${C_{NMIE}}$. Roughly speaking, $\sigma _s^2/\sigma _{n,0}^2\ll1$ means that the noise is the dominant contributor to the variation in the photodetector’s output signal, hence resulting in ${C_{NMIE}}$ ≈ 0. One can infer from this analysis that, for obtaining a substantial value of ${C_{NMIE}}$ with a given γ0, it is required that random variation in the photodetector’s output signal due to yA(t) or yB(t), instead of the noise, dominates the variation of ${\tilde{y}_A}(t)$ or ${\tilde{y}_B}(t)$.When random modulation is performed by Alice and (or) Bob, ${\tilde{y}_{AE}}(t)$ and (or) ${\tilde{y}_{BE}}(t)$ may become partially correlated with ${\phi _A}(t)$ and ${\phi _B}(t)$, and the relevant secret-key capacity CRM is lower bounded by (see Theorem 4.1 in [1])
$${C_{RM}} \ge {C_{RM,LB}} = {I_{RM}}({{\phi_A};{\phi_B}} )- \min ({{I_{RM}}({{\phi_A};{{\tilde{y}}_{AE}},{{\tilde{y}}_{BE}}} ),{I_{RM}}({{\phi_B};{{\tilde{y}}_{AE}},{{\tilde{y}}_{BE}}} )} ),$$
where ${I_{RM}}({\phi _A};{\phi _B})$ is the mutual information of ${\phi _A}(t)$ and ${\phi _B}(t)$ when random modulation is performed, ${I_{RM}}({\phi _A};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$ represents the mutual information of ${\phi _A}(t)$ and [${\tilde{y}_{AE}}(t)$, ${\tilde{y}_{BE}}(t)$] when random modulation is performed, and ${I_{RM}}({\phi _B};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$ stands for the mutual information of ${\phi _B}(t)$ and [${\tilde{y}_{AE}}(t)$, ${\tilde{y}_{BE}}(t)$] when random modulation is performed. In this section, we focus our attention only on examining the lower bound of the secret-key capacity CRM. The lower bound given by Eq. (33) can be regarded as the remainder of subtracting some information rate leaked to the eavesdropper from the information rate between the two legitimate parties. The “min” operation in Eq. (33) is due to the fact that the legitimate parties can utilize the two-way communication over a public channel to opt whether the information rate leaked to Eve is from Alice or from Bob (see Sec. 4.2 of [1]). As stated at the end of Sec. 4, from the perspective of cross-correlation between the legitimate parties’ and eavesdropper’s observations of the CCRSM, the degree of eavesdropping risk in the FDSM-WE scenario can be considered highest. Note that, privacy amplification can be used in practice to eliminate such risk after carrying out the information reconciliation, and performing privacy amplification results in removal of some bits from the reconciled key bit sequence. From this point of view, the privacy amplification is somewhat tantamount to the role that the subtraction of Eq. (33) plays. To provide a view of the lowest limit of the secret-key capacity CRM, hereinafter we develop expressions for the lower bound CRM,LB in the FDSM-WE scenario, because the FDSM-WE scenario may suffer from the highest degree of eavesdropping risk. Hence, we can reveal how random modulation affects the secret-key extraction, from an information-theoretic perspective, when Eve is most capable of acquiring information about the CCRSM.

For the aforementioned FDSM-WE scenario, the mutual information of ${\phi _A}(t)$ and ${\phi _B}(t)$ can be formulated by (see Appendix B)

$${I_{RM}}({{\phi_A};{\phi_B}} )= \frac{{{\gamma _1}}}{{2\pi {{\bar{x}}_B}}}\int_{ - \infty }^\infty {\textrm{d}{\varphi _A}\int_{ - \infty }^\infty {\textrm{d}{\varphi _B}} } {Q^{\prime}_0}({{\varphi_A},{\varphi_B}} ){\log _2}\left[ {\frac{{{{Q^{\prime}}_0}({{\varphi_A},{\varphi_B}} )}}{{{{Q^{\prime}}_1}({{\varphi_A}} ){{Q^{\prime}}_2}({{\varphi_B}} )}}} \right]$$
with $\gamma_1=R^2 s_T^2 \bar{x}_B^2 \bar{h}_M^2 / \sigma_{n, 0}^2$ being the average SNR at the legitimate terminals when random modulation is carried out by Bob,
$${Q^{\prime}_0}({{\varphi_A},{\varphi_B}} )= \int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}} \frac{1}{{\hat{x}}}{p_{\hat{x}}}({\hat{x}} )\int_0^\infty {\textrm{d}\hat{h}{p_{\hat{h}}}({\hat{h}} )\exp \left[ { - \frac{{{{({{\varphi_A} - \hat{x}\hat{h}} )}^2}}}{{2/{\gamma_1}}} - \frac{{{{({{\varphi_B} - \hat{x}\hat{h}} )}^2}}}{{2\bar{x}_B^2{{\hat{x}}^2}/{\gamma_1}}}} \right]}, $$
$${Q^{\prime}_1}({{\varphi_A}} )= \int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}} {p_{\hat{x}}}({\hat{x}} )\int_0^\infty {\textrm{d}\hat{h}{p_{\hat{h}}}({\hat{h}} )\exp \left[ { - \frac{{{{({{\varphi_A} - \hat{x}\hat{h}} )}^2}}}{{2/{\gamma_1}}}} \right]}, $$
$${Q^{\prime}_2}({{\varphi_B}} )= \int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}} \frac{1}{{\hat{x}}}{p_{\hat{x}}}({\hat{x}} )\int_0^\infty {\textrm{d}\hat{h}{p_{\hat{h}}}({\hat{h}} )\exp \left[ { - \frac{{{{({{\varphi_B} - \hat{x}\hat{h}} )}^2}}}{{2\bar{x}_B^2{{\hat{x}}^2}/{\gamma_1}}}} \right]}, $$
where ${\bar{x}_B}$ is the mean of the modulation signal xB(t), and ${p_{\hat{x}}}(\hat{x})$ denotes the PDF of the normalized modulation signal. In arriving at Eqs. (34)–(37), we let $\sigma _{n,A}^2$=$\sigma _{n,B}^2$=$\sigma _{n,0}^2$. Note that the normalized modulation signal is defined as the ratio of the modulation signal xB(t) to its mean. After carrying out an examination of Eqs. (34)–(37) similar to that of Eqs. (30)–(32) presented above, one can find $\sigma _s^2$/$\sigma _{n,0}^2$≪1 does not necessarily lead to ${I_{RM}}({\phi _A};{\phi _B})$ ≈ 0, and ${p_{\hat{x}}}(\hat{x})$ is another decisive factor for ${I_{RM}}({\phi _A};{\phi _B})$. By use of an analysis approach analogous to that used to deal with the inner integration over $\hat{h}$ in Eqs. (31) and (32), one can infer that ${I_{RM}}({\phi _A};{\phi _B})$ ≈ 0 under the conditions of both $\sigma _s^2$/$\sigma _{n,0}^2$≪1 and $R^2 s_T^2 \bar{h}_M^2 \sigma_{x, B}^2 / \sigma_{n, 0}^2 \ll 1$ if ${p_{\hat{x}}}(\hat{x})$ takes on a single-peak shape. The underlying reason for this finding is elucidated as follows: The conditions $\sigma _s^2 /\sigma _{n,0}^2 \ll 1\,{\rm and}\,R^2s_T^2 \bar{h}_M^2 \sigma _{x,B}^2 /\sigma _{n,0}^2 \ll 1$ mean that random variation in the photodetectors’ output signals of legitimate parties due to noises dominates the variation of ${\phi _A}(t)$ and ${\phi _B}(t)$, hence resulting in that ${\phi _A}(t)$ and ${\phi _B}(t)$ are little correlated with hM(t) and xB(t). We can infer from the above observations that, for the FDSM-WE scenario, ${I_{RM}}({\phi _A};{\phi _B})$ with given γ1 and ${\bar{x}_B}$ becomes relatively large if random variation in the photodetectors’ output signals due to either hM(t) or xB(t) is much stronger than that due to the noises. Hence, in the FDSM-WE scenario, the randomness contained in xB(t), besides hM(t), makes positive contribution towards ${I_{RM}}({\phi _A};{\phi _B})$.

Below we proceed by developing the mathematical expressions for ${I_{RM}}({\phi _A};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$ and ${I_{RM}}({\phi _B};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$ in the FDSM-WE scenario. Note that, ${\tilde{y}_{AE}}(t)$ is indeed independent of both ${\phi _A}(t)$ and ${\phi _B}(t)$ in the FDSM-WE scenario. Therefore, there exist the identities ${I_{RM}}({\phi _A};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$=${I_{RM}}({\phi _A};{\tilde{y}_{BE}})$ and ${I_{RM}}({\phi _B};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$=${I_{RM}}({\phi _B};{\tilde{y}_{BE}})$ for the FDSM-WE scenario. By recalling that ${\tilde{y}_{BE}}(t)$ is equal to ${R_E}{s_T}{\tilde{h}_{BE}}{x_B}(t)$ in the FDSM-WE scenario with ${\tilde{h}_{BE}}$ being the deterministic transmission coefficient of the optical channel from Bob to Eve, one can further deduce that ${I_{RM}}({\phi _A};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$=${I_{RM}}({\phi _A};{x_B})$ and ${I_{RM}}({\phi _B};{\tilde{y}_{AE}},{\tilde{y}_{BE}})$=${I_{RM}}({\phi _B};{x_B})$. Following an approach similar to that used to derive Eqs. (34)–(37), one finds

$${I_{RM}}({{\phi_A};{x_B}} )= \sqrt {\frac{{{\gamma _1}}}{{2\pi }}} \cdot \int_{ - \infty }^\infty {\textrm{d}{\varphi _A}\int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}} } {p_{\hat{x}}}({\hat{x}} ){W_0}({{\varphi_A},\hat{x}} ){\log _2}\left[ {\frac{{{W_0}({{\varphi_A},\hat{x}} )}}{{{W_1}({{\varphi_A}} )}}} \right]$$
for the FDSM-WE scenario, where
$${W_0}({{\varphi_A},\hat{x}} )= \int_0^\infty {{\rm d}\hat{h}\exp [{ - {{{\gamma_1} \cdot {{({{\varphi_A} - \hat{x}\hat{h}} )}^2}} / 2}} ]{p_{\hat{h}}}({\hat{h}} )}, $$
$${W_1}({{\varphi_A}} )= \int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}{p_{\hat{x}}}({\hat{x}} )\int_0^\infty {\text{d}\hat{h}\exp [{ - {{{\gamma_1} \cdot {{({{\varphi_A} - \hat{x}\hat{h}} )}^2}} / 2}} ]{p_{\hat{h}}}({\hat{h}} )} }.$$

Similarly, one finds

$${I_{RM}}({{\phi_B};{x_B}} )= \sqrt {\frac{{{\gamma _1}}}{{2\pi }}} \cdot \frac{1}{{{{\bar{x}}_B}}}\int_{ - \infty }^\infty {\textrm{d}{\varphi _B}\int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}} } {p_{\hat{x}}}({\hat{x}} ){W^{\prime}_0}({{\varphi_B},\hat{x}} ){\log _2}\left[ {\frac{{{{W^{\prime}}_0}({{\varphi_B},\hat{x}} )}}{{{{W^{\prime}}_1}({{\varphi_B}} )}}} \right]$$
for the FDSM-WE scenario, where
$${W^{\prime}_0}({{\varphi_B},\hat{x}} )= \frac{1}{{\hat{x}}}\int_0^\infty {\text{d}\hat{h}\exp \left[ { - \frac{{{{({{\varphi_B} - \hat{x}\hat{h}} )}^2}}}{{2\bar{x}_B^2{{\hat{x}}^2}/{\gamma_1}}}} \right]{p_{\hat{h}}}({\hat{h}} )}, $$
$${W^{\prime}_1}({{\varphi_B}} )= \int_0^{1/{{\bar{x}}_B}} {\textrm{d}\hat{x}{p_{\hat{x}}}({\hat{x}} )\frac{1}{{\hat{x}}}\int_0^\infty {\text{d}\hat{h}\exp \left[ { - \frac{{{{({{\varphi_B} - \hat{x}\hat{h}} )}^2}}}{{2\bar{x}_B^2{{\hat{x}}^2}/{\gamma_1}}}} \right]{p_{\hat{h}}}({\hat{h}} )} }. $$

With use of an analysis approach similar to that applied to Eqs. (31), (32) and (35)–(37), one can deduce that ${I_{RM}}({\phi _A};{x_B})$ ≈ 0 and ${I_{RM}}({\phi _B};{x_B})$ ≈ 0 in the FDSM-WE scenario under the conditions of both $\sigma _s^2/\sigma _{n,0}^2\ll1$ and R2$s_T^2\bar{h}_M^2\sigma _{x,B}^2/\sigma _{n,0}^2\ll1$, if both ${p_{\hat{x}}}(\hat{x})$ and ${p_{\hat{h}}}(\hat{h})$ take on a single-peak shape.

To numerically evaluate Eqs. (29) and (33), we need to give the specific mathematical forms of ${p_{\hat{h}}}(\hat{h})$ and ${p_{\hat{x}}}(\hat{x})$. In [30], the inverse Gaussian gamma (IGG) distribution has been proposed to model the statistical behavior of fluctuations in the received optical signal of free-space optical communications through atmospheric turbulence. For exemplifying a quantitative information-theoretic analysis of secret-key extraction from the CCRSM, below we let ${p_{\hat{h}}}(\hat{h})$ be the IGG distribution given by [30]

$${p_{\hat{h}}}({\hat{h}} )= \sqrt {\frac{{2\alpha }}{\pi }} \frac{{{\beta ^\beta }{{\hat{h}}^{\beta - 1}}}}{{\Gamma (\beta )}}\exp (\alpha ){\left( {1 + \frac{{2\beta }}{\alpha }\hat{h}} \right)^{ - \beta /2 - 1/4}}{K_{\beta + 1/2}}\left( {\alpha \sqrt {1 + 2\beta \hat{h}/\alpha } } \right),\hat{h} > 0, $$
where Γ(·) denotes the gamma function and Kβ+1/2(·) represents a modified Bessel function of the second kind; α and β are the two control parameters of the IGG distribution. Because it is required that xB(t) ranges from 0 to 1, in what follows, we assume xB(t) obeys a standard two-sided power distribution [31] with parameters θ = 0 and n > 0; with this in mind, it follows that
$${p_{\hat{x}}}({\hat{x}} )= {\bar{x}_B}n{({1 - {{\bar{x}}_B}\hat{x}} )^{n - 1}}, $$
where ${\bar{x}_B}$ = 1/(n + 1) and $0 \le \hat{x} \le 1/{\bar{x}_B}$. For a standard two-sided power distribution [31] with parameters θ = 0 and n > 0, there exists $\sigma _{x,B}^2$= n/(n + 2).

Figure 5 exemplifies how the lower bound CRM,LB in the FDSM-WE scenario behaves in terms of $\sigma _{x,B}^2$ with different PDFs of the normalized channel transmission coefficient. In producing the curves of Fig. 5, we let γ1 ≡ 20 and used Eqs. (44) and (45) to model ${p_{\hat{h}}}(\hat{h})$ and ${p_{\hat{x}}}(\hat{x})$, respectively. We point out here that CRM,LB with $\sigma _{x,B}^2$ = 0 in the FDSM-WE scenario is actually equal to CNMIE, i.e., the secret-key capacity without random modulation. Moreover, in Fig. 5, $\sigma _{x,B}^2$= 0.24, 0.48 and 0.72 correspond to n = 0.63, 1.85 and 5.14, respectively. With examination of Eqs. (34), (38), (41), (44) and (45), we emphasize that the curves in Fig. 5 are completely determined by the parameters γ1, α, β and n. For given parameter n of the standard two-sided power distribution, both $\sigma _{x,B}^2$ and ${\bar{x}_B}$ are determinate, and there exist numerous possible parameter combinations of the average channel transmission coefficient ${\bar{h}_M}$, photodetector’s noise variance $\sigma _{n,0}^2$, photodetector’s responsibility R, and transmitted optical power sT that can result in the fixed γ1 of 20 used in producing Fig. 5. Notice that, the average channel transmission coefficient ${\bar{h}_M}$ depends on both the channel propagation distance and the divergence angle of the transmitted beam. When designing practical CCRSM-based secret-key extraction systems, to obtain fixed γ1, one has several degrees of freedom for making trade-offs between the channel propagation distance, divergence angle of the transmitted beam, photodetector’s noise variance, photodetector’s responsibility, and transmitted optical power. It is seen from Fig. 5 that, with the same average SNR γ1, the lower bound CRM,LB in the FDSM-WE scenario grows slightly with increasing $\sigma _{x,B}^2$. Notice that, when $\sigma _{x,B}^2$ = 0, CRM,LB can be fully attributed to turbulence-induced optical fluctuations. With the above observations in mind, one can infer that the random modulation does not harm the usefulness of the common-randomness component of the CCRSM due to turbulence-induced optical fluctuations; in other words, although the random modulation causes a proportion of information of the CCRSM to be leaked to Eve, we can consider that the leaked information only stems from the randomly time-varying modulation signal, and the proportion of information arising from the said common-randomness component due to turbulence-induced optical fluctuations is still secure against Eve’s eavesdropping. We emphasize once again that the random modulation is employed in this work only for decorrelating the legitimate parties’ consecutive observations of the CCRSM. From this point of view, by recalling the analysis in Sec. 3, we find that the random modulation can enable us to obtain nearly uncorrelated consecutive observations of both ${\phi _A}(t)$ and ${\phi _B}(t)$ with use of a sampling interval rather smaller than the correlation time of turbulence-induced optical fluctuations, while ensure that the potential of extracting secret keys from the randomness stemming from turbulence-induced optical fluctuations is not weakened. Thus, the random modulation can be utilized to increase the number of extracted uncorrelated secret key bits per second. Moreover, it is found from Fig. 5 that, with other parameters fixed, CRM,LB associated with α = 8.54 and β = 6.11 is different from that related to α = 22.1 and β = 19.1. The underlying physical reason for this is that there is much difference in the shape of ${p_{\hat{h}}}(\hat{h})$ between the two groups of α and β.

 figure: Fig. 5.

Fig. 5. CRM,LB in terms of $\sigma _{x,B}^2$ in the FDSM-WE scenario with γ1 ≡ 20 and different PDFs of the normalized channel transmission coefficient. The modulation signal xB(t) obeys a standard two-sided power distribution [31] with parameters θ = 0 and n > 0.

Download Full Size | PDF

At this point, we make a simple comparison of our approach with the free-space optical secret key agreement (FSO-SKA) reported in [9]. First, a fundamental difference between the FSO-SKA and our approach lies in the physical randomness source used to generate the initial key bits. Specifically, in the FSO-SKA, a physical random number generator at the sender terminal is used to generate the initial key bits, whereas, in our approach, the channel’s inherent randomness is utilized to generate the initial key bits. Second, compared with the FSO-SKA, our approach does not require the use of a probing station to act as a virtual eavesdropper. Indeed, in our approach, the decorrelation of the legitimate parties’ and eavesdropper’s observations of the fluctuating channel transmission coefficient is easily guaranteed by the fact that the eavesdropper has to stay at a position separated from the legitimate parties by a distance much larger than the spatial coherence length of the atmosphere, which is on the order of several centimeters, for keeping the legitimate parties unaware of the eavesdropping action. Similar to the FSO-SKA, our approach needs employing privacy amplification to eliminate the leaked information to Eve; however, we can estimate the leaked information in the possible worst case by using Eqs. (38) and (41), where both ${p_{\hat{x}}}(\hat{x})$ and ${p_{\hat{h}}}(\hat{h})$ can be obtained by the legitimate parties without the help of a virtual eavesdropper; although Eqs. (38) and (41) may overestimate the actual leaked information, the potential of extracting secret keys from the CCRSM’s constituent randomness component due to turbulence-induced optical fluctuations is indeed not harmed and use of such overestimated leaked information in privacy amplification to compress and hence shorten the reconciled bit sequence does ensure that the finial key bits are unknown to Eve. Below, we illustrate what key generation rate can be achieved under the condition of given γ1 when a fixed temporal sampling interval τs of 2.5 × 10−5 ms is used by legitimate parties to produce measurement samples of the CCRSM. Notice that, τs = 2.5 × 10−5 ms means that there are 4 × 107 measurement samples produced per second. According to Fig. 5, one finds that CRM,LB = 0.677 bits / channel measurement when $\sigma _{x,B}^2$ = 0.72 (equivalently, n = 5.14), α = 22.1 and β = 19.1. Hence, theoretically, we can attain a key generation rate of at least 27.08 Mbps when τs = 2.5 × 10−5 ms, γ1 = 20, n = 5.14, α = 22.1 and β = 19.1. This key generation rate is indeed comparable to that of the FSO-SKA.

We note that secret-key generation with artificial randomness has been suggested for RF wireless channels [20,21]. Our work differs from [20] in that random modulation in the CCRSM is carried out by the two communicating parties instead of a helper, and artificial interference produced by the helper in [20] causes an additive distortion rather than the multiplicative distortion induced by the random modulation. The artificial randomness in [21] is produced by use of random unitary precoding matrix, which does not alter the statistical behavior of the MIMO RF wireless channel therein and hence the relevant secret-key capacity; on the other hand, the modulation signal in the CCRSM is only required to vary randomly within the range of [0, 1] and its normalized variance does play a role in determining the lower bound of the secret-key capacity, which can be easily verified by examining Fig. 5. As a final comment, it is easy to find that CRM depends on ${p_{\hat{x}}}(\hat{x})$, i.e., the PDF of the normalized modulation signal. In the exemplary calculations related to Fig. 5, we have considered that the modulation signal obeys a standard two-sided power distribution. Evidently, other distributions with a support ranging from 0 to 1 can also be chosen in practice. An interesting question arises as to how to find the optimum ${p_{\hat{x}}}(\hat{x})$ that can maximize the secret-key capacity. Nevertheless, treatment of this question is beyond the scope of this work. It is a theoretical issue that deserves future attention.

6. Conclusions

By targeting at extracting shared secret keys from single-mode atmospheric optical channels equipped with random modulation, we have first proposed a novel randomness sharing scheme and developed the mathematical models for the relevant common randomness source, which is called CCRSM; Our randomness sharing scheme can be used to implement enhanced secret-key extraction, which can break the restriction imposed by the correlation time of turbulence-induced optical fluctuations on the number of extracted uncorrelated key bits per second. Subsequently, theoretical expressions for both the autocorrelation of the legitimate parties’ observations of the CCRSM and the cross-correlation between the two legitimate parties’ observations of the CCRSM have been further derived, and thereby the random-modulation-caused decorrelation of the legitimate parties’ consecutive observations of the CCRSM has been examined with use of numerical calculations. The cross-correlation coefficient of the eavesdropper’s and legitimate parties’ observations of the CCRSM is theoretically formulated for facilitating the analysis of possible eavesdropping risk, and the dependence of such cross-correlation coefficient on the normalized variance of the time-varying modulation signal is numerically analyzed for the FDSM-WE, PDSM-WE and DM-WE scenarios. Finally, the lower bound of the secret-key capacity in consideration of our randomness sharing scheme is dealt with for revealing how random modulation impacts the key extraction from an information-theoretic perspective. It has been shown that random modulation can be used to effectively decorrelate the legitimate parties’ consecutive observations of the CCRSM and simultaneously does not harm the potential of extracting secret keys from the CCRSM’s randomness component due to turbulence-induced optical fluctuations. From our theoretical and numerical results, it is found that, with use of the CCRSM, random modulation is a viable approach to breaking the restriction imposed by the correlation time of turbulence-induced optical fluctuations on the number of extracted uncorrelated key bits per second.

Appendix A

According to the definition of mutual information [32], one finds

$${I_{NMIE}}({{\phi_A};{\phi_B}} )= \int_{ - \infty }^\infty {\int_{ - \infty }^\infty {\textrm{d}{\phi _A}\textrm{d}{\phi _B}{p_{{\phi _A}{\phi _B}}}({{\phi_A},{\phi_B}} ){{\log }_2}\left[ {\frac{{{p_{{\phi_A}{\phi_B}}}({{\phi_A},{\phi_B}} )}}{{{p_{{\phi_A}}}({{\phi_A}} ){p_{{\phi_B}}}({{\phi_B}} )}}} \right]} } , $$
where ${p_{{\phi _A}{\phi _B}}}({\phi _A},{\phi _B})$ is the joint PDF of ${\phi _A}(t)$ and ${\phi _B}(t)$, ${p_{{\phi _A}}}({\phi _A})$ is the PDF of ${\phi _A}(t)$, and ${p_{{\phi _B}}}({\phi _B})$ is the PDF of ${\phi _B}(t)$. Notice that, nA(t) and nB(t) are AWGNs; they have a Gaussian distribution with zero mean and variance $\sigma _{n,0}^2$. Note that nA(t) =${\phi _A}(t)$RsThM(t) and nB(t) =${\phi _B}(t)$RsThM(t) when xA(t) = xB(t) ≡ 1. With the Gaussian PDFs for nA(t) and nB(t) in hand, by making the change of variables nA(t) =${\phi _A}(t)$RsThM(t) and nB(t) =${\phi _B}(t)$RsThM(t), the conditional PDF for ${\phi _X}(t)$ given hM(t) in the case of xA(t) = xB(t) ≡ 1 can be given by
$${p_{{\phi _X}|{h_M}}}({{\phi_X}|{h_M}} )= \frac{1}{{\sqrt {2\pi \sigma _{n,0}^2} }}\exp \left[ { - \frac{{{{({{\phi_X} - R{s_T}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}}} \right], $$
where X = “A” or “B”. Because nA(t) and nB(t) are statistically independent, their joint PDF factorizes into a product of the PDFs for nA(t) and nB(t). With the joint Gaussian PDF of nA(t) and nB(t) in hand, by making change of variables nA(t) =${\phi _A}(t)$RsThM(t) and nB(t) =${\phi _B}(t)$RsThM(t), the conditional joint PDF of ${\phi _A}(t)$ and ${\phi _B}(t)$ given hM(t) can be written by
$${p_{{\phi _A}{\phi _B}|{h_M}}}({{\phi_A},{\phi_B}|{h_M}} )= \frac{1}{{2\pi \sigma _{n,0}^2}}\exp \left[ { - \frac{{{{({{\phi_A} - R{s_T}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}} - \frac{{{{({{\phi_B} - R{s_T}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}}} \right]$$
in the case of xA(t) = xB(t) ≡ 1. With use of the law of total probability [33], the PDF of ${\phi _X}(t)$ can be written by
$${p_{{\phi _X}}}({{\phi_X}} )= \int_0^\infty {{p_{{\phi _X}|{h_M}}}({{\phi_X}|{h_M}} ){p_h}({{h_M}} )} \textrm{d}{h_M}, $$
where ph(hM) is the PDF of hM(t). Similarly, the joint PDF of ${\phi _A}(t)$ and ${\phi _B}(t)$ can be expressed by
$${p_{{\phi _A}{\phi _B}}}({{\phi_A},{\phi_B}} )= \int_0^\infty {{p_{{\phi _A}{\phi _B}|{h_M}}}({{\phi_A},{\phi_B}|{h_M}} ){p_h}({{h_M}} )\textrm{d}{h_M}} . $$

First introducing Eqs. (A4) and (A5) into Eq. (A1) and then letting φA=${\phi _A}$/(RsT${\bar{h}_M}$), φB=${\phi _B}$/(RsT${\bar{h}_M}$) and $\hat{h} = h/{\bar{h}_M}$, with ${\bar{h}_M}$ being the mean of hM(t), lead us to Eqs. (30)–(32).

Appendix B

When xA(t) ≡ 1, we find that nA(t) =${\phi _A}(t)$RsTxB(t)hM(t) and nB(t) =${\phi _B}(t)/{x_B}(t)$RsThM(t). Similar to the treatment in Appendix A, one can find the following conditional PDFs:

$${p_{{\phi _A}|{h_M}{x_B}}}({{\phi_A}|{h_M},{x_B}} )= \frac{1}{{\sqrt {2\pi \sigma _{n,0}^2} }}\exp \left[ { - \frac{{{{({{\phi_A} - R{s_T}{x_B}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}}} \right], $$
$${p_{{\phi _B}|{h_M}{x_B}}}({{\phi_B}|{h_M},{x_B}} )= \frac{1}{{\sqrt {2\pi \sigma _{n,0}^2} }}\frac{1}{{{x_B}}}\exp \left[ { - \frac{{{{({{\phi_B}/{x_B} - R{s_T}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}}} \right], $$
$${p_{{\phi _A}{\phi _B}|{h_M}{x_B}}}({{\phi_A},{\phi_B}|{h_M},{x_B}} )= \frac{1}{{2\pi \sigma _{n,0}^2}}\frac{1}{{{x_B}}}\exp \left[ { - \frac{{{{({{\phi_A} - R{s_T}{x_B}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}} - \frac{{{{({{\phi_B}/{x_B} - R{s_T}{h_M}} )}^2}}}{{2\sigma_{n,0}^2}}} \right]$$
in the case of xA(t) ≡ 1. We point out that Eqs. (B1)–(B3) represent the conditional PDF of ${\phi _A}(t)$ given hM(t) and xB(t), the conditional PDF of ${\phi _B}(t)$ given hM(t) and xB(t), and the conditional joint PDF of ${\phi _A}(t)$ and ${\phi _B}(t)$ given hM(t) and xB(t), respectively. By use of the law of total probability [33], one finds
$${p_{{\phi _X}}}({{\phi_X}} )= \int_0^1 {\textrm{d}{x_B}{p_x}({{x_B}} )\int_0^\infty {\textrm{d}{h_M}{p_{{\phi _X}|{h_M}{x_B}}}({{\phi_X}|{h_M},{x_B}} ){p_h}({{h_M}} )} } , $$
$${p_{{\phi _A}{\phi _B}}}({{\phi_A},{\phi_B}} )= \int_0^1 {\textrm{d}{x_B}{p_x}({{x_B}} )\int_0^\infty {\textrm{d}{h_M}{p_{{\phi _A}{\phi _B}|{h_M}{x_B}}}({{\phi_A},{\phi_B}|{h_M},{x_B}} ){p_h}({{h_M}} )} } , $$
where X = “A” or “B”, and px(xB) is the PDF of xB(t). In analogy with the derivation of Appendix A, we can arrive at Eqs. (34)–(37) by letting φA=${\phi _A}$/(RsT${\bar{x}_B}{\bar{h}_M}$), φB=${\phi _B}$/(RsT${\bar{x}_B}{\bar{h}_M}$), $\hat{h} = h/{\bar{h}_M}$, and $\hat{x} = {x_B}/{\bar{x}_B}$ with ${\bar{x}_B}$ being the mean of xB(t).

Funding

National Natural Science Foundation of China (61775022, 62275033); Natural Science Foundation of Chongqing (cstc2021jcyj-msxmX0457).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not available at this time but may be obtained from the authors upon reasonable request.

References

1. M. Bloch and J. Barros, Physical-Layer Security: From Information Theory to Security Engineering (Cambridge University Press, 2011).

2. Y. Liang, H. V. Poor, and S. Shamai, “Information theoretic security,” Found. Trends Commun. Info. Theory 5(4-5), 355–580 (2007). [CrossRef]  

3. U. M. Maurer, “Secret key agreement by public discussion from common information,” IEEE Trans. Inf. Theory 39(3), 733–742 (1993). [CrossRef]  

4. R. Ahlswede and I. Csiszàr, “Common randomness in information theory and cryptography-part I: Secret sharing,” IEEE Trans. Inf. Theory 39(4), 1121–1132 (1993). [CrossRef]  

5. H. Endo, T. Sasaki, M. Takeoka, M. Fujiwara, M. Koashi, and Masahide Sasaki, “Line-of-sight quantum key distribution with differential phase,” New J. Phys. 24(2), 025008 (2022). [CrossRef]  

6. E. Otte, I. Nape, C. Rosales-Guzman, C. Denz, A. Forbes, and B. Ndagano, “High-dimensional cryptography with spatial modes of light: tutorial,” J. Opt. Soc. Am. B 37(11), A309–A323 (2020). [CrossRef]  

7. Z. Pan and I. B. Djordjevic, “Secret key distillation over satellite-to-satellite free-space optics channel with a limited-sized aperture eavesdropper in the same plane of the legitimate receiver,” Opt. Express 28(25), 37129–37148 (2020). [CrossRef]  

8. T. Ikuta and I. Kyo, “Intensity modulation and direct detection quantum key distribution based on quantum noise,” New J. Phys. 18(1), 013018 (2016). [CrossRef]  

9. H. Endo, M. Fujiwara, M. Kitamura, O. Tsuzuki, T. Ito, R. Shimizu, M. Takeoka, and M. Sasaki, “Free space optical secret key agreement,” Opt. Express 26(18), 23305–23332 (2018). [CrossRef]  

10. T. A. Eriksson, P. V. Trinh, H. Endo, M. Takeoka, and M. Sasaki, “Secret key rates for intensity-modulated dual-threshold detection key distribution under individual beam splitting attacks,” Opt. Express 26(16), 20409–20419 (2018). [CrossRef]  

11. N. Yamamori and K. Inoue, “Experimental demonstration of intensity-modulation/direct-detection secret key distribution,” Jpn. J. Appl. Phys. 59(2), 022003 (2020). [CrossRef]  

12. K. Banaszek, M. Jachura, P. Kolenderski, and M. Lasota, “Optimization of intensity-modulation/direct-detection optical key distribution under passive eavesdropping,” Opt. Express 29(26), 43091–43103 (2021). [CrossRef]  

13. H. Endo, M. Fujiwara, M. Kitamura, O. Tsuzuki, R. Shimizu, M. Takeoka, and M. Sasaki, “Group key agreement over free-space optical links,” OSA Continuum 3(9), 2525–2543 (2020). [CrossRef]  

14. N. Bornman, A. Forbes, and A. Kempf, “Random number generation and distribution out of thin (or thick) air,” J. Opt. 22(7), 075705 (2020). [CrossRef]  

15. M. D. Drake, C. F. Bas, D. R. Gervais, P. F. Renda, D. Townsend, J. J. Rushanan, J. Francoeur, N. C. Donnangelo, and M. D. Stenner, “Optical key distribution system using atmospheric turbulence as the randomness generating function: Classical optical protocol for information assurance,” Opt. Eng. 52(5), 055008 (2013). [CrossRef]  

16. I. B. Djordjevic, “Atmospheric turbulence-controlled cryptosystems,” IEEE Photonics J. 13(1), 1–9 (2021). [CrossRef]  

17. D. Wang, P. Ren, J. Cheng, Q. Du, Y. Wang, and L. Sun, “Secure transmission for mixed FSO-RF relay networks with physical-layer key encryption and wiretap coding,” Opt. Express 25(9), 10078–10089 (2017). [CrossRef]  

18. N. Wang, X. Song, J. Cheng, and V. C. Leung, “Enhancing the security of free-space optical communications with secret sharing and key agreement,” J. Opt. Commun. Netw. 6(12), 1072–1081 (2014). [CrossRef]  

19. S. N. Premnath, S. Jana, J. Croft, P. L Gowda, M. Clark, S. K. Kasera, N. Patwari, and S. V. Krishnamurthy, “Secret key extraction from wireless signal strength in real environments,” IEEE Trans. on Mobile Comput. 12(5), 917–930 (2013). [CrossRef]  

20. D. Chen, Z. Qin, X. Mao, P. Yang, Z. Qin, and R. Wang, “SmokeGrenade: An efficient key generation protocol with artificial interference,” IEEE Trans. Inform. Forensic Secur. 8(11), 1731–1745 (2013). [CrossRef]  

21. S. Baksi and D. C. Popescu, “Secret key generation with precoding and role reversal in MIMO wireless systems,” IEEE Trans. Wireless Commun. 18(6), 3104–3112 (2019). [CrossRef]  

22. B. T. Quist and M. A. Jensen, “Maximization of the channel-based key establishment rate in MIMO systems,” IEEE Trans. Wireless Commun. 14(10), 5565–5573 (2015). [CrossRef]  

23. M. F. Haroun and T. A. Gulliver, “Secret key generation using chaotic signals over frequency selective fadingchannels,” IEEE Trans. Inform. Forensic Secur. 10(8), 1764–1775 (2015). [CrossRef]  

24. C. Chen and M. A. Jensen, “Secret key establishment using temporally and spatially correlated wireless channel coefficients,” IEEE Trans. on Mobile Comput. 10(2), 205–215 (2011). [CrossRef]  

25. G. Li, A. Hu, J. Zhang, L. Peng, C. Sun, and D. Cao, “High-agreement uncorrelated secret key generation based on principal component analysis preprocessing,” IEEE Trans. Commun. 66(7), 3022–3034 (2018). [CrossRef]  

26. J. H. Shapiro and A. L. Puryear, “Reciprocity-enhanced optical communication through atmospheric turbulence-Part I: Reciprocity proofs and far-field power transfer optimization,” J. Opt. Commun. Netw. 4(12), 947–954 (2012). [CrossRef]  

27. R. R. Parenti, J. M. Roth, J. H. Shapiro, F. G. Walther, and J. A. Greco, “Experimental observations of channel reciprocity in single-mode free-space optical links,” Opt. Express 20(19), 21635–21644 (2012). [CrossRef]  

28. L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, 2nd ed. (SPIE, 2005).

29. A. Jurado-Navas, A. Garca-Zambrana, and A. Puerta-Notario, “Efficient lognormal channel model for turbulent FSO communications,” Electron. Lett. 43(3), 178–179 (2007). [CrossRef]  

30. M. Cheng, Y. Guo, J. Li, X. Zheng, and L. Guo, “Inverse Gaussian gamma distribution model for turbulence-induced fading in free-space optical communication,” Appl. Opt. 57(12), 3031–3037 (2018). [CrossRef]  

31. S. Kotz and J. R. van Dorp, Beyond Beta: Other Continuous Families of Distributions with Bounded Support and Applications (World Scientific Publishing Company, 2004).

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

33. M. H. DeGroot and M. J. Schervish, Probability and Statistics, 4th ed. (Pearson Education, 2012).

Data availability

Data underlying the results presented in this paper are not available at this time but may be obtained from the authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. Schematic diagram of a reciprocal single-mode bidirectional atmospheric optical channel equipped with random modulation.
Fig. 2.
Fig. 2. Correlation coefficient of two consecutive observations of ${\phi _A}(t)$, i.e., ${b_{\phi ,A}}({\tau _s})$, in terms of $\sigma _{x,A}^2$ with γA ≡ ∞ and different $\sigma _h^2$; (a) the single-modulation case with $\sigma _{x,A}^2$ > 0 and $\sigma _{x,B}^2$ ≡ 0; (b) the double-modulation case with $\sigma _{x,A}^2$$\sigma _{x,B}^2$> 0. The correlation time τh,0 and the temporal sampling interval τs are specified as 1 ms and 2.5×10−5 ms, respectively; the correlation time of the random modulation signal is small enough that bx,A(τ)$\sigma _{x,A}^2$≪ 1 and bx,B(τ)$\sigma _{x,B}^2$≪ 1 hold true at a value of τ smaller than τs.
Fig. 3.
Fig. 3. Cross-correlation coefficients of ${\tilde{y}_{BE}}(t)$ and ${\phi _X}(t)$ and of ${\tilde{y}_{AE}}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,B}^2$ and $\sigma _{x,A}^2$ in the FDSM-WE and PDSM-WE scenarios, respectively, where the subscript X = “A” or “B”; ${\gamma _A}$=${\gamma _B}$=20, ${\gamma _{AE}}$= ∞, and ${\gamma _{BE}}$= ∞; (a) and (b) correspond to the FDSM-WE scenario where ${\tilde{y}_{BE}}(t)$=${R_E}{s_T}{\tilde{h}_{BE}}{x_B}(t)$ and ${\tilde{y}_{AE}}(t)$ is independent of ${\phi _X}(t)$; (c) and (d) correspond to the PDSM-WE scenario where ${\tilde{y}_{AE}}(t)$ is partially correlated with ${\phi _X}(t)$ and ${\tilde{y}_{BE}}(t)$ is independent of ${\phi _X}(t)$. (a) $\sigma _h^2 = 0.3$, $\sigma _{h,BE}^2$=0, and $\sigma _{x,A}^2$=0; (b) $\sigma _h^2 = 0.1$, $\sigma _{h,BE}^2$= 0, and $\sigma _{x,A}^2$= 0; (c) $\sigma _h^2 = 0.3$, $\sigma _{h,AE}^2$=$\sigma _h^2$, and $\sigma _{x,B}^2$=0; (d) $\sigma _h^2 = 0.1$, $\sigma _{h,AE}^2$=$\sigma _h^2$, and $\sigma _{x,B}^2$= 0.
Fig. 4.
Fig. 4. Cross-correlation coefficients of ${\tilde{y}_{XE}}(t)$ and ${\phi _A}(t)$, of ${\tilde{y}_{XE}}(t)$ and ${\phi _B}(t)$, and of ${\phi _E}(t)$ and ${\phi _X}(t)$ as a function of $\sigma _{x,B}^2$ ($\sigma _{x,A}^2$), with the subscript X = “A” or “B”. γA = γB = 20, γAE = γBE = ∞, $\sigma _{h,BE}^2$= 0, and $\sigma _{h,AE}^2$ = $\sigma _h^2$. (a) $\sigma _h^2$ = 0.3; (b) $\sigma _h^2$ = 0.1. The worst eavesdropping scenario is considered in producing the curves. Both Alice and Bob perform the random modulation independently with $\sigma _{x,B}^2$ = $\sigma _{x,A}^2$.
Fig. 5.
Fig. 5. CRM,LB in terms of $\sigma _{x,B}^2$ in the FDSM-WE scenario with γ1 ≡ 20 and different PDFs of the normalized channel transmission coefficient. The modulation signal xB(t) obeys a standard two-sided power distribution [31] with parameters θ = 0 and n > 0.

Equations (55)

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

y A ( t ) = s T x B ( t ) h B A ( t ) ,
y B ( t ) = s T x A ( t ) h A B ( t ) ,
y ~ X ( t ) = R y X ( t ) + n X ( t ) ,
ϕ X ( t ) = x X ( t ) y ~ X ( t ) ,
b ϕ , A ( τ ) = [ b x , A ( τ ) σ x , A 2 + 1 ] [ b x , B ( τ ) σ x , B 2 + 1 ] [ b h ( τ ) σ h 2 + 1 ] 1 + b d ( τ ) γ A 1 ( σ x , A 2 + 1 ) ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) ( σ h 2 + 1 ) 1 + γ A 1 ( σ x , A 2 + 1 ) ,
σ x , A 2 = x A 2 ( t ) / x A ( t ) 2 1 ,
σ x , B 2 = x B 2 ( t ) / x B ( t ) 2 1
σ h 2 = h M 2 ( t ) / h M ( t ) 2 1 ,
γ A = R 2 s T 2 h M ( t ) 2 x B ( t ) 2 / σ n , A 2 ,
b x , A ( τ ) = x A ( t ) x A ( t + τ ) x A ( t ) 2 x A 2 ( t ) x A ( t ) 2 ,
b x , B ( τ ) = x B ( t ) x B ( t + τ ) x B ( t ) 2 x B 2 ( t ) x B ( t ) 2 ,
b h ( τ ) = h M ( t ) h M ( t + τ ) h M ( t ) 2 h M 2 ( t ) h M ( t ) 2 ,
b ϕ , B ( τ ) = [ b x , A ( τ ) σ x , A 2 + 1 ] [ b x , B ( τ ) σ x , B 2 + 1 ] [ b h ( τ ) σ h 2 + 1 ] 1 + b d ( τ ) γ B 1 ( σ x , B 2 + 1 ) ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) ( σ h 2 + 1 ) 1 + γ B 1 ( σ x , B 2 + 1 )
γ B = R 2 s T 2 h M ( t ) 2 x A ( t ) 2 / σ n , B 2 ,
b ϕ , A ( τ ) b h ( τ ) ζ ( γ A , σ x , A 2 , σ x , B 2 )
b ϕ , B ( τ ) b h ( τ ) ζ ( γ B , σ x , B 2 , σ x , A 2 )
ζ ( γ , σ 1 2 , σ 2 2 ) = ( σ 1 2 + 1 ) ( σ 2 2 + 1 ) + [ ( σ 1 2 + 1 ) ( σ 2 2 + 1 ) + γ 1 ( σ 1 2 + 1 ) 1 ] σ h 2 .
ρ A B = [ ϕ A ( t ) ϕ A ( t ) ] [ ϕ B ( t ) ϕ B ( t ) ] [ ϕ A ( t ) ϕ A ( t ) ] 2 [ ϕ B ( t ) ϕ B ( t ) ] 2 .
ρ A B = 1 1 + γ A 1 [ ( σ x , B 2 + 1 ) ( σ h 2 + 1 ) 1 / ( σ x , A 2 + 1 ) ] 1 × 1 1 + γ B 1 [ ( σ x , A 2 + 1 ) ( σ h 2 + 1 ) 1 / ( σ x , B 2 + 1 ) ] 1 .
y ~ X E ( t ) = R E s T x X ( t ) h X E ( t ) + n X E ( t ) ,
ρ X E , A = σ x , X 2 / χ A
χ A = ( σ x , X 2 + 1 ) ( σ h , X E 2 + 1 ) 1 + γ X E 1 ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) ( σ h 2 + 1 ) 1 + ( σ x , A 2 + 1 ) γ A 1 ,
σ h , X E 2 = h X E 2 ( t ) / h X E ( t ) 2 1 ,
γ X E = R E 2 s T 2 h X E ( t ) 2 x X ( t ) 2 / σ n , X E 2 ,
ρ X E , B = σ x , X 2 / χ B
χ B = ( σ x , X 2 + 1 ) ( σ h , X E 2 + 1 ) 1 + γ X E 1 ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) ( σ h 2 + 1 ) 1 + ( σ x , B 2 + 1 ) γ B 1 .
ρ C E , X = [ ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) 1 ] / χ C E , X ,
χ C E , X = [ ( σ x , B 2 + 1 ) ( σ x , A 2 + 1 ) ( σ h 2 + 1 ) 1 + ( σ x , X 2 + 1 ) γ X 1 ] 1 / 2 × [ ( σ x , A 2 + 1 ) ( σ x , B 2 + 1 ) ( σ h , A E 2 + 1 ) 1 ] 1 / 2
C N M I E = I N M I E ( ϕ A ; ϕ B ) ,
I N M I E ( ϕ A ; ϕ B ) = γ 0 2 π d φ A d φ B { Q 0 ( φ A , φ B ) log 2 [ Q 0 ( φ A , φ B ) Q 1 ( φ A ) Q 1 ( φ B ) ] } ,
Q 0 ( φ A , φ B ) = 0 exp [ ( φ A h ^ ) 2 + ( φ B h ^ ) 2 2 / γ 0 ] p h ^ ( h ^ ) d h ^ ,
Q 1 ( φ X ) = 0 exp [ ( φ X h ^ ) 2 2 / γ 0 ] p h ^ ( h ^ ) d h ^ ,
C R M C R M , L B = I R M ( ϕ A ; ϕ B ) min ( I R M ( ϕ A ; y ~ A E , y ~ B E ) , I R M ( ϕ B ; y ~ A E , y ~ B E ) ) ,
I R M ( ϕ A ; ϕ B ) = γ 1 2 π x ¯ B d φ A d φ B Q 0 ( φ A , φ B ) log 2 [ Q 0 ( φ A , φ B ) Q 1 ( φ A ) Q 2 ( φ B ) ]
Q 0 ( φ A , φ B ) = 0 1 / x ¯ B d x ^ 1 x ^ p x ^ ( x ^ ) 0 d h ^ p h ^ ( h ^ ) exp [ ( φ A x ^ h ^ ) 2 2 / γ 1 ( φ B x ^ h ^ ) 2 2 x ¯ B 2 x ^ 2 / γ 1 ] ,
Q 1 ( φ A ) = 0 1 / x ¯ B d x ^ p x ^ ( x ^ ) 0 d h ^ p h ^ ( h ^ ) exp [ ( φ A x ^ h ^ ) 2 2 / γ 1 ] ,
Q 2 ( φ B ) = 0 1 / x ¯ B d x ^ 1 x ^ p x ^ ( x ^ ) 0 d h ^ p h ^ ( h ^ ) exp [ ( φ B x ^ h ^ ) 2 2 x ¯ B 2 x ^ 2 / γ 1 ] ,
I R M ( ϕ A ; x B ) = γ 1 2 π d φ A 0 1 / x ¯ B d x ^ p x ^ ( x ^ ) W 0 ( φ A , x ^ ) log 2 [ W 0 ( φ A , x ^ ) W 1 ( φ A ) ]
W 0 ( φ A , x ^ ) = 0 d h ^ exp [ γ 1 ( φ A x ^ h ^ ) 2 / 2 ] p h ^ ( h ^ ) ,
W 1 ( φ A ) = 0 1 / x ¯ B d x ^ p x ^ ( x ^ ) 0 d h ^ exp [ γ 1 ( φ A x ^ h ^ ) 2 / 2 ] p h ^ ( h ^ ) .
I R M ( ϕ B ; x B ) = γ 1 2 π 1 x ¯ B d φ B 0 1 / x ¯ B d x ^ p x ^ ( x ^ ) W 0 ( φ B , x ^ ) log 2 [ W 0 ( φ B , x ^ ) W 1 ( φ B ) ]
W 0 ( φ B , x ^ ) = 1 x ^ 0 d h ^ exp [ ( φ B x ^ h ^ ) 2 2 x ¯ B 2 x ^ 2 / γ 1 ] p h ^ ( h ^ ) ,
W 1 ( φ B ) = 0 1 / x ¯ B d x ^ p x ^ ( x ^ ) 1 x ^ 0 d h ^ exp [ ( φ B x ^ h ^ ) 2 2 x ¯ B 2 x ^ 2 / γ 1 ] p h ^ ( h ^ ) .
p h ^ ( h ^ ) = 2 α π β β h ^ β 1 Γ ( β ) exp ( α ) ( 1 + 2 β α h ^ ) β / 2 1 / 4 K β + 1 / 2 ( α 1 + 2 β h ^ / α ) , h ^ > 0 ,
p x ^ ( x ^ ) = x ¯ B n ( 1 x ¯ B x ^ ) n 1 ,
I N M I E ( ϕ A ; ϕ B ) = d ϕ A d ϕ B p ϕ A ϕ B ( ϕ A , ϕ B ) log 2 [ p ϕ A ϕ B ( ϕ A , ϕ B ) p ϕ A ( ϕ A ) p ϕ B ( ϕ B ) ] ,
p ϕ X | h M ( ϕ X | h M ) = 1 2 π σ n , 0 2 exp [ ( ϕ X R s T h M ) 2 2 σ n , 0 2 ] ,
p ϕ A ϕ B | h M ( ϕ A , ϕ B | h M ) = 1 2 π σ n , 0 2 exp [ ( ϕ A R s T h M ) 2 2 σ n , 0 2 ( ϕ B R s T h M ) 2 2 σ n , 0 2 ]
p ϕ X ( ϕ X ) = 0 p ϕ X | h M ( ϕ X | h M ) p h ( h M ) d h M ,
p ϕ A ϕ B ( ϕ A , ϕ B ) = 0 p ϕ A ϕ B | h M ( ϕ A , ϕ B | h M ) p h ( h M ) d h M .
p ϕ A | h M x B ( ϕ A | h M , x B ) = 1 2 π σ n , 0 2 exp [ ( ϕ A R s T x B h M ) 2 2 σ n , 0 2 ] ,
p ϕ B | h M x B ( ϕ B | h M , x B ) = 1 2 π σ n , 0 2 1 x B exp [ ( ϕ B / x B R s T h M ) 2 2 σ n , 0 2 ] ,
p ϕ A ϕ B | h M x B ( ϕ A , ϕ B | h M , x B ) = 1 2 π σ n , 0 2 1 x B exp [ ( ϕ A R s T x B h M ) 2 2 σ n , 0 2 ( ϕ B / x B R s T h M ) 2 2 σ n , 0 2 ]
p ϕ X ( ϕ X ) = 0 1 d x B p x ( x B ) 0 d h M p ϕ X | h M x B ( ϕ X | h M , x B ) p h ( h M ) ,
p ϕ A ϕ B ( ϕ A , ϕ B ) = 0 1 d x B p x ( x B ) 0 d h M p ϕ A ϕ B | h M x B ( ϕ A , ϕ B | h M , x B ) p h ( h M ) ,
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.