Classical Capacity of the Free-Space Quantum-Optical Channel by Saikat Guha Bachelor of Technology in Electrical Engineering, Indian Institute of Technology Kanpur (2002) Submitted to the Department of Electrical Engineering and Computer Science in partial fulfillment of the requirements for the degree of Master of Science in Electrical Engineering and Computer Science at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY January 2004 @ Massachusetts Institute of Technology 2004. All rights reserved. ................. A u thor ................. Department of Electrical Engineering and Computer Science ( January 30, 2004 Certified by..... jH. Shapiro 'effr Julius A. / trgtonjiofessor of Electrical Engineering Thesis Supervisor Accepted by ...... Arthur C. Smith Chairman, Department Committee on Graduate Students MASSACHUSETTS INSTITUTE OF TECHNOLOGY APR 15 2004 LIBRARIES BARKER Classical Capacity of the Free-Space Quantum-Optical Channel by Saikat Guha Submitted to the Department of Electrical Engineering and Computer Science on January 30, 2004, in partial fulfillment of the requirements for the degree of Master of Science in Electrical Engineering and Computer Science Abstract Exploring the limits to reliable communication rates over quantum channels has been the primary focus of many researchers over the past few decades. In the present work, the classical information carrying capacity of the free-space quantum optical channel has been studied thoroughly in both the far-field and near-field propagation regimes. Results have been obtained for the optimal capacity, in which information rate is maximized over both transmitter encodings and detection schemes at the receiver, for the entanglement-assisted capacity, and also for sub-optimal systems that employ specific transmitter and receiver structures. For the above cases, several new broadband results have been obtained for capacity in the presence of both diffraction limited loss and additive fluctuations emanating from a background blackbody radiation source at thermal equilibrium. Thesis Supervisor: Jeffrey H. Shapiro Title: Julius A. Stratton Professor of Electrical Engineering 2 Acknowledgments First and foremost, I would like to extend my sincerest thanks to my research advisor Prof. Jeffrey H. Shapiro. This work has been made possible only with the continual guidance and encouragement I received from him over the past one and a half years. I am really impressed at the way he efficiently organizes his work, and devotes time to meet each of his graduate students individually on a regular basis, along with managing his numerous administrative responsibilities. I particularly adore his style of 'color-coding' his presentations on both black-board and paper. I thank my graduate advisor Prof. Dimitri P. Bertsekas, for his suggestions and discussions on my progress in graduate school, at the beginning of every term. I am grateful to my office mate Brent Yen for the numerous interesting discussion sessions we have had on a wide variety of topics ranging from quantum information to classical physics to number theory. The discussions we had during the 'ups and downs' of the thermal noise capacity proof were indeed fun. I would also like to thank Dr. Vittorio Giovannetti and Dr. Lorenzo Maccone, post-doctoral associates in RLE, for quite a few useful discussions I have had with them. My family and friends have always been very supportive towards all my endeavors so far. I thank my father Prof. Shambhu N. Guha, mother Mrs. Shikha Guha and my sister Somrita, for their continual encouragement and support over all these years. I am really grateful to my best friends Arindam and Pooja for having been there for me whenever I needed them, and for the nice times I relished spending with them over the last couple of years. Finally, I thank the Army Research Office for supporting this research through DoD Multidisciplinary University Research Initiative Grant No. DAAD-19-00-1-0177. 3 Contents 1 Introduction 14 2 Background 18 2.1 Free-Space Propagation Geometry: The Channel Model . . . . . . . . 3 . . . . . 18 2.1.1 Far Field . . . . . . . . . . . . . . . . . . . . 20 2.1.2 Near Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.2 Thermal Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.3 Quantum Detectors for Single Mode Bosonic States . . . . . . . . . . 22 2.3.1 Direct Detection . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.3.2 Homodyne Detection . . . . . . . . . . . . . . . . . . . . . . . 23 2.3.3 Heterodyne Detection . . . . . . . . . . . . . . . . . . . . . . 24 2.4 Classical Capacity of a Quantum Channel . . . . . . . . . . . . . . . 24 2.5 Entanglement-Assisted Capacity . . . . . . . . . . . . . . . . . . . . . 26 2.6 Wideband Channel Model . . . . . . . . . . . . . . . . . . . . . . . . 27 Capacity using Number State Inputs and Direct Detection Receivers 30 3.1 Number State Capacity: Lossless Channel . . . . . . . . . . . . . . . 30 3.2 Number State Capacity: Lossy Channel . . . . . . . . . . . . . . . . . 31 3.2.1 SM Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.2 Wideband Capacity . . . . . . . . . . . . . . . . . . . . . . . . 34 Constant High Loss and Low Input Power . . . . . . . . . . . 35 Constant Loss with Arbitrary Input Power . . . . . . . . . . . 37 4 4 4.1 4.2 Coherent State Capacity: Lossless Channel . . . . . . . . . . . . . . . 40 4.1.1 SM Capacity: Heterodyne Detection . . . . . . . . . . . . . . 40 4.1.2 SM Capacity: Homodyne Detection . . . . . . . . . . . . . . . 42 4.1.3 Wideband Capacity: Heterodyne and Homodyne Detection . . 43 4.1.4 Wideband Capacity: Restricted Bandwidth . . . . . . . . . . 43 Coherent State Capacity: Lossy Channel . . . . . . . . . . . . . . . . 47 4.2.1 SM Capacity: Homodyne and Heterodyne Detection . . . . . . 49 4.2.2 Wideband Capacity . . . . . . . . . . . . . . . . . . . . . . . . 49 Frequency Independent Loss . . . . . . . . . . . . . . . . . . . 50 . . . . . . . . . . 53 Free-Space Channel: Near-Field Propagation . . . . . . . . . . 57 . 61 General Loss Model: A Wideband Analysis . . . . . . . . . . . 64 Free-Space Channel: Far-Field Propagation Free-Space Channel: Single Spatial Mode - Ultra Wideband 5 39 Capacity using Coherent State Inputs and Structured Receivers Capacity using Squeezed State Inputs and Homodyne Detection 5.1 5.2 68 Squeezed State Capacity: Lossless Channel . . . . . . . . . . . . . . . 69 5.1.1 SM Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 5.1.2 Wideband Capacity . . . . . . . . . . . . . . . . . . . . . . . . 71 Squeezed State Capacity: Lossy Channel . . . . . . . . . . . . . . . . 72 5.2.1 SM Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 5.2.2 Wideband Capacity . . . . . . . . . . . . . . . . . . . . . . . . 74 6 The Ultimate Classical Capacity of the Lossy Channel 76 6.1 Ultimate Capacity: SM and Wideband Lossy Channels . . . . . . . . 77 6.2 Free-Space Channel: Far-Field Propagation . . . . . . . . . . . . . . . 78 6.3 Asymptotic Optimality of Heterodyne Detection . . . . . . . . . . . . 81 6.3.1 Heterodyne Detection Optimality: Free-Space Far-Field Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.2 82 A Sufficient Condition for Asymptotic Optimality of Coherent State Encoding and Heterodyne Detection . . . . . . . . . . . 5 83 7 7.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 7.2 Ultimate Classical Capacity: Recent Advances . . . . . . . . . . . . . 91 . . . . . . . . . . . . . . 94 . . . . . . . . 98 7.2.1 7.3 7.4 8 88 Capacity of a Lossy Channel with Thermal Noise Number States: Majorization Result Capacity using known Transmitter-Receiver Structures 7.3.1 Number States and Direct Detection . . . . . . . . . . . . . . 98 7.3.2 Coherent States and Homodyne or Heterodyne Detection . . . 103 7.3.3 Squeezed States and Homodyne Detection . . . . . . . . . . . 104 Wideband Results and Discussion . . . . . . . . . . . . . . . . . . . . 105 Conclusions and Scope for Future Work 109 111 A Blahut-Arimoto Algorithm 6 List of Figures 2-1 Free space propagation geometry. AT and AR are the transmitter and receiver apertures respectively. . . . . . . . . . . . . . . . . . . . . . . 2-2 19 A single-mode lossy bosonic channel can be modelled as a beam-splitter with transmissivity rj. , is the input mode, I is the environment, which remains unexcited (in vacuum state) in the absence of any extraneous noise other than pure loss. 2-3 a is the output of the channel. . . . . . . . 20 Variation of the fractional power transfers (transmissivities) of the first five successive spatial modes as a function of normalized frequency of transmission f' = 2f/f, [23, 24]. The modes have been arranged in decreasing order, i.e., rm > r/2 > 173 > ... , and fc = cL/ /ATAR is a frequency normalization. . . . . . . . . . . . . . . . . . . . . . . . . . 3-1 21 Capacity (in bits per use) of the SM lossy number state channel computed for very low average input photon numbers, (a) Numerically using the 'Blahut-Arimoto Algorithm' for q7 = 0.001 [blue solid line], and (b) Using approximation (3.5) [black dashed line]. 3-2 . . . . . . . . 33 Capacity achieving probability distribution for r/ = 0.2 and Ft = 2.62 photons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 33 3-3 Capacity achieving probability distributions of photon numbers at the input of the SM lossy bosonic channel for three different values of transmissivity - = 0.1, 0.2, and 0.3 (the black, blue and green lines respectively). All three distributions correspond to the same average number constraint A = 2.62. As the transmissivity decreases, the peaks of the distribution spread out. . . . . . . . . . . . . . . . . . . . . . . 3-4 34 Capacity (in bits per use) CNS-DD(h), of the SM lossy number state channel computed numerically using the 'Blahut-Arimoto Algorithm', and plotted as a function of h, for r7= 0.2, 0.14, 0.1, and 0.06. . . . . 4-1 35 Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a lossless channel using coherent-state inputs and heterodyne detection. (a) 8 f m ,. > f,, (b) f m, < f. .... 46 4-2 This figure illustrates the effect of forced lower and upper cut-off frequencies (fmin and fmax), on the coherent state homodyne and heterothe critical power levels dyne capacities. Given values of fmin and fm,, (the input power at which the critical frequency f, hits the upper-cut off frequency fm., and beyond which power 'fills up' only between fmin and f m ,) have been denoted by Phlr and Pht respectively for the two cases. In this figure, we plot the normalized capacities C' - C/fma vs. a normalized (dimensionless) input power P' = P/hf2.. We normalize all frequencies by fma. and denote them by f' - f/fm.. The 'dash-dotted' (black) line represents the capacity achievable using either homodyne or heterodyne with unrestricted use of bandwidth (Eqn. (4.20)), i.e. C = (1/ In 2) V2P/h bits per second. The 'solid' (green) line, and the 'dashed' (blue) line represent homodyne and heterodyne capacities respectively for any set of cut-off frequencies satisfying fi = 0.2. For this value of f'in P'd = 0.16, and P'h = 0.32. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). 4-3 . . . . . . . . . . . . . . . . . . . . . 48 Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a channel with frequency independent transmissivity TI, using coherent-state inputs and heterodyne detection. (a) f m ax 4-4 > f , (b) fmax < fc. . . . . . . . .. ... . . . . . . . . . . . 51 Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a channel with frequency dependent transmissivity q OC f2, using coherent-state inputs and heterodyne de- tection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 54 4-5 Dependence of C' on P' for coherent-state inputs and heterodyne detection, for different values of the red (dotted) line is for for fj fi. fAni The blue (solid) line is for fi 0, = = 0.1 and the green (dashed) line is = 0.2. Observe that all the three curves coincide in the low power regime, because when P' < PC'rit C' does not depend on fnin. When P' > Pcrit, performance of the channel deteriorates because of restricting ourselves to a non-zero 4-6 f'.. . . . . . . . . . . . . . . . . . Dependence of C' on P'for coherent-state inputs with heterodyne and homodyne detection, for f' = 0. The blue (solid) line is for hetero- dyne detection, and the red (dashed) line is for homodyne detection. . 4-7 56 57 This figure illustrates the coherent state homodyne and heterodyne capacities for the free-space optical channel in the near-field regime. All the frequencies in the transmission range f E [fmin, fma] satisfy Df > 1. In this curve, a normalized capacity C' = C/Af, has been plotted against a dimensionless normalized input power P' = P/(Ahfm), where A = ATAR/C 2 L 2 is a geometrical parameter of the channel (Fig. the critical power levels (the input 2-1). Given values of fmin and fm,, power at which the critical frequency fmax, f, hits the upper-cut off frequency and beyond which power 'fills up' only between fmin and have been denoted by Phr and Ph, f m a) respectively for the two cases. The 'solid' (black and blue) line, and the 'dashed' (red and green) line represent homodyne and heterodyne capacities respectively for any set of cut-off frequencies satisfying fj,1, = 0.3. For this value of fmi, the values of the normalized critical input power are: p'crit = 0.019, and p'it = 0.076 respectively. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). . . . . . . . . . 4-8 62 An example of ultra wideband capacities for coherent state encoding with homodyne or heterodyne detection, for the free-space optical channel under the assumption of a single spatial mode transmission. . 10 65 4-9 The optimal power allocation for the wideband coherent state channel with heterodyne detection depicted as 'water-filling', for frequency dependent transmissivity shown for ql in Fig. 2-3. . . . . . . . . . . . 5-1 67 This figure shows comparative performance of squeezed state encoding and homodyne detection with respect to that of coherent state encoding with homodyne and heterodyne detection. The 'solid' lines (red, green and grey) represent coherent state heterodyne, coherent state homodyne, and squeezed state homodyne capacities respectively for an arbitrary set of cut-off frequencies satisfying f' i= 0.2. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). For the above value of = 0.32, p' malized cut-off power are: P' f' i, the values of the nor- = 0.08, and p'crit = 0.16. The thick dashed (blue) line is the best achievable performance of the coherent state channel (with either heterodyne or homodyne detection), and is given by C' = (1/ In2) 2P'. The thick dotted (red) line is the best achievable performance of the squeezed state channel, and is given by C' = (1/In 2) 5-2 1. . . . . . . . . . . . . . . . . . . . . . . 73 An example of ultra wideband capacities for squeezed state channel and coherent state encoding with homodyne or heterodyne detection, for the free-space optical channel under the assumption of a single spatial mode transmission. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-1 Single Mode Capacities (in bits per use of the channel) for various cases. The transmissivity of the SM channel q = 0.16. . . . . . . . . . 6-2 75 78 Dependence of C' on P' for the ultimate classical capacity compared to C' for coherent-state inputs with heterodyne and homodyne detection. All curves assume f'is = 0. The blue (solid) line is the ultimate classical capacity, the red (dotted) line is for heterodyne detection, and the green (dashed) line is for homodyne detection. 11 . . . . . . . . 81 7-1 The thermal noise channel can be decomposed into a lossy bosonic channel without thermal noise (2.1), followed by a classical Gaussian noise channel 7-2 A plot of E' T 4 (1 ,)F (7.4). o Pn,k . . . . . . . . . . . . . . . . . . . . . . . for a SM thermal noise channel with transmissivity = 0.8, and N = 5 photons, as a function of m; for n = 0, 6, 12, 18,. The dark (black) lines correspond to the values n = 0, 30, 60, 90. 7-3 90 . . . 96 A plot of output entropies of the SM thermal noise channel as a function of 71 for number states inputs In)(n1, corresponding to n = 0, 1, 2, 3 and 4. The value of N = 5. We see that output entropy is minimum for the vacuum state 10) (0 , which is shown in the figure by the dark black line. The results for n > 0 were obtained by numerical evaluation of the pa, from (7.18); the n = 0 entropy result so obtained matches the analytic result g((1 - ,)N ). . . . . . . . . . . . . . . . . . . . . . . . 7-4 97 A plot of optimum apriori probability distributions (in dB) for a pure SM lossy channel (Tj = 0.8, h = 3.7), and a SM thermal noise channel (T = 0.8, h = 3.7, N = 5). In the pure lossy case, there are no peaks at this value of transmissivity and the distribution is close to linear, whereas for the thermal noise lossy channel, the distribution still has multiple peaks. For the optimum distribution of the thermal noise channel to converge sufficiently close to the Bose-Einstein distribution, y has to be even closer to unity. . . . . . . . . . . . . . . . . . . . . . 7-5 102 A sample calculation of the number-state direct-detection capacity (using the Blahut-Arimoto algorithm) of a SM lossy thermal noise channel with transmissivity 7 = 0.8, and mean thermal photon number N = 5. 102 7-6 A sample simulation result of all the different capacities of a SM lossy thermal noise channel we considered so far in the thesis. The channel SN has transmissivity y = 0.8, and mean thermal photon number N = 5.106 12 7-7 Ultra-wideband capacity of a lossy thermal noise channel - optical communication, single spatial mode propagation - cm 2 and AR = Free-space AT = 400 1 cm 2 , separated by L = 400 km of free-space; with frequencies of transmission ranging from that associated with a wavelength of 30 microns (which is deep into the far-field regime), to that associated with a wavelength of 300 nanometers (which is well within the near field regime). Mean thermal photon number N = 5 is assumed to remain constant across the entire range of frequencies. . . . . . . . 7-8 Capacities of a wideband lossy thermal noise channel 108 Free-space - all param- optical communication, single spatial mode propagation eters same as in Fig. 7-7, except that the mean thermal photon number N is now assumed to vary with frequency of transmission according . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Blahut-Arimoto Algorithm . . . . . . . . . . . . . . . . . . . . . . . . 113 to E qn. (7.2) A-1 f 13 Chapter 1 Introduction The objective of any communication system is to transfer information from one point to another in the most efficient manner, given the constraints of the available physical resources. In most communication systems, the transfer of information is done by superimposing the information onto an electromagnetic (EM) wave. The EM wave is known as the carrierand the process of superimposing information onto the carrier wave is known as modulation. The modulated carrier is then transmitted to the destination through a noisy medium, called the communication channel. At the receiver, the noisy wave is received and demodulated to retrieve the information as accurately as possible. Such systems are often characterized by the location of the frequency of the carrier wave in the electromagnetic spectrum. In radio systems for example, the carrier wave is selected from the radio frequency (RF) portion of the spectrum. In an optical communication system, the carrier wave is selected from the optical range of frequencies, which includes the infrared, visible light, and ultraviolet frequencies. The main advantage of communicating with optical frequencies is the potential increase in information that can be transmitted because of the possibility of harnessing an immense amount of bandwidth. The amount of information transmitted in any communication system depends directly on the bandwidth of the modulated carrier, which is usually a fraction of the center frequency of the carrier wave itself. Thus increasing the carrier frequency increases the available transmission bandwidth. For example, the frequencies in the optical range would typically have 14 a usable transmission bandwidth about three to four orders of magnitude greater than that of a carrier wave in the RF region. Another important advantage of optical communications relative to RF systems comes from their narrower transmitted beams [tRad beam divergences are possible with optical systems. These narrower beamwidths deliver power more efficiently to the receiver aperture. Narrow beams also enhance communication security by making it hard for an eavesdropper to intercept an appreciable amount of the transmitted power. Communicating with optical frequencies has some challenges associated with it as well. As optical frequencies are accompanied by extremely small wavelengths, the design of optical components require completely different design techniques than conventional microwave or RF communication systems. Also, the advantage of optical communication derives from its comparatively narrow beam, which introduces the difficulty of high-precision beam pointing. RF beams require much less pointing precision. Progress in the theoretical study of optical communication, the advent of the laser - a high-power optical carrier source the developments in the field of optical fiber-based communication, the development of novel wideband optical modulators and efficient detectors, have made optical communication emerge as a field of immense technological interest [1]. The subject of information theory addresses ultimate limits on communication. It tells us how to compute the maximum rate at which reliable data communication can be achieved over a particular communication channel by appropriately encoding and decoding the data. This rate is known as the channel capacity [20, 2, 31. It also tells us how to compute the maximum extent a given set of data can be compressed so that the original data can be recovered within a specific amount of distortion. Though information theory doesn't give us the exact algorithm (or the optimal code) that would achieve capacity on a given channel, and it doesn't tell us how to optimally compress a given set of data, nevertheless it sets ultimate limits on communication which are essential to meaningfully determine how well a real system is actually performing. The field of information theory was launched by Shannon's revolutionary 1948 paper [20], in which he laid down its complete foundation. Because light is quantum mechanical in nature, quantum information theory is 15 needed to assess the ultimate limits on optical communication. Much work has already been done on quantum information theory [13, 4], which sets ultimate limits on the rates of reliable communication of classical information and quantum information over quantum mechanical communication channels. As in classical information theory, quantum information theory does not tell us the transmitter and receiver structures that would achieve the best communication rate for a given kind of quantum mechanical noise. Nevertheless the limits set by quantum information theory are extremely useful in determining how well available technology can perform with respect to the theoretical limit. In this thesis, we will primarily focus on the lossy free-space quantum optical channel. Some examples of such a channel are satellite-to-satellite communication, wireless optical communication through air with a clear line-of-sight between transmitter and receiver units, and deep space communication. Long distance communication through a lossy fiber or a waveguide can also be treated along similar lines. Quantum information theory work to date has not treated the lossy channel case sufficiently well. This thesis will establish ultimate limits achievable on reliable communication of classical information over such a channel. In addition, it will study the best possible performance of standard transmitter and receiver structures in relation to the ultimate achievable capacity. It will also consider the effects of background thermal noise on the ultimate capacity and other achievable capacities of the channel. The organization of this thesis is as follows. In chapter 2, we elaborate the channel model in detail, establish our notation for the rest of the work, present a short review on the notion of thermal noise in single mode bosonic communication, discuss some standard quantum detection schemes, and review the notions of unassisted and entanglement-assisted classical capacity of a quantum channel. The capacity of the far-field lossy channel using photon number state inputs and unit-efficiency direct detection receivers is dealt with in chapter 3. Chapter 4 evaluates the capacity of the lossy channel using coherent state inputs with homodyne and heterodyne receivers. The calculations in this chapter have been done for a simple frequency-independent loss model and for a more realistic free-space optical channel model in which the 16 transmissivity of the channel depends on the frequency of transmission, in both the near-field and the far-field regimes. We look into the achievable rates using optimized quadrature-squeezed state inputs in Chapter 5. We establish the ultimate classical capacity of a lossy far-field channel in chapter 6, and work out the single mode and wideband capacities in the presence of thermal noise, in Chapter 7. The thesis concludes in chapter 8 with a note on ongoing and future directions of research in the field. 17 Chapter 2 Background This chapter is intended to provide all the necessary theoretical background and framework, with ample references to previous work in the area, for the material to be covered in the thesis. We discuss the model of the free-space quantum optical channel both in the far-field and the near-field propagation regimes. We describe the well known structured receivers for single mode (SM) bosonic states, and lay down their definitions mathematically. We also present a brief discussion on the notion of capacity for transmitting classical information over a quantum channel, followed by a discussion on entanglement assisted classical capacity. The chapter ends with a short note on the formulation of the power constrained wideband channel model that we use for all our subsequent capacity calculations. 2.1 Free-Space Propagation Geometry: The Channel Model Free-space optical communication has attracted considerable attention recently for a variety of applications [25, 26]. In the present work, we shall primarily focus on maximizing the rate at which classical information can be sent reliably over a freespace optical communication link under an average input power constraint by modulating various parameters characterizing the quantum states of the optical field 18 AT AR z-0 Free Space z=L Figure 2-1: Free space propagation geometry. AT and AR are the transmitter and receiver apertures respectively. modes at the transmitter, and by using different detection schemes at the receiver. Consider a line-of-sight free space optical communication channel with circular shaped transmitter and receiver apertures denoted AT and AR respectively, as shown in Fig. 2-1. A quantized source of radiation at the transmitter produces a field E(r, t) which is space limited to the transmitter exit aperture AT { (x, y) :x 2 +y 2 d0 2 /4 and time limited to a finite signalling interval To = {t : to - T < t < to}. After propagation through L m of free space, the field is collected by a receiver whose entrance aperture is AR {(x', y') : + y' 2 < dL2 /4}. The receiver is assumed to time limit the field in its entrance aperture to an interval TL t < to + L/c}. {t : to - T + L/c < The free space Fresnel number for transmission at frequency f is defined as Df = (ATAR/c 2 L 2 )f 2 , where with some abuse of notation, we are now using AT = irdo2 /4 and AR 2 = 7rdL /4 to denote the aperture areas at the transmitter and the receiver respectively. For narrowband transmission at a center frequency f, depending upon the magnitude of D1 , the propagation is broadly classified to be either in the 'far-field' (Df < 1), or the 'near-field' (Df > 1) regimes. 19 Environment c=J0a+ 1-q~b & Output Input Figure 2-2: A single-mode lossy bosonic channel can be modelled as a beam-splitter with transmissivity 7. e is the input mode, b is the environment, which remains unexcited (in vacuum state) in the absence of any extraneous noise other than pure loss. is the output of the channel. 2.1.1 Far Field In the far-field regime, only one spatial mode of the input bosonic field couples appreciable amount of power to the receiver aperture [5], and all other spatial modes are thus insignificant. Further, it has been shown that [5], in the Heisenberg picture, quantum description of the propagation of a single spatial mode of the field through this channel in the far-field regime is obtained by coupling the input field mode & into a beam splitter of transmissivity 7 = Df, with an 'environment' mode b (Fig. 2-2), which is in the vacuum state (2.1); i.e. = d&+ /1 - 7 b, (2.1) where (ATAR f2 c2 L2 ) and d, 22 , and b are the modal annihilation operators of the input, output and the noise modes respectively. Several authors like to call the above model a single-mode lossy bosonic channel, where 1 - q is identified as the 'loss'. Physically, q is the fraction of the input power that gets coupled to the receiver aperture. quadratic dependence on the carrier frequency 20 f in the far-field. Note that, q has a . 1.000 0.00 0 0 - 0.600- IU 0.400 0.200 - eta_1 eta_2 eta_3 eta_4 eta_5 - - 0.000 0.000 2.000 4.000 6.000 8.000 10.000 12.000 normalized frequency (f-prime) Figure 2-3: Variation of the fractional power transfers (transmissivities) of the first five successive spatial modes as a function of normalized frequency of transmission f = 2f/fc [23, 24]. The modes have been arranged in decreasing order, i.e., rq1 > 772 > 773 > ... , and f, = cL/vA ATAR is a frequency normalization. 2.1.2 Near Field As the value of Df increases, more and more spatial modes of the field start coupling appreciable power into the receiver. If we arrange the fractional power transfers of the spatial modes in decreasing order and denote the fractional power transfer of the ith mode by 7ij, then - from the eigenvalue behavior of the classical free-space modal decomposition theory [23, 24] - the qj's depend upon the carrier frequency f as shown in Fig. 2-3. When Df > 1, near-field propagation prevails. In this regime, there are Df spatial modes which have e7 ~ 1, and all the other modes are insignificant [5]. 21 2.2 Thermal Noise In the absence of any extraneous noise sources other than pure loss, the environment mode b (2.1) is in the vacuum state, which is the minimum 'noise' required by quantum mechanics to preserve correct commutator brackets of annihilation operators of the output modes. Different types of noise can be described by different initial states of the environment mode 6. We define a thermal noise lossy bosonic channel to be a channel that describes the effect of coupling the lossy channel to a thermal reservoir at temperature T [7]. The thermal environment mode 6 is excited in the state given by the density matrix PT = - exp A 2 a, 2a)(d (2.3) where 1a) are the coherent state vectors and S is the mean photon number per mode of the thermal state PT, given by the Planck's formula for the mean number of photons per mode of a blackbody radiation source at thermal equilibrium at temperature T: = N 1r~~uu) = hf/JkBT -1 24 where h and kB are the Planck's constant and Boltzmann's constant respectively. 2.3 Quantum Detectors for Single Mode Bosonic States An arbitrary quantum measurement on a single EM field mode can be described in terms of a set of Hermitian operators {lb}, which satisfy (positivity), (2.5) (completeness). (2.6) b ;> 0 nib = f b 22 These operators thus comprise a Positive Operator Value Measure (POVM). When field mode is in the state given by the density operator /a, the probability of an outcome b is given by (2.7) Pr(bla) = Tr(I afb). Some of the basic quantum measurement schemes for single mode (SM) bosonic states that we will consider in this thesis are (a) Direct Detection, (b) Homodyne detection, and (c) Heterodyne Detection. Their POVM descriptions are given below [6]. 2.3.1 Direct Detection Direct detection (or photon counting) refers to a widely used method of optical demodulation that responds to a short time averaged power of the optical field. Quantum mechanically, it refers to a measurement done in the number-state basis {in),n = 0,1, 2,...}. The POVM corresponding to Direct Detection is given by H 2.3.2 for n=0,1,2,... = In)(ni, (2.8) Homodyne Detection Homodyne Detection refers to a quantum detection scheme in which the incoming field is mixed with a strong local oscillator (LO) field through a 50-50 beam-splitter (BS). The LO field is spatially and temporally coherent with the incoming field. The outputs of the BS are detected by two ideal photon counters. The output of the homodyne detector is the scaled difference of the output of the two photon counters. Homodyne detection measures one quadrature of the field in a given direction in phase-space, depending upon the phase of the LO-field. Homodyne detection can be described by the POVM nm = Iai)(oCiI, 23 for ai E R, (2.9) where 11, are projection operators onto the 6-function normalized eigenstates lai) of the quadrature component ai = Re(a), that is in phase with the LO field whose phase, for convenience, has been set to zero. 2.3.3 Heterodyne Detection Heterodyne detection refers to a simultaneous measurement of both the quadratures of the optical field. It mixes the received signal with a strong LO, whose center frequency is offset by a radio frequency -- known as the intermediate frequency (IF) from that of the signal, on a 50-50 BS. Quadrature detection on the IF difference signal from photodetectors placed at the BS output ports then yields the desired measurements. Heterodyne detection can be described by the POVM 1 EH = -c)(aL, 7r (2.10) for aE C, where 1a) is a coherent state. 2.4 Classical Capacity of a Quantum Channel Let us consider an information source that emits discrete independent, identicallydistributed (i.i.d.) classical symbols, indexed by i = 1,2,..., A. The source emits one letter at a time with a probability distribution PN, where PN(i) is the apriori probability that the symbol i is emitted. Our goal is to transmit the output of this source in the best possible manner over our quantum channel (2.1) using appropriate quantum states for encoding and an optimum detection scheme at the receiver. In particular, we wish to compute the capacity for transmitting classical information in bits, per use of a lossy quantum channel. Let us assume that we excite the field mode & (in one signalling interval), in the quantum state ,i E 'R to encode the source-symbol i. Let us denote the state at the output of the channel by <P(i) where the channel <P is a completely positive (CP) trace preserving map. A block code of length n would consist of codewords that in general 24 could be an entangled state 3 E At the output, the most general quantum Ro(n mechanical detection scheme can be described in terms of a POVM, which we shall call a 'decision rule' and denote by X = {X 3 }. Such a generalized measurement can be thought of as a measurement of a quantity X whose possible results of measurement j = 1, 2, ... , M, which can be treated as an index for a set of classical output symbols. The conditional probabilities that a particular output symbol is are labelled by received given that a particular input symbol was transmitted through the channel, i.e. the classical transition probabilities [9, 11, 12] are given by (2.11) P(ji) = Tr(I(pA)X) The classical mutual information and the quantum Holevo information for single use of this channel are respectively given by [13] pN()PJ I(pN, (, X) = E AH(PN where S() = -Tr( log (ZPN(i)i =S P(k) pN(i)S(i -E (2.12) (2.13) log ,) is the Von Neumann entropy of the state 3. We define [13] C (1) = sup sup I(pN, 4*", X) {PNI (2.14) X C,(@)= sup AH(pN, I n), (2.15) {PN(i)} where (DO' = (D 0 4D 0 ... 0 1 represents multiple uses of the channel, and PN is a probability distribution over all states in H71. It can be shown that both of the above quantities are superadditive and that their averages with respect to n, as n -+ oc exist [13]. From Shannon theory we know that, at any rate less than C(I), using product state inputs over many uses of the channel to encode the source letters, one can achieve arbitrarily low error probability by using suitable classical coding techniques 25 [20]. We define CcQ to be the classical information capacity of a quantum channel 1 when the transmitter can encode messages using product state inputs and the receiver can decode using measurements entangled over arbitrarily many uses of the channel. The Holevo-Schumacher-Westmoreland (HSW) Theorem states that [13, 14]. (2.16) CCQ = Ci(D). The ultimate capacity of a quantum channel 1 to communicate classical information is the supremum of the CCQ of n uses of ID (i.e. (pon), divided by n. We call this capacity the 'Classical Capacity' of a quantum channel, and denote it by CQQ or simply by C. The subscript 'QQ' signifies the fact that input-states and decoding can both be entangled over a large number of channel uses. So, the classical capacity of a quantum channel is C = CQQ = sup n 2.5 n (2.17) Entanglement-Assisted Capacity A majority of the striking advantages that quantum information theory enjoys over its classical counterpart stem from the wonders that quantum entanglement [4] can do to communication protocols. Teleportation and Superdense Coding [4] are two such examples. Surprisingly, it turns out that the classical information carrying capacity of a quantum channel can be increased way beyond its Classical Capacity (2.17), by using prior shared entanglement as a resource. The entanglement-assisted classical capacity CE is defined to be the maximum asymptotic rate of reliable bit transmission with the aid of unlimited amount of prior entanglement shared between the transmitter and the receiver. The importance of CE lies in the fact that it gives an upper bound to all the relevant capacities of the quantum channel, including the quantum and classical (unassisted) capacities [21]. In can be shown [21] that the natural generalization of the concept of mutual information between input and output of a channel, to the quantum case, is given 26 by the Quantum Mutual Information I(, ,) of a quantum channel (D, where 3 is the input state, given by I(,y) = S() + S(I'()) - S((W 0 T)PP) (2.18) where P is a purificationi of the input density matrix ,. It can be shown that the entanglement-assisted capacity CE is given by CE= max I(1,D) (2.19) where the maximum is taken over all possible inputs / to the channel [21]. 2.6 Wideband Channel Model The study of quantum limits to wideband bosonic communication rates has been one of the primary applications of quantum communication theory [15, 8]. The wideband capacity of a power-constrained lossless bosonic channel has been shown to be proportional to the square root of the total input power P [8]. Until very recently [18], the exact solution to the wideband classical capacity of the lossy bosonic channel was not known. In this thesis, we will demonstrate the calculation of the wideband classical capacities of the free-space channel (with and without thermal noise) for a variety of cases, viz. ultimate classical capacity, entanglement-assisted capacity, capacity achieved by using a coherent state encoding with heterodyne or homodyne detection, and capacity achieved by using an encoding based on an optimized set of quadrature squeezed states and homodyne detection. For all these cases, the procedure we use to compute the input-power-constrained wideband capacities from the respective SM (bits per use) capacities is essentially the same. The procedure is illustrated below. Let us consider an input-power-constrained frequency multiplexed channel, in 'Purification of a mixed state is a purely mathematical process [4]: one can show that given any system A in a mixed state k^, it is always possible to introduce an additional reference system R and a pure state in the combined state space JAR), such that this pure state reduces to yA when one looks at system A alone, i.e. pA - Tr(IAR)(ARI). 27 which each frequency bin (of width b Hz) is in general a SM thermal noise channel characterized by a transmissivity qi and an input-output relation similar to (2.1). In the following discussion, we shall consider communication using a single spatial mode of the bosonic field at each frequency f E [fmin, fmax], where fmin and fm, for a real bosonic channel would primarily depend upon practical limitations on transmitter and receiver design, and on the optimum power allocation corresponding to the given set of transmitter and receiver structures. We denote the center frequency of the ith frequency bin by fi and use bfih for the average photon transmission rate in the ith bin. The average input power constraint is then given by P= b (2.20) hfiii. In far field free-space propagation ,the transmissivity of the ith bin is ij = 2 where A - ATAR/C 2 L 2 (2.2). Let us denote by Ri the average number of thermal photons in the ith bin of the environment; Ni is given by (2.4) with f = fi. If we denote the SM capacity (in bits per use) of the ith frequency bin by C(7i, fii, Si), then the wideband capacity CWB can be obtained as a function of the average input power constraint P, using a constrained maxization: CwB(P) = MaxZC (qi(fi), fi, (M) (2.21) subject to b hfig ; P. (2.22) The above maximization can be done using a Lagrange multiplier technique either analytically or numerically, depending upon the complexity of the SM capacity formula C (rh(fi), fii, Si(fi)) for the particular kind of capacity in question. Note that this formulation assumes that different frequency bins are employed as parallel channels. Because of superadditivity, it must be shown that optimum capacity, for our channel, does not require entangling different frequency bins. Of 28 course, when seeking the capacity of specific transmitter-structure/receiver-structure combinations, we are free to assume that these structures do not entangle different frequency bins. In the subsequent chapters, we first work through the capacity calculations of all the different cases without thermal noise. We later summarize all the thermal noise capacity results together in one chapter. 29 Chapter 3 Capacity using Number State Inputs and Direct Detection Receivers Number states are eigenstates of the photon number operator N = &f&, where d is the annihilation operator associated with a single mode of the bosonic field. The number state In) is the eigenstate associated with the eigenvalue of n photons. Clearly, as is a Hermitian operator, it is an observable, and its eigenstates {In); n = N 1, 2, ..., oo} form a basis of the entire state space N. A direct detection receiver (or an ideal photon counter) is a device that performs a measurement of N with unit efficiency, on an incident single mode (SM) field. If the state incident on such a receiver is the number state In), then the result of measurement is n with certainty. 3.1 Number State Capacity: Lossless Channel Consider a zero-temperature SM lossless (rj = 1) bosonic channel (2.1), with a maximum average number of photons h, at the input. It is well known that classical capacity (2.17) can be achieved for this channel using photon number states and a unit-efficiency direct detection receiver [15]. The probability distribution of number states that achieves capacity is the Bose-Einstein distribution, given by 30 PN(n) = 1+h 1+h (3.1) for n = 0,1,2,... _ ,n where PN(n) is the probability of transmitting the state In)(nI in one use of the channel. The capacity (in bits per use) of this SM number state channel is given by [15] C(") = g(h) = "-1) log 2 (1+ (3.2) ). + log 2 (1 + We also know that the ultimate classical capacity (in bits per second) of the wideband lossless channel with an average input-power constraint P (in Watts) is given by [15] (3.3) V Cultimate(P) which can be achieved by number state encoding and direct detection [15]. In the rest of this chapter, we shall investigate the impact of loss on the capacities of SM and wideband channels with number state transmission and direct detection. 3.2 Number State Capacity: Lossy Channel Let us transmit a pure number state PIN = In)(nj with probability pN(n) through a zero-temperature SM lossy channel (2.1) with transmissivity 'q. It can be shown [4], that the output is a mixed state given by POUT = E'=o (nk(i - )n~kIk)(kI, when we transmit In). If we detect this state using a unit-efficiency direct detection receiver, the probability of getting m counts is (n)rp,(I - ' for m = 0, 1, 2,..., n. So, the classical transition probabilities (2.11) for a single use of the channel can be written as for m > n 0 (2)r"(1 31 - ) for m ;m. 3.2.1 SM Capacity To get the SM capacity (in bits per use), we have to maximize the classical mutual information I(M; N) (2.12) over all input probability distributions pN(r) subject to a fixed value h of average photon-number at the input. This constrained maximization problem can be solved numerically using an iterative algorithm, known as the Blahut- Arimoto (BA) Algorithm [16] (See Appendix A for the details of the algorithm). For the special case Tpi << 1 (low-power, high-loss), some approximate analytical results can be obtained. In this regime, the SM capacity (in bits per use) of the number state channel is given by C ~ H(7pi) where H(x) H(7), (3.5) X log 2 (x) - (1 - X) log 2 (1 - X) is the binary entropy function. The results of a Blahut-Arimoto simulation for SM capacity in this regime has been compared with the above approximation in Fig. 3-1. Interestingly, on carrying out the BA algorithm [16] for the general case, we find that the capacity achieving probability distributions have multiple peaks (See Fig. 3-2). This suggests that to achieve the best capacity using number state inputs, one would almost preferentially use a set of optimum numbers of photons in each channel use. A sample plot of the optimal probability distribution for 7 = 0.2 and i! = 2.62 is given in Fig. 3-2. A qualitative explanation of the appearance of these peaks in the optimum distribution is as follows. When photons are sent through the lossy channel in bunches of known magnitudes, the effect of the loss is to spread each bunch by some amount about its mean value. If the initial bunches are spaced out well enough, the photon counting receiver is able to resolve them with little or no ambiguity. Thus, we should expect that the peaks of the probability distributions would space out more and more as we increase the loss (i.e., decrease the transmissivity 71), as lesser loss would imply lesser 'spread' of the peaks, and hence for the same level of 'ambiguity' at the receiver, we could afford to have the peaks closer together. Though the above 32 5.000e-4 4.000e-4 WI 3.2OOe-4 - 1.000e-4 - Approximation - Blahut Arimoto -0- 1000 U. f I- 0.000 0.020 0.080 0.060 0.040 I I I I I 0.100 Ii I 0.120 i 1 0.140 Average Photon Number (nbar) Figure 3-1: Capacity (in bits per use) of the SM lossy number state channel computed for very low average input photon numbers, (a) Numerically using the 'BlahutArimoto Algorithm' for r = 0.001 [blue solid line], and (b) Using approximation (3.5) [black dashed line]. 11111111 0 -10 -20 -30 -40 E -50 0 -60 n-2 -70 -80 -90 0 I 20 A. 40 1 60 . 100 80 Photon number (n) 120 140 160 Figure 3-2: Capacity achieving probability distribution for r7 = 0.2 and A = 2.62 photons. 33 0.000 eta = 0.1 -10.000 _ -20.000 - -30.000 - -40.000 - Ct I/ - ~ ~I *\ .~ -50.000 If ~{ \\ V / X *\ -60.000 0.000 - eta = 0.2-eta =0.3 20.000 40.000 60.000 A 80.000 100.000 120.000 number of photons (n) Figure 3-3: Capacity achieving probability distributions of photon numbers at the input of the SM lossy bosonic channel for three different values of transmissivity --- 77 = 0.1,0.2, and 0.3 (the black, blue and green lines respectively). All three distributions correspond to the same average number constraint ft = 2.62. As the transmissivity decreases, the peaks of the distribution spread out. explanation is not completely rigorous, this intuitive idea has been illustrated by actual BA simulation results for the optimum distributions for 27 (Fig. 3-3). = 0.1, 0.2, and 0.3. As one increases the transmissivity to unity, the multiple peaks in the optimum probability distribution gradually merge, and the distribution converges to the Bose-Einstein distribution (3.1). Finally, a plot of the capacity (in bits per use) CNS-DD(!) as a function of h for various values of q is given in Fig. 3-4. As one would expect, for the same value of ii, the capacity increases with increasing transmissivity 71. 3.2.2 Wideband Capacity The wideband capacity of the number state channel can be obtained numerically by maximizing the sum of the SM capacities (obtained using the BA algorithm) 34 2- 0.50 eta = e eta = 0.2 0.14 eta = eta = I I 0 20 10 30 I 0.1 0.06 I 40 Average photon number Figure 3-4: Capacity (in bits per use) CNS-DD(5i), of the SM lossy number state channel computed numerically using the 'Blahut-Arimoto Algorithm', and plotted as a function of h, for q = 0.2, 0.14, 0.1, and 0.06. across the entire frequency range of transmission, subject to the average input power constraint P (Watts), using the Lagrange Multiplier technique described in Section 2.6. We will consider two cases (a) Constant high loss (frequency independent) and low input power at all frequencies; and (b) Frequency independent loss with arbitrary input power. The numerical analysis of the frequency dependent wideband far-field free-space channel turns out to be too complicated for the number state inputs. Constant High Loss and Low Input Power Let us assume that the total input power P is so small that the mean photon numbers fi < 1, Vi, where i is a discrete index corresponding to a narrow frequency bin of width b Hz centered at fi (see Section 2.6). Assume that the transmissivity 7 < 1 and is same for all frequencies. Carrying out the constrained maximization (2.21) for this case, and incorporating appropriate approximations, we obtain the optimum distribution of photons across frequencies as 35 [i+ AhI W = exp f1 , (3.6) where A is the Lagrange multiplier. Substituting this result into the the power constraint equation (2.22), and approximating the sums by integrals by assuming a vanishingly small bin width, we obtain P = fi exp (-[ + Ohfi]) bh (3.8) Xe S (3.7) Ihf-'(3.9) eh021 where 3 = Aln 2/7, and allowing the integration to be over all frequencies is permissible because the contribution of the optimum average photon number becomes vanishingly small as f -- oc. Substituting this relation into the capacity equation (3.5), and approximating the sums by integrals by assuming a vanishingly small bin width, we obtain C = (3.10) [H(pfi) - fiH(77)] b Sb [- - (1-iij) ln (nit) In (1- 71i) 71ln 77+ i(1 - 7) ln(1 - 7)] + [-ri 1n1 - qhi ln hi + (1 - 7nh) In 3 3hln 2 j 0 i In i nh + ifn ln 77 - r77i (1 - I)] (neglecting terms of second order in 71) 217 (1 + x)e-dx = flhln 2~ 36 (3.11) (3.12) Eliminating the Lagrange multiplier coefficient f from the power equation and the capacity equation, we obtain the wideband capacity CWB in bits/sec as a function of input power P: CWB- 2T2 (3.13) P Note that in the limit of low power and high loss, the wideband capacity of a lossy channel with frequency independent transmissivity, achieved by number state encoding and direct detection, is proportional to the transmissivity 77. Constant Loss with Arbitrary Input Power Now, let us consider the general case of arbitrary transmissivity 77, without any restriction on the input power constraint P. The analysis developed in this section can be used to compute the wideband capacity with frequency independent loss for any transmitter-detector combination for which the SM capacity (in bits per use) is known. Let us set up the Lagrangian J(ni, A) as follows: J(hi, A) = b C(, hi) + A(P - b hfih), where C(TI, hi) is the SM capacity in bits per use. Setting &J(nii, (3.14) A)/Ohi = 0, we obtain ____ - Oni (3.15) Ahfi. The above equation can be inverted to express hi as a function of Afi. Making the substitution z = Af, and converting the sums to integrals by going to the limit of vanishingly small bin width, we can define h(z) as a function with a continuous argument. Let us define a function m(z) = zn(z). Also, as we know C(rj, h(z)), we may express the capacity as a function of z, say C,,(z). Now, going through the algebra, one can show that the wideband capacity CwB satisfies 37 CWB, P fo Cizdz 1 (3.16) f" m(z)dz Given the SM capacity, the term in the square braces is just a positive constant which can be evaluated easily. For rI = 0.2 for all frequencies, the capacity was calculated to be CwB,=0.2 CWB,Iossless. = 0.5877/P/h, which is almost r7 times the lossless capacity So, similar to the 'low-power high-loss' case, at r7 = 0.2 with arbitrary input power P, the wideband number state capacity of the lossy channel (with frequency independent loss) is almost proportional to the transmissivity. We later show that the ultimate classical capacity of the wideband lossy channel C = V/CwB,ossIess, which can be achieved by a coherent-state encoding. one important conclusion is that transmissivity - So, at least for the case of frequency-independent in contrast to the lossless case, number state encoding and direct detection is not optimal for the lossy channel. Another important observation from the above analysis is that, for frequency independent transmissivity, the wideband classical capacity of any lossy channel can be expressed as a constant times VP/h, where P is the constraint on total input power in Watts. 38 Chapter 4 Capacity using Coherent State Inputs and Structured Receivers Coherent states {fla)} are defined by the eigenvalue equation for the non-Hermitian annihilation operator d, ja) with, in general, complex eigenvalues {a}. coherent state Ia) (4.1) = ala), The mean complex amplitude of the is given by (d) = a = ai + ia 2 = (el) + i(e 2 ), where ei and &2 are the normalized quadrature operators. Coherent states are quadrature minimum uncertainty states each of whose quadrature components are uncorrelated and have variances ((A&i) 2 ) = ((Aa2)2) = 1/4. One can derive the number-state expansion of a coherent state [15], Ia) = e- /2 -In). (4.2) The two other properties of coherent states that we will use for our analysis are the inner product of two coherent states and the overcompleteness property: 39 (3ce) = e(aO*-a*O)/2e-a-012 /2, = Jd a)(a (4.3) (4.4) 7r In this chapter, we will investigate the classical capacity of the SM lossy channel and the wideband free-space optical channel in both far-field and near-field propagation regimes, achievable using coherent state encoding with homodyne or heterodyne detection (see Section 2.3) at the receivers. The lossless capacities for coherent state inputs have been dealt with in detail in [15], and are explained briefly in the next section. 4.1 Coherent State Capacity: Lossless Channel Let us transmit a coherent state 1, = 1a)(al through a zero-temperature SM lossless (rq = 1) bosonic channel (2.1), with a probability measure pA(a)d2a. Here, A is a complex random variable representing the input alphabet, which takes on the value a, if the coherent state 1a) is transmitted in a single use of the channel. Thus, the unconditional channel density operator (per use) takes the form p = JPA(a)Pad2a, (4.5) and the average input power constraint can be written as i = Tr( &'&) = Jd2aja2pA(a). 4.1.1 (4.6) SM Capacity: Heterodyne Detection As we discuss in Section 2.3, ideal heterodyne detection involves a simultaneous measurement of both quadratures of the field mode, and is described by the POVM [6] 40 U1 (4.7) = 1-11)(I1. 7r The conditional probability density to read out 3 = 1, + i0 2 at the output of the heterodyne detector, given that Ia) was PBIA(f 3 I&e) = transmitted is given by (4.8) Tr(&1 ) 1(#a)12 (4.9) 7r 1 - 7r exp (-13 a12), - (4.10) where B is a complex random variable representing the output alphabet of the channel. From Eqn. (4.10), we conclude that this channel can be looked upon as two independent and identical, parallel additive Gaussian noise channels (corresponding to the two quadratures) with mean squared amplitude constraint of h/2 at the input, and a noise variance of 1/2 for each of the two channels. We know from Shannon theory that the capacity of an additive white Gaussian noise (AWGN) channel is achieved by a Gaussian input probability density. So, the capacity achieving input distribution is given by PA (a) = 7rn (4.11) ) exp(- which makes the unconditional channel density operator (4.5), a thermal state with mean photon number i! (2.3). The SM capacity in bits per use of this channel can be obtained using the classical Shannon formula [20], by adding the contributions of the two quadratures: Ccoherent--heterodyne(f) + g109 2 ( = 2 = log2 (1 + 41 ). (4.12) (4.13) 4.1.2 SM Capacity: Homodyne Detection As we discussed in Section 2.3, ideal homodyne detection involves the measurement of a single quadrature of the field mode, and for the case when the local oscillator phase is oriented with the d, quadrature, is described by the POVM [6] (4.14) N13= If1)(1 The conditional probability density to read out 13,at the output of the homodyne detector, given that Ia) was transmitted is given by PBIAC(311a) = I(Qu1a)1 2 - 27r (1) (exp (1 )2 2(1)4 , (4.15) where we have assumed a is real. From the above equation, we see that this channel is equivalent to an AWGN channel with mean squared amplitude constraint of ft photons at the input, and noise variance, of 1/4. The SM capacity (in bits per use) of the lossless channel for a coherent state input with homodyne detection is thus given by log 2 ( Ccoherent-homodyne(i) - + ) 1 log 2 (1 + 4Ai). 2 (4.17) Note that, at high values of h, 'coherent-heterodyne' does better than 'coherenthomodyne', whereas at small values of h, 'coherent-homodyne' does better. This is because homodyne detection has less noise but less 'bandwidth' than heterodyne detection. So, at low h values, where low-noise operation is more important than 'bandwidth', homodyne detection yields a higher capacity. The opposite situation prevails at high h values, where 'bandwidth' is at a premium, so that heterodyne detection yields a higher capacity. 42 4.1.3 Wideband Capacity: Heterodyne and Homodyne Detection If P is the constraint on the maximum average input power in Watts (averaged across frequencies), then the wideband capacities of the lossless bosonic channel using coherent state encoding with homodyne and heterodyne detection can be calculated using the constrained maximization procedure explained in Section 2.6. The various known wideband capacities (in bits per second) of the lossless channel are summarized below [15]. Cuitimate(P) = r 2P In 2 3h -1 2P Ccoherent -heterodyne(P) Ccoherent-homodyne(P) I (4.18) , - h Note that interestingly, the capacity expressions for Ccoherent-heterodyne(P) Ccoherent-homodyne(P) (.9 9 (4.20) and are identical. Also, we observe that Cuitimate(P) Ccoherent-heterodyne (P) 7r (4.21) N/3 and thus the performance of heterodyne detection is consistently suboptimal at all input power levels. In won't be out of context here to mention that, later in this thesis, we prove a sufficient condition for the lossy wideband channel capacity with heterodyne detection and coherent state encoding to approach the ultimate classical capacity as the input power P -+ oc. 4.1.4 Wideband Capacity: Restricted Bandwidth The wideband capacities for the lossless channel in the previous sub-section were all calculated under the assumption that we have at our disposal photons at all frequencies in the unbounded interval (0, oo), which is not the case in practice. Also 43 it turns out that, in the 'ultimate capacity' case, no matter what the magnitude of the input power P is, the capacity achieving power allocation uses non-zero power at all frequencies f E (0, oc). Whereas, in the case of either heterodyne or homodyne detection, it can be shown that for a given input power level P, the best distribution populates only up to a certain cut-off frequency fe, though the cut-off frequencies for homodyne and heterodyne are different for a given P. Interestingly, as noted in the previous sub-section, after we solve the constrained maximization problem by integrating the respective optimal power allocation and SM capacity expressions between f E (0, fc], the wideband capacity expressions for Ccoherent-homodyne(P) Ccoherent heterodyne(P) and are found to be identical. What happens to the homodyne and heterodyne capacities if we now impose a pair of lower and an upper cut-off frequencies of transmission? Interestingly, it turns out that in this case homodyne and heterodyne wideband capacities are no longer the same. Homodyne performs better at lower values of P, whereas heterodyne wins at higher values of P. Let us work out the wideband capacities for these two cases in detail to see why this happens. The analysis for heterodyne and homodyne capacities are similar. So, we will first work out the heterodyne capacity in detail, then state the homodyne result and compare the two cases. Maximizing the overall information rate for the wideband lossless channel with coherent state inputs and heterodyne detection ln(1 + fij), C= (4.22) subject to total input power P (2.20), we obtain the optimum intensity (W/Hz) distribution as Ih - P b - hfi,* = 1o - hfi, (4.23) where Io = (1/A In 2) and A is a Lagrange multiplier. Replacing the sums by integrals, and imposing Kuhn-Tucker like conditions, we can write the optimal intensity variation as 44 1(f) = {l'o - hf, if > 0 (4.24) otherwise 0, with input power P given by fmax P= J I(f)df. (4.25) fmin The allocation of intensity across different frequencies in the transmission band can be looked upon as a 'water-filling' solution to an infinite set of classical parallel additive white Gaussian noise (AWGN) channels [2] (Fig. 4-1(a)). For a given value of 1 o, due to the lower cut-off frequency constraint, and the optimal intensity expression, f non-zero power is allocated only in the region frequency cut-off fm , is greater than intensity I(f) in the region f where f, = 1 0 /h. Integrating the optimal E [fmin, fc], we get, hf)df = - (fc 2 - 2fcfmin + fmin 2 ). f=Ioh (Io - P f, E [fmin, fc], as long as the upper (4.26) 2 fmin From this result it follows that fc = 2P + fmin. (4.27) 4.7 As the input power P is increased from zero upwards, f m in, until it hits the upper cut-off frequency f m .. fc increases starting from The value of P for which this happens will be termed the 'critical power' for heterodyne capacity, which we denote as Pat . In order to simplify the notation, let us normalize the input power by hf2x to obtain a dimensionless quantity P' = P/hfm2.. Also, let us define a normalized wideband capacity C' = C/fm., and normalize all frequencies by fmax and denote them by f' f/f m ,. It is easy to deduce from Eqn. (4.27), that p' (1 45 - f .i (4.28) het hf hf I0 I I I I I I I I I I I I I I 0 Z Z: ~T-~- -- fmin TZ 7__:_ / f f max f fmin f max 10 10 fch h (b) (a) Figure 4-1: Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a lossless channel using coherent-state inputs and heterodyne detection. (a) f m a fc, (b) f m a < fc. When P' is increased beyond P't, fc keeps increasing beyond fm , but the power 'fills up' vertically in the region f E [fmin, f mn], as shown in Fig. 4-1(b). The above equations can be solved analytically to compute the normalized wideband capacity Ccoherent-heterodyne(P') as a function of the normalized input power P'. For P' < P'cr, we have Ccoherent-heterodyne(PI) and fm. where f'= f/fm, = v2P + fn Ccoherent-heterodyne where f m r < fe, and fc. For P' >P'i, we get fc f min In 2 f' is given f',= (4.29) m 2P/ - fnin = ( Jf) fMm (4.30) n by 1(1 + f' in) + (1 P' - (4.31) If a similar analysis is carried out for the homodyne capacity case, one can easily 46 compute the normalized wideband capacity Ccoherent-homodyne(P'). For this case, the critical power is given by (1 crit P hom - (4.32) 2. 8 For P' < picntm, we have ('f 1m(inf' V Ccoherent-homodyne(P') In 2 where f' = 8P' + f' 1i, and fm Ccoherent-homodyne(P) where f m , < fc, and = > 212 fc. For P I(' (4.33) ki))(4 inin 2 Pr fnin)+ , In f 3 g frnin m- f)4) f' is given by 1 f 4P'______ = 2 (1 + fif) 2 + 4,' (4.35) (1 - fmin) Now, if we use these four boxed expressions to plot the normalized capacities coherent-homodyne(P') and Ccoherent-heterodyne(P'), as a function of P' = P/hfm. (see Fig. 4-2), we find that homodyne detection does better at lower input power levels whereas there is a value of power Pcross-over, beyond which heterodyne detection takes over. As explained in Section 4.1.2, this crossover occurs because of the different noise and bandwidth characteristics of heterodyne and homodyne detection. As expected though, by imposing these additional constraints by forcing lower and upper cutoff frequencies, the overall performance in either case deteriorates from the optimal homodyne (or heterodyne) performance given by C = (1/In 2) 4.2 2P/h (4.19),(4.20). Coherent State Capacity: Lossy Channel Let us transmit a coherent state Ia) (aI with a probability measure pA(a)d 2a, through a zero-temperature SM lossy channel (2.1) with transmissivity r7. It can be easily 47 2.006 ,, do fJ I -. ~.- I I * I- 410 1.00 'p - H cross-oer lei N 0.500 . - Homodyne -- P < P-crit-hom Homodyne -- P > P-crit-horn -Heterodyne -- P < P-crit-het Heterodyne P > P-crit-het)--- Cht /p peltt Het Hom Honr--Het 0.306 0.00 0.300 ILI I0 - (no bandwudth restriction). *I oe 1.200 - . 1.5ON P' (normalized input power) Figure 4-2: This figure illustrates the effect of forced lower and upper cut-off frequencies (fmin and fm), on the coherent state homodyne and heterodyne capacities. Given values of fmin and fm., the critical power levels (the input power at which the critical frequency f, hits the upper-cut off frequency fm, and beyond which power 'fills up' only between fmin and f m ) have been denoted by Ph" and Ph t respectively for the two cases. In this figure, we plot the normalized capacities C' = C/fmax vs. a normalized (dimensionless) input power P'=-P/hfm. We normalize all frequencies by f m . and denote them by f' =_ f/f ma. The 'dash-dotted' (black) line represents the capacity achievable using either homodyne or heterodyne with unrestricted use of bandwidth (Eqn. (4.20)), i.e. C = (1/ln2)V2P/h bits per second. The 'solid' (green) line, and the 'dashed' (blue) line represent homodyne and heterodyne capacities respectively for any set of cut-off frequencies satisfying fi = 0.2. For this value picrit of fin, P'ho = 0.16, and P'ht = 0.32. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). -ci 48 shown that when a coherent state = ca (aI is sent through this channel, the /IN output is also a coherent state whose mean is contracted towards the origin of the phase space by the factor /i, i.e. POUT =I V 177a) (Vja. If this output state is detected using homodyne (or heterodyne) measurement, we can use arguments similar to those in the previous section to show that the channel in this case can be represented by one (or a set of two parallel) AWGN channel(s); and hence capacity is achieved by a Gaussian distribution of coherent states at the input. 4.2.1 SM Capacity: Homodyne and Heterodyne Detection The SM capacities of the lossy channel with transmissivity 71 for coherent state encoding with homodyne and heterodyne detection respectively are given by Ccoherent-hornodyne(7, h) = Ccoherent-heterodyne (7, A) = log 2 (1+ 477A) log 2 (1 + I and (4.36) (4.37) which can be easily deduced from the results in [15] by observing that we can think of the above as the capacity of a lossless channel with an input 1,/ija) (.f/-aI and a power constraint Tr(a al qa)(Aaj) ; an. Note again that, at high values of h, 'coherent-heterodyne' does better than 'coherent-homodyne' whereas at small values of A, 'coherent-homodyne' does better. 4.2.2 Wideband Capacity Wideband capacities achievable using coherent state inputs can be calculated from the corresponding SM capacities using the technique illustrated in Section 2.6. In this section, we will consider the wideband capacity calculation for four cases - (a) Frequency independent loss, (b) Free-space channel in the far-field regime, (c) FreeSpace channel in the near-field regime, and (d) Single spatial mode propagation of 49 the Free-space channel - Ultra wideband. Frequency Independent Loss As is evident from comparing Eqn. (4.36) and Eqn. (4.37), the calculations for the case of homodyne detection will be very similar to that of heterodyne detection. So once again we shall go through the heterodyne detection case, and thereafter state the corresponding final results for the homodyne detection capacity. Let us assume that rI = r,Vj. Maximizing the overall information rate (4.38) ln(1 + qfi), C =b subject to total input power P (2.20), we get the optimum intensity (W/Hz) distribution as Pb Is - - = hf fi? = Io - hf- *,(4.39) where lo = (1/A In 2) and A is a Lagrange multiplier. Replacing the sums by integrals over all positive frequencies, and imposing Kuhn-Tucker like conditions, we can write the intensity variation as L I0 - Iif 0, if > 0 (4.40) otherwise with input power P given by fmax (4.41) P= JI(f)df. fmin The allocation of intensity across different frequencies in the transmission band can be looked upon as a 'water-filling' solution to an infinite set of classical parallel additive white Gaussian noise (AWGN) channels [2] (Fig. 4-3(a)). The above equations can be solved analytically case - exactly as we did for the lossless to obtain the wideband capacity in bits per second. The wideband capacities 50 h7 hf 1 )7 0 'io- P 7 fmin fmax 1o77 4 f fmin f max f 4 h (a) 10 fch (b) Figure 4-3: Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a channel with frequency independent transmissivity 7, using coherent-state inputs and heterodyne detection. (a) f m a > fe, (b) f m a < fc. for the lossy channel in the ideal case, when we calculate the capacities under the assumption that we have at our disposal photons, at all frequencies of transmission (i.e. no lower and upper cut-off frequencies), are given by Ccoherent-heterodyne Ccoherent-homodyne(17 P) = P) = 2njP 1 127h' 1 1n2 and 2 P h (4.42) (4.43) So, the performance of homodyne and heterodyne in this ideal case are the same. Next, let us impose a set of lower and upper cut-off frequencies given by fmin and fm. respectively. Carrying out our analysis along similar lines to the lossless case, and using the same normalization as in Section 4.1.4, the critical power level for heterodyne capacity is found to be ( For p' ; p'crt, we have 51 2 (4.44) Ccoherent -heterodyne (77, P') where f' = 2rP' + fmin - (4.45) in fmj., and f m , > fc. For P > P' , we get Ccoherent-heterodyne where f mr ( 2rP' I = P frn) - =q In 2 fc in + m in/ - (fmm ) (4.46) . < f, and f' is given by f',= 1(1 77P' + f'in) + (4.47) . If a similar analysis is carried out for the homodyne capacity case, one can easily compute the normalized wideband capacity CCoherent-homodyne(P'). For this case, the critical power is given by ) (1 - p/critf hor8 (4.48) For P' < P'/[cA, we have Ccoherent-homodyne(17, where f' = V8rP' + fmin, P') = 2T/Pl - 1 fmin I n ( '2 mm > pycrit and fmax fc. For P' > fr) Ccoherent-homodyne(77, P') where f mr In 2 i+ oP' n fc - w (4.49) n c e we get finn (/1) (4.50) < fc, and f' is given by f'C = 1(1 + fnin) + 4 7P' -f) (4.51) So all the wideband results pertaining to the case of frequency independent loss can be obtained from the corresponding expressions of the lossless case, by replacing 52 P with TIP. Also, imposing cut-off frequencies has the effect of twisting apart the homodyne and the heterodyne curves so that homodyne does better than heterodyne in the power-limited regime and vice versa in the bandwidth-limited (high-power) regime. Free-Space Channel: Far-Field Propagation The next case we consider is the free-space quantum optical channel (see Section 2.1), with an additional assumption that in the entire frequency range of transmission, the propagation remains well within the far-field regime, i.e. [fmin, fm,,]. DJ = q(f) < 1, Vf E In this case we can assume that at any given frequency in the above range, only one spatial mode of the field couples appreciable power into the receiver aperture with fractional power transfer being almost equal in magnitude to the freespace Fresnel number at that frequency D1 , and the rest of the spatial modes of the field can be neglected [5]. Let us denote by i, as we did in Section 2.6, a discrete frequency index, so that 2 the transmissivity of the ith frequency bin can be written as 7i = Afi , where A = ATAR/C 2 L 2 is a geometrical parameter of the channel. We define a constant K = h/A which has dimensions of power. On carrying out the analysis along similar lines as the frequency independent loss case, the optimum intensity distribution for coherent state encoding and heterodyne detection is found to be I(f) = { I0 - L' if > 0 ' 0, (4.52) otherwise where Io = (1/A In 2) and A is a Lagrange multiplier. In this case, we see that capacity is achieved with preferential use of high-frequency photons. So, it might seem that we can achieve as much capacity as we would like to, by using higher and higher frequencies. But this is not true, because the highest frequency of our transmission band must adhere to the far-field condition for the single-spatial-mode assumption to remain valid. Eqn. (4.52) suggests that we normalize all the quantities with respect to fmna, the maximum allowed frequency of our transmission band. Let us define 53 k 1 f f f' 11 Figure 4-4: Optimal allocation of intensity across the transmission bandwidth to achieve maximum capacity on a channel with frequency dependent transmissivity r7 cc f2 , using coherent-state inputs and heterodyne detection. I' -_ - - I(f)fma I'(f') where f' = f/fma, mT 0, K p if > 0 -w(4.53) otherwise and 10' = Iofmax/K. Let us define a dimensionless (normalized) power P' = P/K, where P is the total average input power in Watts, and a normalized capacity C' = C/f m a, where C is the wideband capacity of the channel in bits/sec. It is now straightforward to show that C'= P'= where f'in = In 2 ln(1 + I'(f')f')df', and (4.54) "i I'(f')df', /1|in (4.55) fmin/fmax. Analogous to the channel with frequency-independent loss, we can look upon the above power-allocation as similar to the classical idea of 'water-filling' for parallel AWGN channels. Interestingly, in this case, the 'water' fills up starting from the 54 highest allowed frequency towards lower frequencies. Depending upon the value of f',5 P' has a critical value Pc'rit, after which the 'water' fills up vertically in the region defined by f' E [fVrin, 11. Carrying out the integrations above in the low-power regime, when P' < Pcrit, i.e. 1/6 > f, we find that the normalized capacity C' is given by C'coherent--heterodyne(P') where = In 16 In 2 (I01 - 1+ (4.56) , 16 is a dimensionless parameter, which is obtained from the power constraint equation Eqn. (2.20), that takes the following form in this case: P' = 1/ - 1 - nI6. In the high-power regime, when P' > Pcrit 1 i.e. 1 C'coherent-heterodyne(PI) P' = I;(1 < (4.57) fin, the corresponding ( n IV - 1 + f'in - fa) + - frin equations IOf'in)), and(4.58) (4.59) ln(fl'i). In Fig. 4-5, the normalized capacity C' = C/fm, has been plotted vs. normalized power P' = P/K for coherent-state inputs and heterodyne detection, for different values of frin. We find that the performance degrades steadily, at high powers, as f min is increased from zero, because of the increasingly stringent bandwidth constraint. The interesting thing to note in this case is that the far-field capacity is directly proportional to the maximum frequency of the transmission band, fm,. This should not be interpreted as a possibility to have infinite capacity by increasing fm . without bound, because the single-spatial-mode approximation requires that even the highest frequency photon be sufficiently deep into the far-field (Dfmax < 1). The calculations for the case with coherent-state inputs and homodyne detection can be performed along similar lines. In Fig. 4-6, we plot C' vs. P' for both homodyne and heterodyne 55 65-4- U 2 1 0 0 20 40 60 Normalized Power P' 80 100 Figure 4-5: Dependence of C' on P' for coherent-state inputs and heterodyne detection, for different values of fai.. The blue (solid) line is for fjni = 0, the red (dotted) line is for flnin = 0.1 and the green (dashed) line is for fani = 0.2. Observe that all the three curves coincide in the low power regime, because when P' < PC'rit, C' does not depend on f.in. When P' > Pcrit, performance of the channel deteriorates because of restricting ourselves to a non-zero fn.in. 56 2- 1l.5 1 N0. 0 0 0 2 6 4 Normalized Power, P' 8 Figure 4-6: Dependence of C' on P' for coherent-state inputs with heterodyne and = 0. The blue (solid) line is for heterodyne detection, homodyne detection, for f and the red (dashed) line is for homodyne detection. detection with f'ai. = 0. We see, as before, that homodyne detection performs better in the low-power (power-limited) regime, because the lower noise-variance associated with homodyne detection overwhelms the bandwidth advantage of heterodyne detection. But, in the high-power (bandwidth-limited) regime, the bandwidth advantage of heterodyne detection makes it outperform homodyne detection. Free-Space Channel: Near-Field Propagation Let us again consider the free-space optical channel that we did in the previous subsection, but now with the assumption that the entire frequency range of transmission remains well within the near-field propagation regime, i.e. D > 1, Vf E [fm"i, fmaxj. It is well known that [5] in this regime at any given frequency of transmission f, there are approximately Df spatial modes of the bosonic field, each one coupling power perfectly (without loss, q = 1) to the receiver aperture. The remaining spatial modes have insignificant contribution to the overall power transfer, and hence may be neglected. The wideband capacity calculation for this case can be set up as follows. 57 Let us assume that the frequency axis in the region f E [f m in, f m ax] into small bins of width b Hz, and denote the center frequency of the is divided up jth bin by f. A = ATAR/C 2 L 2 is again a geometrical parameter of the channel (Fig. 2-1). The free-space Fresnel number for the frequency fi is given by Df, are Df, independent spatial modes of the field at frequency fi, = Afi 2 . So, there each of which couple power perfectly into the receiver aperture. Without loss of generality, assume that each spatial mode at the frequency f, is excited in a state with the same mean photon number denoted by iii. So, the heterodyne detection capacity C, of this wideband channel is given by C = max b (Afi2)log 2 (1+ i) , (4.60) subject to the maximum average input power constraint P, which now takes the form P= (4.61) (Af2 )hfib. Solving this constrained maximization problem along similar lines as the SM lossless case, we obtain the optimum intensity (W/Hz) distribution P-= (Afi 2 )hfi* = Afi 2 (Jo - hfi), (4.62) b where Io = (1/A in 2) and A is a Lagrange multiplier. Replacing the sums by integrals, and imposing Kuhn-Tucker like conditions as before, we can write the optimal intensity variation as - hf), if > 0 Af2(I 0, (4.63) otherwise with input power P given by P J I(f)df. fmin 58 (4.64) For a given value of Io, due to the lower cut-off frequency constraint, and the opti- f E [fmin, fc], fc, where f, = 10 /h. mal intensity expression, non-zero power is allocated only in the region as long as the upper frequency cut-off fmax is greater than Integrating the optimal intensity I(f) in the region P= fc=Io/h In Af 2 (Io j fai Ah - hf)df = 12 12 f E [fm in, LI, (fe4 - 4fe f m in3 From the above equation, we see that as expected, substitute f, = f m in fc we get, + 3fmin 4 ). = fmin, (4.65) for P = 0. Let us + g(fmin, P), where it can be easily shown that g(fmin, P) is a function satisfying 2 2 g(fmin, P)4 + (4fmin)g(fmin, P) 3 + (6fmin )g(fmin, P) = 12P Ah (4.66) It is easy to observe that the expression on the left hand side is a strictly increasing function of g(fnin, P) for fm in > 0, in the region g(fmin, P) > 0. Also, g(fnin, 0) = 0, and g(fmin, P) increases as P increases. So, given a value of fmin, and P > 0, there exists a unique value of g(fmin, P), which corresponds to a unique value of As the input power P is increased from zero upwards, it hits the upper cut-off frequency f m ,. fc f. increases from fmin, until The value of P for which this happens will be termed the 'critical power' for heterodyne capacity, which we denote as P .r~t To simplify the expressions in the calculations that follow, let us denote by P' P/(Ahfm), a normalized (and dimensionless) input power. We define a normalized and normalize all frequencies by the upper cut-off frequency capacity C' - C/Afm, fmax, i.e. f' = f/fmax. Note that the normalization we use for C and P here, Also, note that with the above is different from what we used in Section 4.1.4. normalization, Eqn. (4.65) and Eqn. (4.66) take the following form: p,= 12P' l(' 12 = = g(fnin, - 4f'f m' _ P') i 3 + 3f' i 4), ')+(4fin)g'(fmnin, 59 (4.67) PF ')+(6fin2 in,2,(.) where f' = f.'i + g'(f.'in, P'). So, following an argument similar to the one we used f 'in before, we can prove that for a given set of values of and P' > 0, there exists a unique value of f'. It is easy to deduce from Eqn. (4.65) that p/crit Pht= (1 - fi'in3 ) 3 When P' is increased beyond this value, power only 'fills up' in the region f _ (1 - f, keeps f(469) increasing beyond fma, but the E [fmin, fma]. The above equations can be solved analytically to compute the normalized wideband capacity P' = fc/fma, For < fc, and (4.70) we get Coherent-heterodyne(P f ma 1 + 3 In f/ is obtained for a given value of P' as explained above, and fma > fc. For P' > Pt' where "3 91 2 Ccoherent-heterodyne (P') f' CCoherent-heterodyne(P'). we have h P', where (.9 4 f' = 912 91n2 "3 ) + 3lnf' - 3ffni 3 ln in (4.71) fMin. is given by 12P' + 3(1 - fin 4(1- fm(in.7 (4.72) If a similar analysis is carried out for the homodyne capacity case, one can easily compute the normalized wideband capacity CCoherent-homodyne(P'). For this case, the normalized critical power is given by crit Fhor' For F' < p',,i _ (1- 3 fin (1 - fiin 4) 16 12 we have 60 (4.73) Ccoherent-homodyne(P') where f m , 1 = 3 3 + 3I fl) (474 > fe, and f' is calculated following a similar procedure to that of the heterodyne case. The relationship between normalized input power P' and f' in this case is given by 1 (f P 48 C - 4f 3fina + 3fm'i4). (4.75) In (4.76) For P' > p'crit we get Ccoherent-homodyne(P) 1 3 [) -2 where fma.a < fc, and f' is given by = 48P' + 3(1 4(1 - f' ) 3 ) Now, we use all the above four boxed expressions for to plot the (normalized) capacities Ccoherent-homodyne(P') and Ccoherent-heterodyne(P), as functions of P' (see Fig. 4-7). We find that, similar to the lossless case, in the wideband free-space channel in the near-field regime, homodyne detection does better at lower input power levels and there is a value of power Pcross-over, beyond which heterodyne takes over as having higher capacity. Free-Space Channel: Single Spatial Mode - Ultra Wideband In the previous two sections, we calculated the coherent state capacities of the freespace channel, first assuming that the entire transmission bandwidth is in the farfield (Df < 1) propagation regime, and next assuming that the entire transmission bandwidth is in the near-field (D > 1) regime. For both these cases, the expressions for wideband capacity achievable using coherent state encoding with either homodyne or heterodyne detection could be calculated explicitly. What happens if there is no 61 IsI 0.400 ' - ~0.300 .200 Homodyne -- P < P-crit-hom Homodyne P > P-rit-hom Heterodyne 0.000 -- P > P-crit-het I 0.000 0.100 0.300 0.200 0.500 0.400 (normalized input power) 0.600 P ma= Ahf~ Figure 4-7: This figure illustrates the coherent state homodyne and heterodyne capacities for the free-space optical channel in the near-field regime. All the frequencies in the transmission range f E [fmin, fmaxl satisfy Df > 1. In this curve, a normalized capacity C' = C/Afm. has been plotted against a dimensionless normalized input 2 2 power P' = P/(Ahf,), where A = ATAR/C L is a geometrical parameter of the channel (Fig. 2-1). Given values of f min and fm a, the critical power levels (the input power at which the critical frequency f, hits the upper-cut off frequency f ma, and beyond which power 'fills up' only between fmin and fma) have been denoted het respectively for the two cases. The 'solid' (black and blue) line, b pithond m~ and pc and the 'dashed' (red and green) line represent homodyne and heterodyne capacities respectively for any set of cut-off frequencies satisfying f' i = 0.3. For this value = 0.019, and of f'in, the values of the normalized critical input power are: p, h't = 0.076 respectively. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). 62 restriction on the bandwidth of communication over the input power limited freespace channel? For example, consider an ultra wideband satellite to satellite freespace communication link with transmitter and receiver apertures of areas AT ~~400 cm 2 and AR % 1 In 2 , separated by L = 400 km of free-space. Assume that the frequency of transmission range from that associated with a wavelength of a few tens of microns (which would typically be deep into the near-field regime - Df ~~0.001), to that associated with a wavelength of several hundred nanometers (which could be considered well into the near field regime Df ~ 25). In such a case, any capacity calculation would have to take into consideration the actual variation of modal transmissivity with frequency of transmission over a broad range of frequencies (see Fig. 2-3). In Section 2.1.2, we observed that as the value of Df increases from the near-field regime, more and more spatial modes of the field start coupling appreciable amount of power into the receiver. If we arrange the fractional power transfers of the spatial modes in decreasing order and denote the fractional power transfer (transmissivity) of the ith mode by 77i, then - from the eigenvalue behavior of the classical free-space modal decomposition theory [23, 24] -- the rhi's depend upon the carrier frequency f as shown in Fig. 2-3. To calculate the general capacity of the ultra wideband free-space channel, we must sum the information capacities of all the contributing spatial modes at a frequency, taking into account the actual values of modal transmissivity for each one of them. As there is no good analytical expression for their functional dependence on frequency of transmission, it is seemingly very difficult to come up with an explicit formula for capacity. So, what we shall do is to find the capacity of the channel assuming that only a single spatial mode propagates at all frequencies. As capacity is additive, an analysis very similar to this can be used in conjunction with the relative magnitudes of transmissivity for different spatial modes to compute the true wideband capacity. In this section, we will calculate the capacity under the assumption of single spatial mode transmission. For our analysis, we will assume that we have photons at our disposal at all frequencies, i.e. for any f E (0, oo). We justify later why this assumption is 63 reasonable. Setting up the calculation in the usual manner, we have to maximize the sum of SM capacities C(r(f), A) across the entire frequency band of interest. For heterodyne detection capacity, we have to maximize C = max C(q(f,), hi) = max b log 2 (1 + 1(fi)i)] , (4.78) subject to the maximum average input power constraint P= hfibi. (4.79) Because of the nonlinear frequency dependence of transmissivity, we have to do this maximization numerically (using the Lagrange multiplier method). The point to note here is the following. At very low frequencies this channel looks like the far-field channel we analyzed earlier, in which q oc f 2 . So in that region, we might expect that the optimal intensity allocation uses high frequency photons preferentially, and that the intensity goes to zero at low frequencies. At higher frequencies, this channel is closer to the lossless wideband channel we considered earlier, for which we know that, the optimal intensity allocation goes to zero at very high frequencies. So, in the ultra wideband case, we would expect the intensity allocation to vanish both for very low and very high frequencies. This is why our assumption of f E (0, oc) is reasonable. As an example of our calculation, we illustrate the results of one of our numerical evaluations for the ultra wideband capacities of homodyne and heterodyne detection, using the set of geometrical parameters listed in the first paragraph of this sub-section (Fig. 4-8). In this case, the frequency range of transmission that has been used, is f E [0 Hz, 3000 THz], corresponding to wavelengths in the range [300 nm, cc). General Loss Model: A Wideband Analysis Let us now analyze the wideband capacity calculation for the coherent state channel with homodyne and heterodyne detection for a general frequency dependent loss. It 64 .-7-0 00 1 -1 P (input power in Watts) 161 Figure 4-8: An example of ultra wideband capacities for coherent state encoding with homodyne or heterodyne detection, for the free-space optical channel under the assumption of a single spatial mode transmission. can be shown' that, for any frequency dependent transmissivity 77(f), Ccoherent-homodyne coherent-heterodyne(4 (- To compute the general wideband capacity using coherent states and heterodyne detection, we have to perform the constrained maximization (4.78). The optimum frequency allocation of intensity I(f) is normalized by the Planck's constant to obtain I where h 1' - 9(A) if > 0 0, otherwise A)) and g(f) = n' is a Lagrange multiplier, f/(f) (-1 is a function that is no less than f, Vf E (0, oo). Let us denote the total input power normalized by the Planck's 'This can be shown by comparing the two constrained maximization problems and realizing that by suitably scaling the capacity and the input power, one maximization problem can be represented in terms of the other. This simple idea was proved by B.J. Yen, and S. Guha, Research Laboratory of Electronics, MIT 65 constant by P' = P/h. lo' is obtained from the power constraint equation: P' = if (Io' - g(f)) df, EA(P') (4.82) where A(P') is a subset of (0, oc) in which the integrand is non-negative. This power allocation can again be looked upon as a classical 'water-filling' solution, where the power P' 'fills' up into the g(f)-curve, starting at the global minimum of g(f). Increasing the value of P'increases the value of Io'. The optimal power allocation for the rl transmissivity shown in Fig. 2-3, has been plotted in Fig. 4-9. The capacity C is found to be: Ccoherent-heterodyne (P') - In 2 InnI 2 + [n(P/ n(A(P')) k A(P') g(f)df) JA (P') Ing(f)df - ni(A(P'))] /J (4.83) where p(A(P')) is the measure of the set A(P') in Hz. For the lossless case, g(f) = f and A(P') = [0, V2P']. Substituting this A(P') into Eqn. (4.83), one readily obtains the lossless heterodyne capacity C = v2P'/ln 2 (4.19). This analysis suggests an interesting way to look at the heterodyne (or homodyne) detection capacity and an obvious numerical algorithmic approach to evaluate the wideband capacity for an arbitrary frequency dependent transmissivity. 66 0P - g~gf))-/- A(P') frequency (f) Figure 4-9: The optimal power allocation for the wideband coherent state channel with heterodyne detection depicted as 'water-filling', for frequency dependent transmissivity shown for r/1 in Fig. 2-3. 67 Chapter 5 Capacity using Squeezed State Inputs and Homodyne Detection We saw in the previous chapter that coherent states are minimum uncertainty product states, whose quadrature components have equal variances. A coherent state 1a) is sometimes represented on the phase plane by a quantum 'error-circle', representing the vacuum noise, whose center is at the tip of the vector representing the mean complex amplitude a, and radius being equal to the quadrature uncertainty ((A& 1 )2)1/ 2 1/2. Squeezed states are also quadrature minimum uncertainty product states, but their quadrature components have unequal variances. To define quadrature-squeezed states, we need to introduce the squeeze operator [15] S(r, #) = exp (2 - where r is called the 'squeezing parameter' and -t5ei) 4 determines , (5.1) the phase of the squeez- ing. The squeeze operator transforms the modal annihilation operator according to S(r, q)d [S(r, 0) t = d cosh r + &te2 io sinh r. (5.2) A quadrature-squeezed state Ia)(,,) = D(6, a)S(r, #) 0) 68 (5.3) is obtained by 'squeezing' the vacuum state and then displacing it. D(d, a) = As choosing 0 amounts to a rota- exp(aet - a*&) is the displacement operator. tion in the phase-plane, we can always arrange to set q = 0, which we do henceforth. The mean complex amplitude of the squeezed state Qa(r,,) is a, similar to a coherent state. The quadrature components are uncorrelated and have variances ((,A&)2) _l -2r, (5.4) _e2r. (5.5) 4 ((A=2)2) 4 These states thus constitute the entire class of quadrature minimum uncertainty product states. If r > 0, the first quadrature will have a variance reduced below the vacuum level, whereas the second will have a variance increased above the vacuum level. We usually recognize et as the 'squeezed quadrature' and &2as the 'amplified quadrature'. A squeezed state can be represented on the phase plane much in the same way as a coherent state, except that the 'error-circle' is replaced by an 'error-ellipse'. 5.1 Squeezed State Capacity: Lossless Channel Let us transmit a squeezed state Pa, = Ial)(r,O)(r,O)(Ci Iwith a probability measure PA1 (al)dal, at each use of the SM channel, where a, is real and r > 0. The uncon- ditioned channel density operator per use, is given by P= JPA (a)a daj, (5.6) and the mean photon number constraint takes the form i! = Tr(&f&) = where 69 .2 + sinh 2 (r), (5.7) 01 = J a2PA(ai)dai (5.8) is the second moment of the distribution PA, (a,). The important thing to note here is that the sinh 2 (r) contribution to n represents the excess excitation of the amplified quadrature, which through the power constraint, limits the degree of squeezing. SM Capacity 5.1.1 As we know [5], ideal homodyne detection measures a quadrature component, and hence for measurement along the di quadrature is described by the POVM n2x where { = lx1)(Xil, (5.9) xi) } are the eigenstates of &1. The conditional probability density to read x1 at the output given that a1 was transmitted, is given by PX1IA 1 (x 1 Iai) - (5.10) 1 2t) Tr(, (5.11) I(Xil l)(r,O)12 1 27r( exp e-2r) (x1 -ai)2~ - X. 2(e-2r) (5.12) Similar to the case of coherent state channel, the channel noise is additive and Gaussian. So, the mutual information is maximized by a Gaussian input probability density a12) e PA1 (a,) = /2r 2 exp 2/.2 (5.13) , and the SM capacity (in bits per use) is given by C(fI, r) = log 2 (+ _ = log 2 [1 + 4e 2 r( 70 - sinh 2 r)]. (5.14) A further maximization with respect to the squeeze parameter r yields the SM capacity per use of the squeezed state channel: Csqueezed-homodyne 5.1.2 (h) = 1og 2 (5.15) (1 + 2h). Wideband Capacity As is evident from the expression of the SM capacity of the lossless squeezed state channel, the wideband capacity calculation closely parallels that of the coherent state heterodyne detection channel. The wideband capacity (in bits per second) in the case when we assume that there is no restriction on the available bandwidth, is given by Csqueezed-homodyne(P) (5.16) h = where P is the total input power constraint at the input. It can again be similarly shown that if we impose additional restraint on the available bandwidth by imposing a pair of lower and upper cut-off frequencies fmin and fma, the wideband capacity deteriorates from the ideal case. Carrying out our analysis along similar lines to the coherent state channel, the normalized critical power level for squeezed state coherent detection capacity is found to be p crit Ici= _ pcrit sq hfm2a sq 1- fI _fmi 4 (5.17) )2 (-7 For P' <pcrit , we have - C queezedhomodyne where f' = V47M - f(P/inin (5.18) + f'in , and fma > fc. For P' > pcrit we get Csqueezed-homodyne(PI) 1 [(1 - fi)+ 71 Infc - fmin ln (5.19) where f m ax < fe, and f' is given f' by 1 = - (1 + fmi) + 2 -(1 2P' ,.(.0 - fmin) The above expressions have been plotted in Fig. 5-1, in their respective regions ' of validity for an arbitrary pair of cut-off frequencies satisfying = 0.2. Squeezed state encoding with homodyne detection consistently performs better than coherent state encoding and homodyne detection, which is expected because whatever performance is achieved by a coherent state encoding can also be achieved using a squeezed state encoding (using squeeze parameter r = 0 all the time). The figure shows comparative performance of squeezed state encoding and homodyne detection with respect to that of coherent state encoding with homodyne and heterodyne detection. 5.2 Squeezed State Capacity: Lossy Channel Let us transmit a squeezed state &, = Ial)(r,o)(r,o)(ail with an apriori probability measure PA, (al)dai, at each use of the zero-temperature SM lossy channel of transmissivity 17, where a 1 is real and r > 0. There is a class of quantum states known as 'Gaussian States', that can be characterized solely by specifying their first and second moments, (d), (dt d), and (z5 2 ), because their Wigner characteristic function is of Gaussian form. All squeezed states are Gaussian states. It can be shown that if the input modes & and b in the SM lossy channel ( = V77e + '1 - 17b) are excited in Gaussian states, then the output mode e is also in a Gaussian state. The mean and the variance of the first quadrature of the output state are respectively given by (&1) = (1 (5.21) and \Vce, - + (1 - 1), (5.22) where the second term in the variance is the (T = 0) vacuum noise contributed by the b mode. 72 2.000 - Coe - ~ ~ Li---- ( /-Coherent-Hom - Coherent-Hor ---P < P-crit-hom C.- ---P<> P-crit-hom - Coherent-Het ---P < P-cnt-het Coherent-Het ---P >P-crit-het Squeezed-Hom ---P < P-crit-sq Squeezed-Horn --- P >P-crit-sq Squeezed-Hom (no bandwidth restriction) Coherent-[Horn 6.010 1.300 = Hed] (no bandwidth restriction)-0.900 0.608 --- 1.200 - 1.50 P (normaIized input power) Figure 5-1: This figure shows comparative performance of squeezed state encoding and homodyne detection with respect to that of coherent state encoding with homodyne and heterodyne detection. The 'solid' lines (red, green and grey) represent coherent state heterodyne, coherent state homodyne, and squeezed state homodyne capacities respectively for an arbitrary set of cut-off frequencies satisfying fmj. = 0.2. Different capacity expressions hold true for power levels below and above the respective critical powers (Shown by the thick and the thin lines in the figure respectively). For the above = = 0.32, P't value of Ain, the values of the normalized cut-off power are: P't 0.08, and P' = 0.16. The thick dashed (blue) line is the best achievable performance of the coherent state channel (with either heterodyne or homodyne detection), and is given by C' = (1/ In 2)v/2-. The thick dotted (red) line is the best achievable 4 . performance of the squeezed state channel, and is given by C' = (1/ ln 2)vri 73 5.2.1 SM Capacity The conditional probability density to read x, at the output given that a 1 was transmitted, is given by pXiIAi(XlIOZ1) (5.23) Tr(/S1iXXi)I = exp rj) ([-gg-2 27 i -. 2(1 [1 -,q+ (- (5.24) (7-ae2 7e-2r Because the channel noise is additive and Gaussian, the mutual information is maximized for a Gaussian input density (5.13), and the SM capacity is given by C(7, i, r) = -log 2 2 1+ 1 1-77+ e-2r 4 (5.25) .sinh2r r7 A further maximization with respect to the squeezing parameter r yields the SM capacity (in bits per use): Csqueezed-homodyne(77, ii) = -log 2 2 + 2 1+ (1 (x(77, h) + x(77, f)) + X(, h) 1 , (5.26) where the function x(ij, h) is given by X(, 5.2.2 )= 7 + 2h)12 -1 (I+ .(5.27) Wideband Capacity It is evident from the SM formula for the capacity of the squeezed state channel that calculation of the wideband capacity analytically is very difficult. So, to compute the wideband capacity of the squeezed state channel, we perform the constrained maximization problem numerically using the Lagrange multiplier method. We illustrate one example of our numerical calculation of squeezed state capacity 74 l17Ul 0 ~1 U Ut .0 N N U .0 U Coherent-Homodyne Cohrent-Hetrodyne-- Squeezed-Homodyne - - - - - I / Q 1I 5 i . . . . . .. . . . . . . . .. i 10-2 1-4 134 10 P (Input power in Watts) Figure 5-2: An example of ultra wideband capacities for squeezed state channel and coherent state encoding with homodyne or heterodyne detection, for the free-space optical channel under the assumption of a single spatial mode transmission. to show how it typically compares with the coherent state capacities. This example has been done for the ultra wideband single-spatial-mode capacity for the free-space channel parameters we used in Fig. 4-8. As expected, squeezed state capacity is consistently higher than the coherent state homodyne detection capacity, and heterodyne detection eventually outperforms both the homodyne detection schemes in the bandwidth-limited (high-power) regime. 75 Chapter 6 The Ultimate Classical Capacity of the Lossy Channel In the previous chapters, we calculated and compared classical information capacities of the single mode and wideband lossy bosonic channel, using various combinations of encoding schemes and receiver structures. In this chapter, we discuss very recent progress made by collaborative efforts of several people in, and associated with the research group of Prof. Jeffrey H. Shapiro, at RLE, MIT [18], in the direction of finding the ultimate information capacity (2.17) of the lossy bosonic channel. We will perform the calculation of ultimate wideband capacity for the free-space channel in the far-field, as an example. We also show that if certain conditions are satisfied, one can achieve the ultimate wideband capacity of a lossy channel asymptotically in the high power regime, using coherent states and heterodyne detection. We stated earlier that the ultimate classical capacity of a quantum channel (2.17) is the maximum of the 'Holevo Information' at the output of the channel, maximized over all possible sets of input states and apriori input probability distributions. Also, we mentioned that the ultimate classical capacities of the SM and the wideband lossless channels are known [15], and that a number-state encoding along with a direct detection receiver can be used to achieve these capacities (3.3). 76 6.1 Ultimate Capacity: SM and Wideband Lossy Channels Recently, it was shown that the ultimate classical capacity can be explicitly calculated for the T = 0 SM and wideband lossy bosonic channels [18]. The SM capacity of a lossy channel with transmissivity 71 is given by Cuitimate(7, ii) = g("i), (6.1) where g(.) is as defined in (3.2). To evaluate the wideband capacity for a specific case, one has to perform the maximization C = max where mi is g(lifii), (6.2) the transmissivity of the ith mode, and the maximum is evaluated subject to the usual average input power constraint (2.20). It can also be argued that this capacity can be achieved using a single use of the channel by using a random code on coherent states, factored over frequency modes [18]. This means that for this channel neither non-classical states for encoding, nor entangled codewords (either entangled over successive channel uses, or entangled over modes) are necessary to achieve capacity. It might be possible, however, that one could achieve capacity using quantum encodings, and that such encodings might have lower error probabilities for finite length block codes than those of the capacity-achieving coherent state encoding. The wideband capacity for a frequency independent modal transmissivity q, is given by c. Cultimate -1 1n 2 2nP(63 3h. (6.3) Now that we know the lossy channel capacities for so many cases, including the ultimate classical capacity, it is a good time to put them all all together in one 77 3210 Coherent-..Homodyne Coherent--Heterodyne 2DO -Squeezed--- Homodyne Ultimate Capacity Number-State.--Direct-Detection --- - - - - . iaoo 1200 Q 0.600 0000 3000 6000 9000 12000 150O0 180O Input average photon number constraint (nbar) Figure 6-1: Single Mode Capacities (in bits per use of the channel) for various cases. The transmissivity of the SM channel q = 0.16. place and compare. In Fig. 6-1, we plot all the SM capacities of the lossy bosonic channel with transmissivity 7 = 0.16. All the plots are evaluations of the respective analytic SM capacity expressions as functions of the average input photon number A, except the plot corresponding to the case of number-state and direct detection, which has been obtained by using a numerical Blahut-Arimito procedure. We notice that as the transmissivity is so low, homodyne detection with optimized squeezed states performs only marginally better than coherent states and homodyne detection, though eventually at high photon numbers, coherent states and heterodyne detection outperform both of them. The solid (green) line in the figure corresponds to the SM ultimate capacity of channel, given by g(0.16A). 6.2 Free-Space Channel: Far-Field Propagation In this section, we will consider the calculation of wideband capacity of the free-space optical communication channel in the far-field regime. Interestingly it turns out that 78 exactly the same normalization of capacity and power used in the far-field analysis of the coherent state channel can be used for the ultimate capacity. Starting our wideband analysis with a frequency axis divided up into small bins of 2 size b Hz each, we denote the transmissivity of the ith frequency bin as 77i = Afi , where ATAR/C 2L 2 is a geometrical parameter of the channel. A K = We define a constant h/A which has dimensions of power. On carrying out the analysis along similar lines as the coherent state heterodyne detection case, the optimal number allocation is found to be n.* 2 =Afi 64 (6.4) (exp (A In 2)K) Aff2 where A is the Lagrange multiplier. The superscript (*) signifies that this is the optimum distribution of mean photon numbers across all frequencies. Also, the optimal allocation of intensity (energy per mode) across the frequencies is given by i = hfish* = (. fiexp (6.5) -A In2) Let us define a new parameter lo = (1/A In 2) which we will treat as a Lagrange multiplier. If we now let our frequency bin width b become very small, so that we can treat the frequency f as a continuous parameter, the optimal mean number distribution W*(f), and the optimal intensity allocation 1(f) are respectively given by: = n(f) exp(_K_) expk " , - if > 0 (6.6) otherwise 0, and K I(f) = , 0, Note that 1(f) > 0 +-l if > 0 (6.7) KIf-1 otherwise Io > 0. Also note that I(f) -- + o, as 79 f -+ oc. So, to keep this far-field analysis valid, we must impose an upper frequency cutoff fmax that satisfies the far-field condition Dmax < 1. Eqn. (6.7) suggests that we normalize all the quantities with respect to fmax, the maximum allowed frequency of our transmission band. Let us define I~f~fmx IYf where f' = f/fmax, and Io' = emoi'_1 e /07 Kf)fa K (0, _ ,if > 06. otherwise 68 Iofmax/K. Let us define a dimensionless (normalized) power P' = P/K, where P is the total average input power in Watts, and a normalized capacity C' = C/f m a, where C is the wideband capacity of the channel in bits/sec. It is now straightforward to show that the normalized Capacity C' is given by In + C'ultimate(P') = ln ,o, df1 , (6.9) where the parameter 10' is computed from the power constraint equation = J elIo 4'' - df' (6.10) The normalized capacity C' = C/fmax has been plotted vs. the normalized power P' = P/K, along with the rates achievable using coherent-state encoding with heterodyne and homodyne detection receivers, in Fig. 6-2. It is observed that the classical capacity of the far-field optical channel is proportional to the maximum frequency of transmission fmax. Again, as in the coherent-state channel, this does not imply a pos- sibility of infinite capacity because this result depends on the single-spatial-mode approximation and the quadratic frequency dependence of modal transmissivity, which requires that the highest frequency photon resides sufficiently deep into the far-field (Dm. < 1). 80 10- ............. ................ ... 8M E z0 00 250 500 750 1000 Normalized Power, P' Figure 6-2: Dependence of C' on P' for the ultimate classical capacity compared to C' for coherent-state inputs with heterodyne and homodyne detection. All curves assume I = 0. The blue (solid) line is the ultimate classical capacity, the red (dotted) line is for heterodyne detection, and the green (dashed) line is for homodyne detection. 6.3 Asymptotic Optimality of Heterodyne Detection Interestingly, it is observed in some of the cases we have considered so far, that the performance of coherent-state encoding with heterodyne detection asymptotically approaches the optimal capacity in the limit of large input power P. In this context, we will first consider the free-space channel in the far-field regime and show analytically that one can achieve ultimate capacity for this channel using coherent states and heterodyne detection in the limit of high input power. After that, we will demonstrate a sufficient condition for this to hold true for any general wideband bosonic channel. 81 6.3.1 Heterodyne Detection Optimality: Free-Space Far-Field Propagation Using the normalization from the previous section, it can be shown that the normalized capacity for coherent state encoding and heterodyne detection, C' is given by C'= in2 inIs -1 + '01 (6.11) , where I is a dimensionless parameter obtained from the power constraint equation, that takes the following form: P' = IO/- 1 - In I/. (6.12) Now, let us compare the performance of the coherent state heterodyne detection case with the ultimate capacity, in the high power regime. For simplicity, we will only consider the case of fmi = 0. In the ultimate capacity case, from Eqn. (6.10), we infer that high values of P' will be obtained by high values of the parameter 1o'. For 10' P'/ - (o'f') df'. fo P - oc, (6.13) So, for large values of Io', Io' ~ P'. We see from Eqn. (6.12) that the same holds for coherent state - heterodyne detection case as well (as in Io' and 1 are negligible in comparison to Io' for large 1o'. So, for large values of normalized input power P', 82 C'ultimate (P) of+ C'coherent-heterodyne (P) f In in I0 - 1 + df 1 + f ln (Io'f') df' In Io' - 11111j'-i(6.15) 1 + (In Io' - 1) In lo' - 1 ln(P') ln(P') - 1 1. .4) (-5 (6.16) (6.18) From the above, and the definitions of normalized capacity and normalized power, we conclude that as P -- 00, ln(P/K) ln(P/K) - 1 Cuitimate(P) Ccoherent -heterodyne(P) - 1. (6.19) So, unlike the case of the lossless channel and the case of frequency independent transmissivity, in which cases the above ratio is found to be ir/v/5 independent of the input power P, we see that in the case of the far-field free-space channel coherent state encoding and heterodyne detection is indeed asymptotically optimal, in the limit of large input power. 6.3.2 A Sufficient Condition for Asymptotic Optimality of Coherent State Encoding and Heterodyne Detection We saw that in the case of frequency independent loss the capacity achieved by coherent state encoding and heterodyne detection does not approach the ultimate capacity in the high power limit, whereas it does in the case of the far-field free-space channel. One important difference in these two cases is that the optimal intensity distribution across frequencies I(f), goes to zero at high frequencies for the frequency independent loss, whereas 1(f) -+ Io as f -- oc for the far-field free-space channel. Because of this behavior of the optimal intensity distribution for the free-space channel, we have to 83 impose an upper cut-off frequency fma, in our capacity calculation (In a real optical communication system, the upper cut-off frequency will be decided by the available hardware and other resources). As the total input power P -+ 00, the optimum mean photon number at every frequency W*(f, P) -* oc. Note that one can obtain *(f, P) by substituting the Lagrange multiplier A in the expression for h*(f,A) in terms of the input power P by solving the power constraint equation. The introduction of the upper cut-off frequency causes the above convergence to become 'uniform'. In the absence of an upper cut-off frequency, however high the power level P might be, there wouldn't be any single N, such that h*(f, P) > N, Vf E (0, oo), which is the condition of uniform convergence. It can be rigorously shown that, for any arbitrary frequency dependence of modal transmissivity, a uniform convergence condition can be set up as a sufficient condition for the asymptotic optimality of heterodyne detection. Consider the following lemma [19]: Lemma 1 (SGY) For wideband quantum communication with an arbitraryfrequency dependence of modal transmissivity r(f), and a given frequency band for communica- tion F (fmin, fm.), if the ratio of optimum mean photon numbers for coherent-state heterodyne detection case hi*e(f, P) to that of ultimate capacity case h*lt(f, P), uniformly converges to unity as P as P -> -- oc, and *l( f, P) uniformly converges to infinity oc, then coherent state encoding with heterodyne detection asymptotically approaches optimum performance in the limit of high average input power P. In particular,if hiet(f, P) uniformly 1,VfEY ilt(f, P) as P -> o, (6.20) and (f * P uniformly , F) fE oo,Vf E - as P -+ 00, then given any e > 0, 3Po, s.t. VP > PO, 84 (6.21) 1 < Chet(P) < 1 + C. Chet(P)- Proof - (6.22) The left hand inequality of Eqn. (6.22) has already been proved in [18]. Here, we will prove the right-hand side of the inequality. Given e > 0, we want to show that there exists a Po s.t. VP > PO, the right hand side inequality of Eqn. (6.22) is satisfied. Choose any 6 > 0. Eqn. (6.20) implies that, given any 6 > 0, ]P*, s.t. VP > P* and Vf E F, 1-6 < het' P) <1 + 6. nUlt (f, P) (6.23) Define - log(1 + (1 - J)Wu;t (f,P*)) log(1 + h*lt(f, P*)) As log(1 + x) is an increasing function of x for x > 0, E"(6, f) > 0, Vf E F and for all 6 > 0. Define a function f (x) = log(1 + (1 - 6)x) - (1 - E"(6, f)) log(1 + X). (6.25) It is easy to see that h*It (f,P*) is a zero of f(x), and that f(x) is an increasing function of x, for all E"(6, f) > 0. Now, we use Eqn. (6.21) to conclude that given any N however large, ]P**, s.t. N= P > P** => i4(f, P) > N, Vf E F. Choose *It (f,P*). We conclude from Eqn. (6.24), and the fact that f(x) is an increasing function of x, that VP > P**, log(1 + (1 - 6)h*t(f, P)) > (1 - e"(6, f)) log(1 + h*t(f, P)). (6.26) Let us define E"(6) = sup (E"(6, f)), f 85 (6.27) so that we may now rewrite Eqn. (6.26) as, VP > P**, log(1 + (1 - 6)*it(f, P)) > (1 - E"(6)) log(1 + init(f, P)). (6.28) We know from the previous sections that: Cult(P) _ Chet (P) g9(h*It(f,Ip)) df fm (6.29) f f*na log (1 + haet (f, p)) d From Eqn. (6.29) and Eqn. (6.23), and the fact that log(1 + x) is an increasing function of x, we conclude that for all P > P*, g (h*It(f P))df <Cuit(P) (f, P))df Chet(P) + 6)i*1l(fIP)d fmax Imin log(1 + (1 ffax ff - g (f*It (f,P))df nmmn fj-x log(1 + (1 - 6)It(f, P))df (6.30) Next, we define P*** = max(P*, P**) (6.31) Using Eqn. (6.28) and the right hand inequality of Eqn. (6.30), we conclude that for all P > P***, Cult (P)fX(i*(f, Chet (P) ~ (1 - P))df log (1 +i*I(f,P))df I"(6)) ff"a We know that the function g(x)/ log(x) - (6.32) 1 from above, as x -+ oo. So, given any E',7 x*, s.t. VX > X* < - g log(1+ X) < 1 + 6'. (6.33) Choose c' = (1 + e)(1 - E"(6)) - 1, and denote the corresponding x* = M. From the uniform convergence of i*l(f, P) (Eqn. (6.21)), we can say that given this M, P****, Is.t. VP > P****, in*(f, P) > M. Therefore VP > P****, 86 1< '~P)) f -< 1 + [(1 + W)( - E"(6)) - 11 . (6.34) Po = max(P****, P***) (6.35) -log(1 + i1*1t(fI P)) Finally, we define so that for all P > Po and Vf E T, P)) ; [log(1 + n*t(f, P))] (1 + f)(1 - E"(6)). g(A1*a(f, Substituting Eqn. (6.36) into Eqn. (6.32), and using the fact that Cult(P) (6.36) Chet(P) [18], we conclude that < Cuit(P) < 1 + Chet(P) - which is the statement we set out to prove. 87 (6.37) Chapter 7 Capacity of a Lossy Channel with Thermal Noise In the previous chapters, we observed that free-space transmission of a single spatial mode of the bosonic field through circular apertures can be treated as an interaction of the input field mode & with an environment mode b through a beam splitter of transmissivity 1, where q is a function of some geometrical parameters and the center frequency f, of the carrier (2.1). In the absence of any extraneous noise sources other than pure loss, the environment mode 6 is in the vacuum state, which is the minimum 'noise' required by quantum mechanics to preserve correct commutator brackets of annihilation operators of the output modes (or in other words, to satisfy the Heisenberg uncertainty relation at the output). Different types of noise can be described by different initial states of the environment mode 6. In this chapter, we will define a thermal noise lossy bosonic channel, which describes the effect of coupling the lossy channel to a thermal reservoir at temperature T [7]. The ultimate classical capacity for this channel is not known yet, though there have been some very recent advances done in this direction by several people in, and associated with the research group of Prof. Jeffrey H. Shapiro, at RLE, MIT. This chapter will summarize some of that work. We will also show that wideband and SM capacities of the thermal noise lossy bosonic channel can be calculated explicitly (either analytically or numerically) for all the quantum encodings and receiver structures considered earlier in this thesis. 88 Background 7.1 A thermal noise lossy bosonic channel at temperature T > 0 (referred to as the thermal noise channel from now on to save on nomenclature), is defined to be a single mode (SM) bosonic channel (2.1), in which the environment mode 6 is excited in a zero-mean isotropic Gaussian state PT -- 7FN pT, given by the density matrix: a) (aId 2 a, exp ( _IN (7.1) where 1a) is a coherent state. N is the mean photon number of the thermal state PT, given by the Planck's blackbody formula N = Tr~p ) = ehf/kBT -1 where h and kB are the Planck's constant and the Boltzmann's constant respectively, f and is the frequency of excitation of the mode. The thermal state PT has the diagonal representation: 1 In)(nI, PT = IV,_ (7.3) where In) are the eigenvectors of the number operator btb, corresponding to eigenvalues n = 0,1 .It is interesting to note that this channel can be further decomposed into a lossy bosonic channel without thermal noise (2.1), followed by what we define below as a classical Gaussian noise channel D(1@)_ (see Fig. 7-1). A classical Gaussian noise channel is characterized by the completely positive (CP) trace preserving map (N(P) 1 ?N which transforms an input state 3 into the state given by the expression: 4N(p) =1 D(a) D(a)d2G, 7rN where 89 (7.4) b(p,,N) Environment (in thermal state) a c=J)&+ A-27 Output Input 77 (a) Thermal Noise Channel (with loss) III Environment (in vacuum state) a da Input (b) Pure lossy channel + D(-7; C Classical Gaussian noise channel Figure 7-1: The thermal noise channel can be decomposed into a lossy bosonic channel without thermal noise (2.1), followed by a classical Gaussian noise channel <D(_,) (7.4). D(ce) = exp(atd - a*&) (7.5) is the unitary displacement operator in the phase space. This thermal noise model can be applied to our free-space quantum optical channel, now with a background blackbody noise source at thermal equilibrium at temperature T. The thermal noise channel as explained above can also be used to describe single mode (SM) quantum optical communication through a lossy optical fiber or a waveguide in the presence of thermal noise. In the power constrained wideband version of the thermal noise channel, a much simpler version of the above model occurs in the case of 'white thermal noise', in which we have one fixed value of mean number of thermal photons N for all frequencies of transmission. This case, though not too realistic, is easier to handle analytically for capacity calculations. Additional N(f) possibilities might also be considered. For example, the noise introduced by an optical amplifier at any particular frequency will have density operator of the form (7.1), but its average photon number - 90 as a function of frequency - will not follow Planck's law, nor will it be a constant. 7.2 Ultimate Classical Capacity: Recent Advances Consider a SM lossy thermal noise channel of transmissivity q, at temperature T. Let there be a mean photon number constraint of i! at the input. Let us denote the mean photon number of the thermal environment by N, which is given by Eqn.(7.2). Let us denote the completely positive (CP) map describing the action of this thermal noise channel on an input state, by EN. Also, we use {} and {p} to denote the set of input states forming the input alphabet of the channel, and their apriori probability distribution respectively. It was shown earlier this year [7], that the capacity of this channel C(r, ii, N), can be lower-bounded by C(7, ii, N) g(nii + (1 - T)N) - g((1 - 77)N). (7.6) Proceeding along similar lines to the ultimate capacity proof in [18], one can upper-bound the SM capacity as shown below. C(q ,f, N) =max [S(EN(ZPi)) [pIn1 < 33 max S(&SN(pjp)) - (7.7) ZPjS(,E(N)) [i - min [S(zE( 71 P j (78 (7.9) where all the above maximizations are done under the mean input photon number constraint h = E pTr(Wdf&p). As the thermal environment state P/T is a zeromean state, the mean photon number at the output of the SM channel is given by rqih + (1 - q)N. As all the output states of this channel will have this mean photon number, the first term on the right-hand side of the last inequality (7.9) can be further upper-bounded by max,3 S(p), with the constraint Tr(dt&3) < "n + (1 - 77)N. But, 91 we know that the maximum entropy subject to a mean number constraint fl, is given by g(ii). So, we now have: C(, ii, N) < g(jn + (1 - i)N) - min [S(EN())]. (7.10) The next thing we need to do is to find the second term in Eqn. (7.9), i.e. the minimum output entropy of a SM thermal noise channel. We stated in the previous section that the thermal noise channel EN can be decomposed into a composition of two channels, viz. F N where S' represents the SM lossy channel - 4 (1-9)N without thermal noise, and I(1-n)N is the classical Gaussian noise channel defined in the previous section. The minimum output entropy of S1 is zero, because any coherent state at the input produces a coherent state at the output, which has zero entropy by virtue of it being a pure state. In particular, the vacuum state input 10) sent through I produces a vacuum state at the output, which clearly has zero entropy. Also, the minimum output entropy of ((1-,)N is less than or equal to that of £,, because in general there need not exist any input state to S that would produce at the output of 6', the minimum output entropy achieving input state for the classical channel 4 (1-o)N (see Fig. 7-1). Thus, we have min [S(E,())] > min [S(Q(1-7)N(5))] , (7.11) which along with (7.10), implies C( 1 , h, N) g(7ii + (1 - 71)N) - min [S(<P(1-n)N())] (7-12) Clearly, when they act on the vacuum input state 10), both the channels 4I(1-q)N produce the thermal output state of mean photon number N, given by 92 and SEN (10)(01) (7.13) 4(1-n)N 1)(1 = I ' ( (1-r/I)N +1 Y-- S) (1 - ))N(IO)n)(nO (7.14) (1-r/I)N+1I exp (1 2a. -a)(ad (7.15) A coherent input state Ia) produces a displaced version of the above thermal state, whose mean complex amplitude is equal to ,/a, from that of the vacuum-state input, i.e. g((1 that the minimum output entropy of D(1-,)N but whose entropy is unchanged 7 )N). So, if we are able to show is achieved by a coherent state input, that would imply that the minimum output entropies of both these channels are the same, and given by g((1 - 7)N). This argument follows from the decomposition 1, = D(1-,)N o , , and the fact that any coherent state can be obtained at the output of S by a suitable coherent state input. From the concavity of Von Neumann entropy S(), we may infer that the minimum output entropy of both these channels can be achieved by pure input states. Also, it can be shown that displacing a state in the phase plane does not change its output entropy. Hence, we may limit our search (for input states that yield the minimum output entropy) to zero-mean pure states. The purpose of introduction of the classical Gaussian noise channel '1 N was that this channel looks much simpler than the general SM thermal noise channel F, and to find the minimum output entropy of E suffices to prove that the minimum output entropy of 'I(1-7)N , it is achieved by coherent state inputs. Conjecture 1 (GGLMSYY) The minimum output entropy of the SM thermal noise channel SN is achieved by coherent input states, and hence is given by min [S(EN())] g ((1 - r)N) (7.16) Though the above conjecture has not been proved yet, several partial results have been derived. If this conjecture is proved to be true, then from Equations (7.9) and 93 (7.6), we would be able to conclude that the bits per use capacity of the SM thermal noise channel 6', is equal to g(pi + (1 - q)N) - g((1 - q)N). Some intuition behind the preceding conjecture can be explained as follows. If the input to the thermal noise channel is an isotropic state (i.e. circularly symmetric in the phase plane), then the output is also an isotropic state, the reason being that the anti-normally ordered characteristic function of the output of a beam-splitter for uncorrelated input states can be decomposed into the product of the anti-normally ordered characteristic functions of the input states; and the product of two isotropic functions is isotropic as well. It can be shown that if we limit ourselves to zero-mean Gaussian input states, the one that achieves minimum output entropy is the vacuum state, which clearly is isotropic. So it seems that one might expect the minimum output entropy of S' to be achieved for an isotropic state at the input. If this belief is true, then we have numerical evidence to show that in the class of the all isotropic input states, the vacuum state has the least output entropy. We stated earlier that from the concavity of Von Neumann entropy, we could conclude that the minimum output entropy of the channel can be achieved by pure input states. It can be shown analytically, that the number states, which include the vacuum state, are the only states which are pure as well as isotropic. Also, we know that if one state majorizes another state, its entropy is smaller, i.e. ,1 >- P2 =4 S( 2 ) < S( 2 ). So, if we can show that the output state corresponding to a 'lower' number state majorizes the output state corresponding to a 'higher' one, then we would be able to conclude that among all number states, the vacuum state 10) has the least output entropy; and hence would be able to assert that, in the class of all isotropic states, the vacuum state input yields the least output entropy, which we know is equal to g((1 -q)N). 7.2.1 Number States: Majorization Result In this sub-section, we will try to demonstrate the assertion we made towards the end of the last section regarding the majorization of output states corresponding to input number states. 94 Conjecture 2 Given two photon number states In1)(nil and |n 2 )(n 21, n 1 < n 2 -> E N(Ini)(nil) >- S,'(1n2)(n2l) In the section on the capacity of the thermal noise channel Evidence (Numerical)- with number state inputs and direct detection (later in this chapter), we show that the output of the thermal noise channel for an input number state In) (nI is given by sN( n)(n1) Pn,mIm) (ml, = (7.17) m=O where the coefficients Pn,m are given by m + n\ (1 Pl m m = 2 )m+n(l + N)nN m (1 + (1 - r7)N)m+n+l -- m-n; -(i x) F[_ 7(1 with F[a, 3; -y; z] - F(a, ;;z) 2 F 1 (G, = 1+ + n), (N - (1l + N))(1 + (1 - r)N)1 - 77)2N(1 + N) (718) 3; -y; z) being the hypergeometric function 7.-1 z + a(a+1)0(0 +1)z2 -y(- + 1) - 2! a(& + 1)(a + 2)/3(3 + 1)(0 + 2) 3 + +(y + 1)(7 + 2) -3! (7.19) From the definition of majorizationwe have, m m if Ipni,k E Pn2,k, k=O k=O Then, S N(Ini)(nil) >In Fig. 7-2, we plot EM_ 0 , Vm - [0,Oo3, SN(In2)(n2l). (7.20) as functions of m for several values of n, for a thermal noise channel with r/ = 0.8 and N = 5 photons. As expected, we find that EkAO Pnk - 1, as m -* oc, because Pn,k's are probability distributions. More- over, we observe that as n increases, the curves shift downwards. Thus, at any given 95 I1860 (r DEEM) - - I/ 0600 - / / / *1* I / I I I / I k 0 OAE4 02M0 I I# I // / n90 2' 0Da00.00 2000 4000 80 60M10 100A M Figure 7-2: A plot of E'U 0 Pn,k for a SM thermal noise channel with transmissivity The dark r/ = 0.8, and N = 5 photons, as a function of m; for n = 0, 6, 12, 18,. (black) lines correspond to the values n = 0, 30, 60, 90. value of m, _ e Pn,k decreases with increasing n, supporting our majorization con- jecture. Given the nature of the expression for the eigenvalues pn,m of the output states, it seems difficult to construct a general analytical proof of the majorization conjecture. However, numerous calculations with different values of transmissivity and mean thermal photon numbers support our conjecture. As a result of our numerical work, we deem it highly likely that output states corresponding to 'smaller' number-state inputs majorize (and thus have smaller values of Von Neumann entropy than) the output states for 'larger' number-state inputs (Fig. 7-3). Hence, we assert from our earlier discussion that - in the class of all isotropic input states, the vacuum state input yields the least output entropy for the SM channel E' which we know to be equal to g((1 - r)N). 96 0. - -21=,00= 1100 S-n=4, 0.300 ,N=5 - n=3, N=5---N=5 ------- 0.200 ---- -n 00 OAOO .0800 1.000 eta (transmissivity) Figure 7-3: A plot of output entropies of the SM thermal noise channel as a function of q for number states inputs In)(n1, corresponding to n = 0, 1, 2, 3 and 4. The value of N = 5. We see that output entropy is minimum for the vacuum state 10) (01, which is shown in the figure by the dark black line. The results for n > 0 were obtained by numerical evaluation of the p.,m from (7.18); the n = 0 entropy result so obtained matches the analytic result g((1 - 7)N). 97 7.3 Capacity using known Transmitter-Receiver Structures We saw in the last section that the ultimate classical capacity of the lossy thermal channel is as yet an unsolved problem. Nevertheless, we have intuitive reasons and evidence to believe that the SM capacity of the thermal noise channel is given by CQft, N) = g(rpi + (1 - q)N) - g((1 - q)N). In this section, we will evaluate capacities of the thermal noise channel E,N that can be achieved by some of the standard encoding schemes and receiver structures. 7.3.1 Number States and Direct Detection Let us transmit a pure number state AIN = In)(n with probability pN(n) through a SM lossy thermal noise channel with transmissivity q and thermal mean photon number N. Let us denote the channel by the Heisenberg evolution (see Fig. 7-1(a)): 6 = AT& + /1 -ab (7.21) where &and 6 are the annihilation operators of the input and output modes respectively, and b denotes the thermal environment. The initial states of the input modes are given by pa b = In)(n 00 Nn + 1)n+1 In)bb(nI n=O (N (7.22) (7.23) The first step in finding the capacity is to find the output state corresponding to the input state In) (nI and express it in the number state basis. Using this representation, we can read off the transition probabilities which can be plugged into a Blahut Arimoto algorithm routine to obtain the optimum input distribution PN(n) and the capacity. 98 We can find the anti-normally ordered characteristic function of the output state as follows [22]: ( * ) = (e -C*eC (7.24) ) /hae 5 Vat) (e-* \beC = (e-* = xi(VP = (e~KI(n e = (e-71K2 *,V()x( 1-*, C( e -VC(* l)) (7.25) (v/FZ-Ibt) (x ( (7.26) -1 () 1 - g(*, /1 - (nIleC te-NvC*arln)) (e-(1+N)(12(1 (7.27) (7.28) -) where we have used the fact that the input to the channel is independent of the thermal environment. In the penultimate step above, we have expressed the antinormally ordered characteristic function of a in terms of its normally ordered form. Also, it can be shown easily, that the anti-normally ordered characteristic function of a thermal state with mean photon number N is given by e-(1+N)1|2 . To calculate (nIe 0a7 e-vC*61n), we express the exponent operator in its Taylor series, and use the 'annihilation' property of d, to obtain (njeOI 6t e-01 *61 (n- kI|/))) S( 2)k k=O (/-1(*)k n | n - k) (7.29) (7.30) ! () (7.31) =Cn(- 1(12),7 where the Cn(x) are the Laguerre polynomials. Now, we find the output state by the operator inverse Fourier transform. 99 = (7.32) ((*, ()e--C eC*6 X J ( _-(I+N)1(12(l -'),Cn 7Ce- 771(2 )-( _ q(2 te(* (7.33) Because the input states are diagonal in the number state basis (7.22) (7.23), it turns out that the output state of the channel is also diagonal in the number state basis, as can easily be verified by evaluating (mifIm') for m $ m' by doing the integration in polar coordinates. The mh diagonal matrix element of fc is given by (mI,3cIm) = (M~firIir2f7rr =- = J d 2x - N-y)(| (7.34) TI~2'mjj)(-4 0-0+-1)(2L re-(1+N-N)r2 ]d] e-(1+N-N)x 2(,r 2)dr Cn(7x)2m(x)dx. Note that if we substitute 77 = 1 into the above expression, we get (mIcIm) (7.35) (7-36) = 6mn, by virtue of the orthonormality relation for the Laguerre polynomials, which complies with the lossless channel. The lossy-case integral can be evaluated explicitly in terms of the Hypergeometric function [10]. The final expression for (mfi.c1m) is given by M m + n (1- )m+n(1 + N)"N ( m ) (1 + (1 - i)N)m+n+1 x F [m, -ni; -(in + n, (N - 7 (1 + N))(1 + (1 - 7 )N)] (1 - -) 2 N(1 + N) (737) Thus, the output state of the channel is given by 00 fc = Z (mjicIm)Im)(mT, (7.38) m=0 hence, the transition probabilities (2.11) for a single use of the channel, when a unity 100 efficiency direct detection receiver is used at the output, can be written as PMIN~mn c -Tll) (7.39) To get the SM capacity (in bits per use) of the thermal noise channel using number states and direct detection, we have to maximize the classical mutual information I(M; N) (2.12) over all input probability distributions PN(n) subject to a fixed value A of mean photon number at the input. This constrained maximization problem can be solved numerically using the Blahut Arimoto (BA) Algorithm [16] (See Appendix A for the details of the algorithm). Interestingly, on carrying out the BA algorithm [16], we find that the capacity achieving probability distributions have multiple peaks similar to the pure lossy case (See Fig. 7-4). This again suggests that to achieve the best capacity using number state inputs one would almost preferentially use a set of optimum numbers of photons in each channel use. In the pure lossy case, when the transmissivity of the channel is increased towards unity, the peaks disappear and the optimum distribution converges to the exponential Bose-Einstein distribution (linear on a log scale). We have numerically evaluated the optimum distributions for the pure lossy case (,q - 0.8, A = 3.7), and the thermal noise channel (,q = 0.8, A = 3.7, N = 5). It can be seen from plots of these distributions (Fig. 7-4), that in the pure lossy case there are no peaks at this value of transmissivity and the distribution is close to linear, whereas for the thermal noise lossy channel the distribution still has multiple peaks. For the optimum distribution of the thermal noise channel to converge sufficiently close to the Bose-Einstein distribution, T1 has to be even closer to unity. To obtain the SM capacity of the thermal noise channel, we have to execute the BA algorithm multiple times for different h constraints. A sample result is shown in Fig. 7-5 for 77 = 0.8, and N = 5 photons. 101 0300 I Lossy chann e(eta=0.8) --no thennal noise Lossy chann 1 (eta=0.8) -- with thermal noise (N-5) -101300 - -20N - -- a4 ~30.00- N i N 0300 10a00 20300 30.000 50.000 40.001 A. : e e e -40O 60.000 70.300 -I 1 9020W 801300 Number of transmitted photons (n) Figure 7-4: A plot of optimum apriori probability distributions (in dB) for a pure SM lossy channel (q = 0.8, h = 3.7), and a SM thermal noise channel (q = 0.8,ii = 3.7, N = 5). In the pure lossy case, there are no peaks at this value of transmissivity and the distribution is close to linear, whereas for the thermal noise lossy channel, the distribution still has multiple peaks. For the optimum distribution of the thermal noise channel to converge sufficiently close to the Bose-Einstein distribution, n has to be even closer to unity. IA41N 1.200 i - 1100 - 0. 0.300 - I 0.600 0A0 - 0.200 - ' 0.000 0.000 21300 41300 61000 8100 101100 1 12AM0 Mean input photon number (nbar) Figure 7-5: A sample calculation of the number-state direct-detection capacity (using the Blahut-Arimoto algorithm) of a SM lossy thermal noise channel with transmissivity i = 0.8, and mean thermal photon number N = 5. 102 7.3.2 Coherent States and Homodyne or Heterodyne Detection To evaluate the capacities of the thermal noise channel with coherent state inputs, it is useful to think in terms of the decomposition of the thermal noise channel E' into a pure lossy channel g followed by a classical Gaussian noise channel 7-1). If a coherent state Ia)(al is input to the classical channel 4 N, 'I(1-n)N (Fig. the output is a displaced thermal state given by exp 1 7r N 1- N) I!)(O3d N 2 0. (7.40) It can be easily shown that if heterodyne detection is done on a displaced thermal given by (7.40), where a = a 1 +ja 2, the output is a pair of Gaussian random variables with mean and covariance function given by ail J )(aj) ~N( -- Heterodyne - N+1 ((a2 With the help of the decomposition gN S 0 ,(2 0 0(1-q)N, (7.41) 2- and the above result, (7.41), we can use reasoning similar to the one we gave in Chapter 4, to evaluate the SM heterodyne detection capacity in bits per channel use, with the following result: Ccoherent -heterodyne(n ,if, N) = 2 = log2 log 2 1 + 1 + (l 7/2 (7.42) l). (7.43) (1 - 71)N + 1 The homodyne capacity can be computed using arguments similar to the ones in Chapter 4, along with the fact that the quadrature variance of homodyne measurement on the thermal state (7.40), is equal to (2N+1)/4. This can be shown by finding the distribution of the homodyne detection outcome by integrating out the second 103 quadrature of the Wigner function of the thermal state. The homodyne capacity (in bits per channel use) is thus given by Ccoherent-homodyne(7, = h, N) 1 -log 2 = 7.3.3 10g2 2 1 1 + (2(1-'q)N + 1)/4) (7.44) + (7.45) N .1 2(1 - rq)N + 1 Squeezed States and Homodyne Detection Let us transmit a squeezed state & = Iai)(r,)(r,O) (a1 with an apriori probability measure PA1 (ai)dai, at each use of the SM lossy thermal noise channel 8 N. All squeezed states are Gaussian states. It can be shown, via characteristic functions, that if the input modes & and b in the SM lossy channel ( = V/& + l - rjb) are excited in Gaussian states, then the output mode a is also in a Gaussian state. The first quadrature mean and the variance of the output state are respectively given by (&1) = Vai, (1 (A12) (7.46) and 2N + 1 4 + (14- ) (7.47) The conditional probability density to read x1 at the output given that a 1 was transmitted, is therefore given by PX1IA 1 (Xi a i ) (7.48) = Tr(pc1Xi)(XiI) 1 /27r (~{(1 - rj)(2N + 1) + exp e-2r]) - 2 (1 [(1 - q)(2N + 1) +,,,-2r (7.49) Since the channel noise is additive and Gaussian, the mutual information is max104 imized for a Gaussian input density (5.13), and the SM capacity is given by C(r, ii, N, r) = 1 - log 2 2 (1+ 1 4 [ h -- sinh 2 r sn) 2 (1-rq)(2N+1) + q; (7.50) r e-2,r A further maximization with respect to the squeezing parameter r yields the SM capacity (in bits per use): Csqueezed-homodyne(r7, 4fi + 2 - (x(7/, h, N) + x(I, h, N)- 1 ) 1 10o 2 2 h,N) (1-rq)(2N+1) + X(TI, N) (, .1 (7.51) where the function x(rq, h, N) is given by x(r/, h, N) = (1- r)(2N + 1)) +2(1 - )(2N + 1) (1 - r)(2N + 1) +/+2) (7.52) 7.4 Wideband Results and Discussion We saw that the capacity of the SM lossy thermal noise channel can be evaluated for standard encoding schemes and receiver structures. The wideband capacities can be found by proceeding exactly in the same manner as we did for the pure lossy channel. The only difference in this case is that there is now an additional frequency dependence of the thermal mean photon number per mode (7.2) that one has to take into consideration while doing the Lagrange multiplier maximization. This makes the calculation of thermal noise capacities very involved analytically. Before turning to wideband calculations, however, we introduce one more SM capacity measure, and then compare all our capacity results for the SM lossy thermal noise channel. It was shown recently [7], that the 'entanglement assisted classical capacity' of a 105 3100 B,- 2500 2.000 Ultimate capacity (???) Coherent-Heterodyne Coherent-Homodyne - Squeezed-Homodyne - -- - ------ ------Entanglement assisted Number state Direct detection 0.500 OyC10 0.000 6100 41100 200 101100 Sam 12.00 Input mean photon number (nbar) Figure 7-6: A sample simulation result of all the different capacities of a SM lossy thermal noise channel we considered so far in the thesis. The channel S has transmissivity r = 0.8, and mean thermal photon number N = 5. SM lossy thermal noise channel EN is CEntanglement -asst.(I, i, N) = g(i) + g(N') - g( +n D -N' 2 1 D- + N'- 1 2 ' (7.53) /(N + N'+ 1) where N' _ ih + (1 - q)N, and DE 2 -4]N(N + 1). As noted in Chapter 2, this capacity provides an upper bound on the ultimate capacity of the SM lossy thermal noise channel. Now, we are in a position to plot all the SM capacities of the thermal noise channel. For the case q = 0.8 and N = 5, we have plotted the different capacities in Fig. 7-6. The plot with the '?'-sign pertains to what we think the ultimate (unassisted) classical capacity is. We note that the entanglement-assisted capacity is indeed an upper bound to all the other SM capacities. Also, as one would expect, the conjectured ultimate capacity is greater than all the structured transmitter-receiver capacities, and is less than the entanglement-assisted capacity. 106 The wideband capacities can be found by performing the constrained maximization numerically. We consider two cases. The first case is 'white Gaussian noise', where we consider the free-space optical channel with single spatial mode propagation (ultra wideband) with same geometrical parameters and bandwidth as we considered in Chapter 4, and assume that across the entire frequency band the mean thermal photon number per mode N = 5 remains constant (Fig. 7-7). The second case we consider (Fig. 7-8), is the free-space channel with single spatial mode propagation with the same parameters as above, but with a frequency dependent thermal mean photon number distribution (7.2) given by the Planck's formula. In both the above cases, it can be seen that the relative behavior of the various wideband capacities are the same. Coherent State encoding and heterodyne detection is smallest in the power limited regime, and catches up with both homodyne detection capacities (with coherent state and squeezed state encoding respectively) in the bandwidth limited (high power) regime. The (conjectured) wideband ultimate capacity is greater than all the structured transmitter-receiver capacities at all values of input power. 107 2 B0Oe17 Ultimate (???) 1e00.17 - .1.200,17 - .e '1 J8130016 4D0. - Squeezed-Homodyne Coherent-Homodyne ------Coherent-Heterodyne - -- - Entanglement Assisted ------ - ----------------- 1 QMO 8.000 6.00 4.00 2.00 0.000 10-000 Input power (Watts) Figure 7-7: Ultra-wideband capacity of a lossy thermal noise channel - Free-space 2 optical communication, single spatial mode propagation - AT = 400 cm and AR = 1 cm 2, separated by L = 400 km of free-space; with frequencies of transmission ranging from that associated with a wavelength of 30 microns (which is deep into the far-field regime), to that associated with a wavelength of 300 nanometers (which is well within the near field regime). Mean thermal photon number N = 5 is assumed to remain constant across the entire range of frequencies. 2OO0e16 00* -- 4.oooe15 I 0.00 0.000 0.050 0.100 0.150 Ultimate Capacity (???) Squeezed-Homodyne - Coherent-Homodyne -----Coherent-Heterodyne - - - I I 0 .200 0.250 - 0.300 Input power P (Watts) Figure 7-8: Capacities of a wideband lossy thermal noise channel - - Free-space optical communication, single spatial mode propagation - all parameters same as in Fig. 7-7, except that the mean thermal photon number N is now assumed to vary with frequency of transmission f according to Eqn. (7.2) 108 Chapter 8 Conclusions and Scope for Future Work The fields of quantum information and quantum communication have grown tremendously over the last two decades, and there is a lot of interesting research that is currently going on in these fields. Quantum Information theory sets theoretical limits on achievable information rates that can be sent reliably over quantum channels. In this thesis, we studied a broad class of classical information capacities of singlemode and wideband lossy bosonic channels. A lossy bosonic channel is a general model of a quantum communication channel, which can be used for physical modelling of communication over free-space, waveguides, lossy optical fibres, etc. Our focus in this thesis was the free-space quantum-optical communication channel. We evaluated the ultimate classical capacity of the broadband free-space optical channel both in the far-field and near-field propagation regimes and compared it with achievable rates using known transmitter and receiver structures. We found that, for the free-space channel operating in the far-field regime, the performance of coherent-state encoding and heterodyne detection asymptotically approaches the far-field capacity in the limit of high input power. We also generalized this fact by proving a set of sufficient conditions, for heterodyne detection capacity to approach the ultimate capacity in the high-power limit. We found the capacities achievable using structured transmitter encoding and detection schemes, for the single-mode and the wideband lossy bosonic 109 channels with additive thermal noise. We also evaluated the wideband entanglementassisted capacity of the free-space channel. The ultimate classical capacity of the thermal noise lossy bosonic channel is yet to be found. We have conjectured this capacity, but the proof is not yet done. The capacity of the general multimode bosonic channel is not known. It would be interesting to know how atmospheric effects like turbulence, fog, and dust affects the capacity of classical communication using quantized light. Another interesting thing to know would be the kind of receiver structures that could achieve the ultimate capacities for the lossy bosonic channel with and without thermal noise. Not much work has been done yet on the quantum information capacity of quantum channels, which could also be interesting to pursue, because it is conceivable that with the advent of robust quantum information hardware in the future, with quantum information running through networks of quantum computers, a good understanding of quantum information capacities might become necessary. 110 Appendix A Blahut-Arimoto Algorithm Consider the following problem [2]: given two convex sets A and B in R, we would like to find the minimum distance between them dmin min d(a, b) acA,bcB where d(a, b) is the Euclidean distance between a and b. (A.1) An intuitively obvious algorithm to do this would be to take any point x E A, and find the y E B that is closest to it. Then fix this y and find the closest point in A. Repeating this process, the distance can only decrease. It has been shown that [17], if the sets are convex and if the distance satisfies certain conditions, then this alternating minimization algorithm indeed converges to the absolute minimum. In particular, if the sets are sets of probability distributions and the distance measure is the relative Shannon entropy, then the algorithm does converge to the minimum relative entropy between the two sets of distributions. This is the central idea on which the famous BlahutArimoto (BA) algorithm is based. Given a channel transition probability matrix P(ji), this algorithm can be used to find the channel capacity under possibly more than one linear constraint by finding the optimum input probability distribution. This algorithm works by iteratively improving the estimates of the input and the conditional distributions. Let pi denote the input distribution of a classical channel, and let P(jli) denote 111 the transition probabilities. Suppose there is a linear constraint at the input of the channel given by E, pjej ; E. Then the steps of the BA algorithm to find the optimum (capacity achieving) input distribution p* are illustrated in Fig. A1. If ej = j, Vj then the linear constraint becomes a constraint on the mean of the input distribution. s is an independent parameter that has to be chosen before executing the algorithm. The final value of the linear constraint E that the algorithm eventually converges to depends upon the selection of the value of s. Unfortunately, there is no very good general way to estimate the value of s which yields a desired value of E. At each iteration, IL and Iu are universal lower and upper bounds to C(E) - sE. The iterations are terminated when the absolute difference between Iu and IL becomes smaller than some pre-decided e. This algorithm works very well to find capacity achieving distributions for the lossy single mode number state channel, with or without thermal noise. 112 START Assign {pj(O)} r =0 ci =exp P(i)10 log P(I -) se zp '"P(j Ii) (max S= l org(+ p , L" = p , ,.g~ STO '.c i I F pi(r+1) = C, i(r) Figure A-1: Blahut-Arimoto Algorithm 113 STOP Bibliography [1] Gagliardi, R. M., Karp, S. Optical Communications John Wiley & Sons, Inc., 1976. [2] Cover, T. M., Thomas, J. A. Elements of Information Theory John Wiley & Sons, Inc., 1991. [3] Gallager, R. G. Information Theory and Reliable CommunicationJohn Wiley & Sons, Inc., 1968. [4] Nielsen, M. A., Chuang, I. L. Quantum Computation and Quantum Information Cambridge University Press, 2000. [5] Yuen, H. P., Shapiro, J. H. Optical Communication with Two-Photon Coherent States-PartI: Quantum-State Propagationand Quantum-Noise Reduction. IEEE Transactions on Information Theory, Vol 24, No. 6, pp. 657-668, Nov. 1978. [6] Yuen, H. P., Shapiro, J. H. Optical Communication with Two-Photon Coherent States-PartIII: Quantum measurements realizable with photoemissive detectors. IEEE Transactions on Information Theory, Vol 26, No. 1, pp. 78-92, January 1980. [7] Giovannetti, V., Lloyd, S., Maccone L., Shor, P. W. Broadband Channel Capacities Physical Review A, vol. 68, 062323, 2003. [8] Yuen, H. P., Ozawa, M. Ultimate Information Carrying Limit of Quantum Sys- tems Phys. Rev. Lett., vol. 70, pp. 363-366, 1993. 114 [9] Davies, E. B. Quantum Theory of Open Systems Academy Press, London, 1976. [10] Gradshteyn, I. S., Ryzhik, I. M. Table of Integrals, Series, and Products Academic Press, 2000. [11] Holevo, A. S. Probabilistic and Statistical Aspects of Quantum Theory North Holland, Amsterdam, 1982. [12] Kraus, K. States, Effects, and Operations: Fundamental Notions of Quantum Theory Springer-Verlag, Berlin, 1983. [13] Holevo, A. S. Coding Theorems for Quantum Channels eprint arXiv:quantph/9809023 v1, September 10, 1998. [14] Schumacher, B., Westmoreland Sending ClassicalInformation via a Noisy Quantum Channel Physical Review A, vol. 56, pp. 131-138, 1997. [15] Caves, C. M., Drummond, P. D. Quantum Limits on Bosonic Communication Rates Rev. of Mod. Physics. vol. 66, pp. 481-537, 1994. [16] Blahut, R. E. Computation of Channel Capacity and Rate-Distortion Functions IEEE Trans. on Inf. Th., vol. 18, no. 4, pp. 460-473, July 1972. [17] Csiszar, I., Tusnady, G. Information geometry and alternatingminimization procedures. Statistics and Decisions, Supplement Issue 1, pp. 205-237, 1984. [18] Giovannetti, V., Guha, S., Lloyd, S., Maccone L., Shapiro, J. H., Yuen, H. P. Classical Capacity of the Lossy Bosonic Channel: The Exact Solution Phys. Rev. Lett. vol. 92, 027902, January 2004. [19] This idea and the proof has been jointly done by J. H. Shapiro, S. Guha, and B. J. Yen, Research Laboratory of Electronics, MIT. [20] Shannon, C. E. A Mathematical theory of communications Bell System Technical Journal, Vol. 27, pp. 379-423 (part one), pp. 623-656 (part two), Oct. 1948. 115 [21] Bennett, C. H., Shor, P. W., Smolin J. A., Thapliyal, A. V. EntanglementAssisted Capacity of a Quantum Channel and the Reverse Shannon Theorem IEEE Trans. on Inf. Th., vol. 48, no. 10, pp. 2637-2655, Oct. 2002. [22] Mandel, L., Wolf, E. Optical Coherence and Quantum Optics Cambridge Uni- versity Press, 1995 [23] Slepian, D. Analytic solution of two apodization problems J. Opt. Soc. Am., vol. 55, pp. 1110-1115, Sept. 1965. [24] Slepian, D. Prolate spheroidal wave functions, Fourier analysis and uncertainty - IV: Extension to many dimensions; Generalized prolate spheroidal functions Bell Syst. Tech. J., vol. 43, pp. 3009-3057, Nov. 1964. [25] Szajowski, P. F., Nykolak, G., Auborn, J. J., Presby, H. H., Tourgee, G. E. High-power optical amplifier-enabled 1550-nm terrestrialfree-space optical datalink operating at 10 Gb/s Proc. IEEE Mil. Commun. Conf, Atlantic City, NJ, pp. 687689, Oct.Nov. 313, 1999. [26] Chu, P. B., Lo, N. R., Berg, E., Pister, K. S. J. Optical communication using micro corner cuber reflectors Proc. IEEE MEMS Workshop, Nagoya, Japan, pp. 350355, Jan. 2630, 1997. 116