15
COMMUNICATION RANGE EQUATION AND LINK ANALYSIS

15.1 INTRODUCTION

This chapter deals with numerous issues related to the design of the data modulator, transmitter, antennas, receiver, and demodulator subsystems as they relate establishing the system performance over a communication link. A communication link, as defined here, involves issues related to the transmission of data from one point to another and, in addition to the various systems, includes the transmission channel. The focus is on the received signal and noise powers and the resulting signal‐to‐noise ratio required to meet the overall system performance specification. The channel is generally viewed as the transmission medium between a transmitter and receiver location and the emphasis in this chapter is on wireless channels involving transmissions through the atmosphere and ionosphere. The link might include, for example, a satellite relay with unique uplink and downlink channels. The satellite relay may be thought of as a “bent pipe” with frequency translation and power amplification or a full processing satellite in which the data is demodulated and then remodulated for transmission on an independent downlink. The emphasis in this chapter is on the application of the communication range equation in analyzing the system performance over wireless channels. Chapter 18 introduces the subject of fading channels. Chapters 19 and 20 focus is on the impact of atmospheric and ionospheric propagation on the communication performance and discusses mitigation techniques to maintain reliable communications.

The communication range equation is reviewed in the remainder of this introductory section highlighting the major parameters regarding the transmitter, antenna, and receiver designs. In this characterization the impact of the channel is identified by the free‐space signal loss Lfs and the atmospheric loss, Latm. Sections 15.215.10 review a variety of system issues related to establishing the link budget, including antenna gains and patterns, and losses involving rain, polarization, multipath, component interfaces, and miscellaneous system losses. Section 15.11 discusses various aspects of system nonlinearities including solid‐state power amplifiers (SSPAs) and traveling wave tube amplifiers (TWTAs) and their impact on the system performance. Sections 15.12 and 15.13 outline the modeling of the nonlinearities for computer simulations and Section 15.14 provides a case study of the impact of a SSPA on the bit‐error performance of a communication link. This chapter concludes in Section 15.15 with an example link budget highlighting the various parameters involved in the link evaluation and their impact on meeting the system performance specification.

The following analysis makes extensive use of the various subsystem temperatures in degrees Kelvin (°K) and Table 15.1 provides conversion formula for degrees Celsius (Centigrade) and Fahrenheit. The standard noise temperature To = 290°K is defined as the temperature at the input of a device or subsystem that corresponds to a standard noise power based on the bandwidth and noise figure as characterized, for example, by a receiver noise figure Fn and bandwidth B expressed in (15.11).

TABLE 15.1 Commonly Used Temperature Conversion Formulas

deg‐K = deg‐C + 273.15 Celsius → Kelvin
deg‐K = deg‐F + 459.67 Fahrenheit → Kelvin
deg‐C = (deg‐F − 32)(5/9) Fahrenheit → Celsius

Consider a transmitter that delivers a power of PPA watts from the high‐power amplifier (HPA) to an antenna having a peak gain of Gt in the direction of a receiver. The power density of the transmitted signal at a range R meters is given by

(15.1) images

where the factor of 4π radians accounts for the solid angle of the spherical radiation pattern of an ideal dipole antenna in free space. A receiver located at a range R from the transmitter will intercept a fraction of the incident energy. The received power intercepted by the antenna is simply the power density times the effective area, Ae, of the receiver antenna. The antenna effective area is defined as the physical area Aa of the antenna aperture times the aperture efficiency ηa, so the power intercepted by the receive antenna is

(15.2) images

The antenna gain is related to the physical and effective areas of the antenna by1

Upon substituting images into (15.3), the received power becomes

The second expression in (15.4) is a convenient form and results from using the free‐space loss expressed as

Although (15.4) is a concise expression for the received power, additional insight into the system design is obtained by including three additional sources of power loss. These losses account for transmit and receive system radio frequency (RF) losses (Lts, Lrs) and the losses associated with atmospheric absorption (Latm). Generally, the transmitter power is specified in terms of the HPA output power PPA; however, there are transmitter and receiver antenna losses associated with waveguide feeds, rotary joints, radomes, and so on, that must also be considered. Eventually all of the losses will be accounted for; however, currently the losses of interest are as follows: Lts, Latm, Lfs, and Lrs and the received power at the receiver low‐noise amplifier (LNA) input is expressed as

where Pt is the power into the transmit antenna. The atmospheric absorption loss is a function of range and is normally specified in terms of the attenuation‐per‐kilometer (α dB/km), for which, the atmospheric loss is given by

(15.7) images

where Rkm is the range extent through the atmosphere in kilometers. For terrestrial links there is also a frequency‐dependent terrain loss (Lter) associated with terrain fluctuations, foliage, surface dielectric constant, and conductivity; for ELF through LF frequencies this loss is often denoted as a ground loss Lgnd. Terrestrial links are discussed in Chapter 19. A number of contributing sources to the system losses are identified in the following sections and, when included with the RF losses, provide further fidelity to the communication range equation in predicting the overall system performance. There are also several contributing sources to the atmospheric loss and, when taken collectively, these losses comprise the system loss budget.

Equation (15.6) is often used to determine the received signal power at various points throughout the receiver to evaluate the dynamic range requirements and possible saturation, especially at the input to the analog‐to‐digital converter (ADC). The ADC has an instantaneous dynamic range of about 6 dB/bit and is often designed to saturate or clip with input levels on the order of milli‐watts. Therefore, with the aid of (15.6) the specification of the receiver dynamic range and AGC requirements can be identified. There are a number of variations in the communication range equation that are useful in predicting the system performance under different circumstances and these are introduced in the following sections.

15.1.1 EIRP and Power Aperture

Commonly used performance measures for the transmitter system are the effective isotropic radiated power (EIRP) and power aperture. The EIRP is defined as the product of the power into an isotropic antenna and the antenna gain in a given direction relative to the isotropic antenna gain. Usually the EIRP is specified in the direction of the maximum antenna gain. The power aperture is defined as the product PPAAta of the transmitter power and the transmit antenna area. The maximum gain along the antenna boresight axis corresponding to images and the EIRP2 is defined as

(15.8) images

To express the communication range equation in terms of the EIRP, the transmitter loss (Lts) is removed from the link RF losses resulting in the communication range equation

(15.9) images

The communication range equation is expressed in terms of the power‐aperture product, PPAAta, by substituting for Gt using (15.3), with the result

(15.10) images

In this expression Ata = Aa is the physical area of the transmit antenna. The power‐aperture product is used as a performance measure of the transmitter when the physical size of the antenna is an important consideration, for example, when the transmit power and physical size of the antenna tend to dominate the cost of the transmit system.

15.1.2 Signal‐to‐Noise Ratio

Expressing the range equation in terms of the signal‐to‐noise ratio, γ, is necessary when relating the system performance to the signal detection and false‐alarm probabilities or the acquisition and bit‐error probabilities. The noise floor at the input to the receiver is characterized in terms of the thermal noise power in the receiver noise bandwidth Bn and is given by

where images W/s‐°K is Boltzmann’s constant, To = 290°K is the standard temperature,3 Bn is the noise bandwidth in Hertz, and Fn is receiver noise figure. The receiver noise figure accounts for the additional noise contributed by the receiver relative to the intrinsic kTBn thermal noise. The noise bandwidth, as used here, is related to the RF bandwidth at the input to the LNA with prefiltering frequency response H(f) and is defined as

(15.12) images

where fo is the frequency corresponding to the maximum of H(f).4 This definition of the noise bandwidth corresponds to an ideal unit amplitude rectangular noise filter having the same area as |H(f)|2. Using these results and (15.6), the communication range equation is used to express the received signal‐to‐noise ratio as

where Fns is the system noise figure that is somewhat higher than that of the receiver because of receiver antenna related noise sources as discussed in Section 15.2.3.

15.1.3 Maximum Range

The maximum range of a communication link is defined as the range beyond which the signal‐to‐noise ratio falls below the value required to maintain the specified system performance; typically this is viewed as the signal‐to‐noise ratio corresponding to the specified bit‐error probability; however, a more stressful requirement might be the signal‐to‐noise ratio required to support the waveform acquisition probability. In any event, using (15.5) to express the free‐space loss in terms of the range, (15.13) is then used to evaluate the maximum range to achieve the specified performance criteria in an additive white noise environment with the result

(15.14) images

where γmin is the minimum received signal‐to‐noise ratio corresponding to the system performance specifications.

15.2 RECEIVER AND SYSTEM NOISE FIGURES AND TEMPERATURES

The noise figure of a receiver [1] is defined as the ratio of the input and output signal‐to‐noise ratio as

In this section the noise figure of a receiver comprising a cascade of amplifiers, mixers, and attenuators is examined in terms of the noise characteristics of the individual components or devices. For all of the major components in the receiver, the key parameter used to characterize the noise is the effective noise temperature. For devices that contribute to the receiver gain, the effective noise temperature is determined from the device noise figure as defined in (15.15); however, for lossy devices it is determined from the physical temperature of the devise and the loss. The analysis that follows is primarily based on the work of Charton [2].

15.2.1 Receiver Noise Figure

The receiver noise figure is developed in terms of the effective noise temperatures of the individual components of the receiver starting with those that contribute to the system gain. In these cases the input and output signal‐to‐noise ratios are defined by

(15.16) images

and

(15.17) images

where PS is the signal power, To = 290°K is the standard temperature, Bn is the noise bandwidth, Gi is the gain of the device, and k is Boltzmann’s constant. The term GikToBn is the output noise power resulting from the input noise power kToBn and ΔPni is the output noise power contributed by the device.5 Using these results and (15.15), the noise figure of the device is expressed as

(15.18) images

Upon defining the effective noise temperature as images °K the device contribution to the output noise is defined as images so, in terms of the effective noise temperature, the device noise figure becomes

Charton points out that this characterization of the noise figure applies to a heterodyned receiver where the signal image is blocked or filtered and when the image is not blocked (15.19) is expressed as images; the factor of two results when the image power equals the desired signal power.

Because the noise figure of a device is usually specified by the manufacturer, the effective noise temperature is computed as

(15.20) images

The interpretation of the effective noise temperature is depicted in Figure 15.1 in terms of noisy and noiseless devices both with gain Gi. In Figure 15.1a the output temperature is computed as images, whereas, in the noiseless representation of Figure 15.1b, the output temperature is computed as images. Therefore, using the equivalent noise temperature, the output temperature is obtained simply by multiplying the input temperature by the device gain. It will be seen that this also applies for lossy devices when the gains are replaced by the reciprocal of the losses, that is, for a lossy device with gain images the gain is replaced by images so the output temperature becomes images.

Two diagrams illustrating noisy amplifier (left) and noiseless amplifier (right) with boxes labeled (Fni, Gi) and (Gi), respectively.

FIGURE 15.1 Comparison of noisy amplifier with equivalent noiseless amplifier.

Using these results, a cascade of successive amplifiers, such that i = 1, 2, 3, …, with different gains and noise figures, results in an overall receiver noise figure given by

or equivalently in terms of the effective noise temperatures

where Trec is the effective receiver noise temperature given by

From these results it is apparent that the input device should be a high‐gain, LNA, in which case it essentially determines the receiver noise figure. For this reason, LNAs with high gain are typically used at the receiver input.

The noise figure of a lossy device, such as an attenuator, is evaluated in a similar way; however, there are two major differences to consider: the gain of the lossy device is less than unity so the loss is given by images, the second consideration is that the physical temperature (Tpi) of the lossy device is typically used to characterize the effective noise temperature performance instead of the noise figure.6 The equivalent noisy and noiseless attenuators are shown in Figure 15.2 in terms of the physical and effective temperatures. For the noisy attenuator, the output temperature is given by images and, for the equivalent noiseless attenuator, the output temperature is images so, upon equating these results, the equivalent noise temperature for a lossy device is

(15.24) images
3 Diagrams illustrating noisy attenuator (left), noiseless attenuator (middle), and cascade of noiseless attenuators (right) with 3 boxes labeled (Tpi,Li), (Li), and (L1L2…), respectively.

FIGURE 15.2 Comparison of noisy attenuator with equivalent noiseless attenuators.

Applying these results to a cascade of N attenuators i = 1, 2, …, N, as shown in Figure 15.2c, each with loss Li and physical temperature Tpi, the effective noise temperature of the cascaded attenuators is

(15.25) images

and the output temperature is computed as

(15.26) images

Usually the cascade of receiver devices alternates between amplifiers and passive filters or attenuators as shown in Figure 15.3. In this context, evaluation of the overall system noise temperature and noise figure involves the appropriate application of the results in this section.

Flow diagram of receiver configuration from antenna to lossy device to amplifier to demodulator.

FIGURE 15.3 Receiver configuration.

15.2.1.1 Example: Evaluation of Receiver Noise Figure and Temperature

As an example application of the results in Section 15.1.2, consider a receiver as shown in Figure 15.3 having a 20 dB gain LNA with a 2 dB noise figure. The LNA is located as close to the receive antenna output as possible, typically it is attached directly to the antenna waveguide, so that the loss prior to the LNA is associated directly with the antenna. The output of the LNA is coupled to the receivers first RF mixer, usually by a short length of coaxial cable. In the following example, the coaxial cable is assumed to have a physical temperature of Tp2 = 290°K and a loss of L2 dB.

In examining this example, L2 is considered to be the AGC control parameter. The signal from the coaxial cable is then input to the second amplifier with a gain of G3 = 25 dB and a noise figure of Fn3 = 10 dB. This is in turn followed by an AGC gain control attenuator with L4 = 0–70 dB of gain control and a physical temperature of Tp4 = 290°K. Finally, the output of the gain control is amplified by a G5 = 40 dB gain amplifier with a noise figure of Fn5 = 10 dB. The receiver noise figure for this configuration is shown in Figure 15.4 for LNA gains of 20 and 40 dB as a function of the AGC attenuator setting (L4) for various conditions of loss (L2) between the LNA and the second amplifier.

Graph of AGC attenuation vs. noise figure illustrating eight ascending curve plots with labels LNA gain 20 dB and 40 dB.

FIGURE 15.4 Example evaluation of receiver noise figure.

In this example, the system noise figure for an LNA gain of 20 dB and low values of AGC attenuation is degraded from the 2 dB LNA noise figure by 0.25–0.95 dB for feed losses (L2) between 0 and 6 dB. These low AGC attenuation levels are associated with very weak signals relative to the thermal noise floor (kToBn) at the receiver input. As the input signal becomes stronger the AGC lowers the overall gain of the receiver by increasing the AGC attenuation to keep the output level constant. However, in spite of the increasing receiver noise figure, the resulting impact on the system performance is not significant because of the correspondingly high output signal‐to‐noise ratio associated with the higher input signal levels.

For example, consider the case with L2 = 3 dB, a minimum detectable signal with the AGC attenuation L4 = 0 dB, and a system requirement that the minimum output signal‐to‐noise is γo = 6 dB. Under these conditions the input signal‐to‐noise ratio is images or 6 + 2.5 = 8.5 dB. When the input signal level increases by 20 dB the ideally adjusted AGC attenuation becomes 20 dB and the receiver noise figure is about 2.6 dB. Under these conditions the corresponding output signal‐to‐noise ratio is images or 8.5 + 20 − 2.6 = 25.9 dB. As the input signal continues to increase another 20 dB the AGC attenuation becomes 40 dB and the receiver noise figure increases to 9 dB. In this case, the output signal‐to‐noise ratio is 39.5 dB. Therefore, in spite of the dramatic increase in the receiver noise figure from 2.5 to 9 dB, the system requirement that γo ≥ 6 dB is more than satisfied.

The curves in Figure 15.4 for the 40 dB LNA gain demonstrate how the sensitivity of the system noise figure is reduced by increasing the LNA gain. Furthermore, with the higher gain LNA the system noise figure is hardly degraded for AGC attenuations up to 20 dB even with a feed loss of 6 dB following the LNA. The higher LNA gain also provided for a wider dynamic range which may be desirable if the feed loss L2 becomes excessive. From these examples it is clear that the loss between the LNA and the second amplifier plays a significant role in determining the receiver noise figure; however, the focus should be on impact of the receiver noise figure with the minimum detectable input signal, that is, when the AGC attenuation is 0 dB. For example, when the loss L2 decreases to 0 dB the effective noise temperature becomes images = 0°K even though the device remains at the physical temperature of Tp2 °K. Therefore, the lossless feed does not influence the receiver noise performance. In this case, the receiver noise figure is influenced by the noise figure of the second amplifier, that is, Fn3, and, as the LNA gain is increased, this influence is diminished providing improved system performance with the minimum input signal power.

15.2.2 Antenna Temperature

With the evaluation of the receiver noise figure and temperatures in hand, the focus is turned to the antenna temperature Ta which is the remaining unknown parameter required to evaluate the system noise temperature and noise figure. The antenna temperature is characterized at the output of the antenna subsystem as indicated in Figure 15.5. The antenna is depicted as having a radome and a radiating dish with a feed horn as shown in Figure 15.5a. The feed horn is located at the focal point of the dish and, in the receive mode, is used to collect the incident electromagnetic wave energy. A short length of waveguide associated with the feed horn provides a flange for connecting the feed horn to the receiver subsystem. The radome protects the reflecting or radiating surface of the antenna from the elements of weather. Although the radome is optional, in Figure 15.5a the radome, reflecting dish, and antenna feed are considered to be an integral part of the antenna subsystem with their associated losses and physical temperatures. The overall antenna efficiency is made up of the aperture efficiency and the radiation efficiency. The radiation efficiency includes the spillover loss of a reflecting antenna, the reflector and feed ohmic losses and, if used, the radome loss [3]. These losses are included in the antenna gain and sidelobe measurements and are only used here to characterize the noise introduced by the antenna subsystem. Combining the antenna reflector loss with the antenna feed loss as the product of their individual losses, that is, images is valid only if the reflector and feed are at the same temperature (see Problem 2). As indicated in Figure 15.5a, the external noise sources that are within the field of view of the antenna, including the antenna sidelobes, also contribute to antenna temperature. These noise sources include discrete sources (ds) such as stars and our Sun and Moon and background (bck) or distributed noise sources such as stellar gas, our galaxy, the Earth’s surface and thermal noise resulting from atmospheric absorption.

Top: Physical representation with labels free space, atmosphere, radome, reflector, etc. Bottom: Model representation with boxes labeled Discrete sources, Distributed sources, Noiseless atmosphere, etc.

FIGURE 15.5 Contributors to antenna temperature.

To begin the evaluation of the antenna noise temperature, the noise temperature of all external noise sources is characterized at the input to the antenna as images. These noise sources are indicated in Figure 15.5b and are identified as follows: Tb is the noise temperature of distant noise sources entering through the main antenna beam, ρlsl is the fraction of the ground noise temperature seen by the lower sidelobes (lsl) of the antenna, ρusl is the fraction of the atmospheric noise temperature Tatm seen by the upper sidelobes (usl) of the antenna, and images is the effective noise temperature of the antenna radome. The notations Tatm and Tgnd represent the respective physical temperatures of the atmosphere and ground. With the exception7 of images, all of the temperatures at the antenna radome are dependent on the grazing (or elevation) angle and the antenna beamwidth. For example, as the antenna viewing angle is changed the temperature Tb will change as different noise sources come into view. Also, as the grazing angle decreases the factor ρlsl increases with a commensurate increase in the ground noise at the antenna radome; the factor ρusl will also fluctuate impacting the atmospheric noise at the radome. The factors ρlsl and ρusl are evaluated by integrating the appropriate noise temperatures over the sidelobes of the antenna beam in consideration of the attenuation with range. At the elevation angle θm, corresponding to the maximum ground noise temperature seen by the lower sidelobes, the ground temperature contribution factor to the antenna temperature is approximated as

where Gr(θm,φ) is the antenna gain corresponding to θm relative to the peak gain Gr(max). Example values are [2]: ρlsl = ρusl = 0.1 for grazing angles above 15° with ρlsl = 0.49 and ρusl = 0.1 for angles below 15°. The link analysis program discussed in Section 15.15 uses (15.27) with Gr(θ, ϕ) evaluated using the circular aperture antenna described in Section 15.3.2.

Referring once again to Figure 15.5b, the antenna temperature is characterized using the relationships developed in the preceding sections and is evaluated as

where Tain is the temperature of the external noise sources. Evaluation of Tb is discussed in the following two sections and the culmination of the general case is expressed in (15.36) in Section 15.2.2.2. Section 15.2.3.1 provides an example involving the computation of the antenna temperature.

15.2.2.1 Contribution of a Single Sky‐Noise Source to the Antenna Temperature

The analysis of the antenna temperature is simplified by considering one external noise source. In the context of Figure 15.5, images is defined as the temperature of either a distributed or a single discrete noise source, that is, images = Tbck or Tds. In evaluating the antenna noise temperature due to a distant radiating source, the radiation intensity or brightness (b) of the source is characterized by the Rayleigh–Jeans approximation [4] of the radiated flux density per steradian, expressed as8

where k is Boltzmann’s constant. The approximation in (15.29) is valid if the ratio images not to large which is the case for most applications. For example, the error is about 1% for f = 150 GHz and images = 300°K; if the error is unacceptable then Planck’s law must be used, for which the brightness is expressed as [5]

where h = 6.63e−34 W‐s2 is Planck’s constant, c = 3e8 m/s is the free‐space speed of light, the bandwidth B ≪ f, and k = Boltzmann’s constant, previously defined along with the remaining parameters. The factor of two in (15.29) and (15.30) results from the Omni polarized electrometric waves emitted from the radiating source; for a linearly polarized receiver antenna these results must be divided by two.

Considering a lossless path, corresponding to Latm = 1 in Figure 15.5, with an effective aperture of Ae m2 and one polarization, the total power density at the receiver antenna is evaluated as

Upon substituting (15.29) into (15.31), the temperature of the source at the antenna input becomes

where Ωs is the solid angle of the sky‐noise source and Ωa = λ2/Ae is the solid angle of the received antenna beam. Equation (15.32) applies for an antenna using single polarization; with dual polarization this result must be doubled. Also, as indicated, it applies when the solid angle of the source is less than that of the antenna, that is, Ωs < Ωa; when Ωs ≥ Ωa the result is

(15.33) images

If the medium between the radiation source and the antenna exhibits a loss, for example, when Latm > 1 with an effective temperature images, then, using the previous results, the contribution of this external noise source at the antenna input is

15.2.2.2 Contribution of Multiple Sky‐Noise Sources to the Antenna Temperature

When all of the sky noise sources shown in Figure 15.5 are considered, the brightness temperature is evaluated by summing the background temperature with the temperatures of the discrete sources in consideration of overlapping regions as viewed by the antenna beam. To simplify this evaluation it is assumed that, except for the Sun, all of the discrete sources are completely within the antenna beam and do not overlap each other or the Sun. Evaluating spaced‐based cross‐links where the Sun may occupy a significant portion the beam area is a special case. Encounters involving the Moon simply involve replacing the Sun’s noise temperature with that of the Moon. With these considerations the brightness temperature of the noise sources with no atmospheric affects is computed as

(15.35) images

and, by analogy with (15.34), when atmospheric affects are involved this result becomes

The noise power of the Sun varies approximately as the inverse square of the frequency. The piecewise linear dependence of a quiet Sun’s noise temperature on frequency is described by the approximate relationship

where fg is the frequency in GHz. Equation (15.37) is based on the results of Kuiper [6] and Blake [5]. During sunspot activity, the noise temperature of the Sun may be 102–104 times higher than indicated by (15.37) lasting for several seconds followed by temperatures about 10 times higher lasting several hours.

The noise power from cosmic sources, such as our galaxy and interstellar gases, varies approximately as the inverse square of the frequency and the frequency dependence of the noise temperature varies approximately as [7] 1/f2.5. For example, denoting the galactic noise temperature at 100 MHz as T100, the galactic temperature at another frequency is expressed as [8]

(15.38) images

Approximate maximum and minimum values of T100 are given by Brown and Hazard [9] as 18,650 and 500°K with a geometric mean value of 3,050°K. The dependence of cosmic noise sources on frequency is discussed in more depth by Hogg and Mumford [10], Ko [11], Smerd [12], and Strum [13]. The inclusion of other discrete sources depends upon the knowledge of their noise characteristics and location relative to the receiver antenna beam pointing coordinates.

The antenna noise temperature is also a function of the loss through the atmosphere as characterized by the atmospheric noise temperature which is dependent on the elevation or grazing angle of the antenna, the temperature, T(r), and the absorption coefficient, α(r), at a range r from the antenna. The effective temperature of the atmosphere is evaluated using the relationship [14]

A simplified evaluation of (15.39) results by using the average temperature and Gardner [15] has reported that this average temperature is approximately 84% of the Earth’s surface temperature (Tgnd) so that images.

15.2.3 System Noise Figure

The preceding analysis of the receiver and antenna subsystems, in terms of their noise temperatures and the receiver noise figure, was undertaken so that the system noise temperature and noise figure can be established to evaluate the performance of the entire receiver system. If the antenna is connected to the receiver input through the antenna to receiver feed subsystem, then the antenna temperature is transferred to the receiver input using the concepts developed in Section 15.2.1. In this case, the system noise figure at the receiver input is defined in terms of the corresponding system temperature so the sensitivity of the overall receiver performance is easily determined from knowledge of the antenna noise temperature and the antenna feed temperature and loss.

Consider the antenna and receiver system shown in Figure 15.6, where the antenna is connected to the receiver input amplifier or LNA using a coaxial cable or waveguide with a loss Lar and physical temperature Tpar °K. Using the antenna noise temperature (Ta) and the receiver noise figure (Fn), the receiver system temperature Trs and the corresponding system noise figure Fns at the input to the receiver are related as indicated in Figure 15.6. The relationship between Trs and Fns is similar to that expressed by (15.19), that is,

Flow diagram of antenna and receiver interface configuration from antenna noise temperature to receiver feed to demodulator.

FIGURE 15.6 Antenna and receiver interface configuration.

Therefore, upon evaluating the system noise temperature at the receiver input it is a simple matter to compute the system noise figure.

The evaluation of Trs follows the procedures discussed in the preceding section with the result

where the effective noise temperature of the antenna‐to‐receiver feed is evaluated as

The receiver noise temperature is evaluated using (15.23) or, more directly, in terms of the receiver noise figure Fn using (15.22) as

Using (15.43) and (15.42) to compute Trs, the system noise figure is evaluated using (15.40) and is plotted in Figure 15.7 as a function of the feed loss Lar for several antenna temperatures. In this plot a receiver noise figure of 2 dB (1.58:1) is used and the physical temperature of the feed network is assumed to be Tparf = 290°K.

Graph of antenna to receiver feed loss vs. system noise figure illustrating three solid curve plots with labels 100, 200, and 400.

FIGURE 15.7 Example evaluation of system noise figure (Fn = 2 dB, Tpar = 290°K).

The asymptotic value of the system noise figure as Lar approaches infinity is evaluated as images = 4.12 dB so the system noise figure is degraded by the physical temperature of the feed; this is of academic interest because the received signal power also experiences this loss and approaches zero as the feed loss increases. A more informative plot is the loss in the receiver noise figure as a function of the antenna noise temperature; this is shown in Figure 15.8 for various losses in the antenna‐to‐receiver feed with Fn = 2 dB and Tpar = 290°K.

Graph of loss in system noise figure with antenna temperature illustrating five ascending plots with labels 0, 1, 2, 3, and 6 and 2 dotted plots meeting at a certain point.

FIGURE 15.8 Loss in system noise figure with antenna temperature (Fn = 2 dB, Tpar = 290°K).

The receiver noise figure loss is characterized by substituting (15.42) into (15.41) and recognizing that the increase in the receiver temperature relative to Trec is

The resulting loss in the receiver noise figure is defined as

Equation (15.45), expressed in decibels, is plotted in Figure 15.8 with (15.44) substituted for ΔTrec.

A receiver noise figure loss of 2.12 dB is incurred for all feed losses when the antenna temperature is equal to the physical temperature of the feed, that is, Ta = Tpar. Furthermore, as the feed loss increases, ΔTrec approaches Tpar independent of the antenna temperature so the noise figure loss corresponds to a constant loss given by

(15.46) images

This result verifies the system noise figure in Figure 15.7, in that, as images. Figures 15.7 and 15.8 are based on the unique condition that images.

On occasions it is useful to refer the system noise temperature, as computed by (15.41), to the antenna input using the relationship

This section concludes by characterizing the gain temperature ratio (G/T) of the receive antenna under several conditions. The G/T ratio is a measure of the quality of the receive system much like the EIRP is a quality measure of the transmit system. The antenna G/T gain temperature figure of merit is typically expressed in units of dB/°K.

One application of the antenna system noise temperature referred to the antenna is to use the specification images. This definition characterizes the antenna performance under operational conditions including various sky and ground noise sources and is used in the design and evaluation of the antenna subsystem. Another definition is images when Tain = 0. This definition provides a figure of merit for the antenna subsystem with no external noise sources as might be required for a performance acceptance test of the antenna.

15.2.3.1 Example: Evaluation of Antenna Temperature and Receiver G/T and C/No Ratios

Consider an Earth station antenna with a gain of Ga = 55 dB and a spot beam with beamwidth θB = 1° (0.0174 rad) viewing a satellite at a low elevation angle. The antenna does not use a radome; however, the reflector and feed losses are assumed to be Larf = 0.5 dB with a physical temperature of Tparf = 290°K. The Sun, at a temperature of Tsun = 500,000°K, is totally within the satellite beamwidth and is viewed against cosmic background noise at a temperature of Tbck = 40°K. The ground temperature in the vicinity of the antenna is Tgnd = 290°K and Gardner’s approximation to the atmospheric temperature is used. The atmospheric loss is 2 dB, that is, Latm = 1.585, and the lower and upper antenna sidelobe contribution factors to the ground and atmospheric noises are ρlsl = ρusl = 0.1. Furthermore, the antenna is connected to the receiver LNA through a feed with a physical temperature of Tar = 290°K and a loss of 1 dB, that is, Lar = 1.26; this loss results from a coaxial cable with an impedance mismatch. Applying the feed loss, gain, and AGC conditions used in the example in Section 15.2.1.1, with L2 = 3 dB, LNA gain = 20 dB, and an AGC attenuation of 10 dB, the receiver noise figure is 2.5 dB.

Based on these conditions the antenna temperature (Ta), the antenna system noise temperature (Tas), and the receiver G/T ratio are evaluated. Some necessary constants are Sun’s radius is images km and the Sun’s distance from Earth is images km.

The first task is to determine the contribution of the sky noise temperature Tb at the antenna input. The solid angles of the Sun and antenna are

(15.48) images
(15.49) images

where images is the diameter of the beam at a range of Rs; a spot beam has a circular pattern normal to the pointing direction. Because images and Latm = 1.585, the sky and atmospheric noise temperatures at the antenna input are computed using (15.36). In this computation, the effective atmospheric noise temperature, based on the ground noise, is approximated as

(15.50) images

These results are used to compute the background noise temperature at the antenna input as

(15.51) images

The ground and atmospheric noise temperature entering through the antenna sidelobes are images and images where Tatm = 660°K is the computed physical temperature of the atmosphere. From these results and (15.28) the total temperature at the antenna input is Tain = 88,649°K.

Since there is no radome, only the effective antenna reflector and feed temperatures are involved in referring the antenna input temperature to the antenna output. Given the combined reflector and feed loss (Larf) and physical temperature (Tparf), the effective antenna reflector and feed temperature is computed as

Using (15.52) with Tain = 88,469°K and (15.28) with Lrad = 1.0 the antenna temperature is computed as

To determine the antenna system noise temperature the receiver noise temperature is required and, using (15.22) or (15.43) and the specified receiver noise figure, the result is

Using (15.53), (15.54), and (15.41), with a feed loss of 1 dB, temperature Tpar = 290°K, Lar = 1.26 and, from (15.42), images = 75.4°K, the receiver system temperature is evaluated as

(15.55) images

The Sun has significantly degraded the system sensitivity as seen from the increase in the system noise figure Fns = 23.4 dB. The system temperature referenced to the antenna input is computed using (15.47) with the result Tas = 79,400°K and, using this result, the antenna‐receiver system G/Tas ratio is found to be

This example demonstrates the procedures in the evaluation of the antenna G/T ratio as seen at the receiver input; it also demonstrates the horrific impact on the system performance when the antenna is pointed directly at the Sun. Problem 3 examines the impact of the Sun’s temperature when the antenna beam is partially illuminated by the Sun.

The receiver carrier power‐to‐noise density ratio, C/No, is an important receiver parameter, in that, the demodulator Eb/No performance is determined as Eb/No|dB = C/No|dB − 10log10(Rb) where Rb is the user data rate. From the viewpoint of the link budget, Eb/No includes the theoretical value, corresponding to a specified Pbe, and includes the demodulator losses. Based on (15.13) the C/No ratio for a point‐to‐point link is expressed in decibels, using previously defined parameters, as

(15.57) images

where images.

15.2.4 Remarks on the System Noise Figure

Often the minimum received power at the input to the LNA is specified along with a corresponding system performance requirement, for example, the bit‐error probability at a given Eb/No(req’d) signal‐to‐noise ratio. This defines the required receiver sensitivity and allows for the design and development of the receiver and demodulator to proceed without the antenna subsystem. Defining the minimum received power as

(15.58) images

the minimum signal‐to‐noise ratio is evaluated using (15.13) and is expressed as

In this computation Fns is the system noise figure so a specification of the system antenna temperature Ta must be provided. The temperature of the feed or cable required to connect the antenna to the LNA must also be considered in the computation of Fns. With these caveats and the bandwidth expressed in terms of the bit rate Bn = Rb, and No = kToFns, (15.59) specifies the minimum (or theoretical) signal‐to‐noise ratio Eb/No(min) that is achievable based on the receiver sensitivity. Therefore, using these results, the receiver and demodulator design margin is

(15.60) images

If the specified Eb/No corresponds to the theoretical value of Pbe based on the selected waveform modulation and coding, then this margin must include all receiver and demodulator losses and hopefully some to spare. Examples of various receiver and demodulator losses are given in Sections 15.415.10 and the magnitude of the demodulator losses is identified in the simulations discussed in Section 15.14.

15.3 ANTENNA GAIN AND PATTERNS

This discussion of antennas is intended to provide insight into the antenna patterns used in the evaluation of the communication link performance and, as such, draws upon the volumes of literature [3, 16–18] on the subject of antenna theory and performance. Because antennas that produce spot beam radiation patterns are in such widespread use in terrestrial and satellite communications, the circular aperture antenna pattern is of particular interest. The antenna efficiency (ηant) includes the following efficiencies: the aperture efficiency (ηa), the spillover efficiency (ηs) of the reflector, and the antenna reflector surface and feed efficiency (ηarf). The aperture efficiency is defined as the ratio of the effective aperture to the physical aperture, such that, Ae = ηaAa, and is dependent on weighting function w(x) that is applied to the aperture illumination. The weighting function can significantly reduce the sidelobes at the expense of producing a wider beamwidth and reduced aperture efficiency. Uniform aperture weighting results in the highest aperture efficiency and lowest half‐power beamwidth; however, the sidelobes are only 13 dB below the main lobe. The radiation efficiency is defined as the product of the antenna spillover efficiency and the antenna reflector surface and feed efficiency, that is, ηr = ηsηarf. The reflector surface and feed efficiencies are related to the antenna ohmic loss, denoted as Larf. The spillover efficiency is associated with the loss when the antenna feed pattern is not completely concentrated on the reflecting surface. The spillover loss accounts for a large part of the radiated energy behind the reflector. Blocking of the aperture by the antenna feed and support structures lower the gain and raise the sidelobes of the antenna pattern. The offset‐feed antenna configuration eliminates aperture blocking.

In the previous sections, the antenna gain is related to the physical and effective antenna area by the relationships

(15.61) images

The antenna gain is defined in terms of the directive antenna gain by the relationship

(15.62) images

where images and Ωa is defined as the solid angle of the antenna beam, expressed as

(15.63) images

The angles θ and ϕ represent the orthogonal azimuth and elevation 3 dB beamwidth angles of the antenna pattern and the 3 dB beam area is approximated as images. Using this relationship with GD as given earlier, the antenna gain is expressed as

(15.64) images

15.3.1 Rectangular Aperture Antenna Pattern

The antenna pattern for a linear one‐dimensional aperture of length La is characterized in terms of the far‐field electric intensity as

where, in general, W() is the complex aperture weighting function and λ is the wavelength of the radiating frequency. The far field occurs when the range r ≫ λ and is formally characterized by the Fraunhofer region where the received electric field is represented by a plane wave; typically, the far field is considered to be r ≥ images. Under this condition, the phase error at the 3 dB points of a uniformly illuminated aperture antenna corresponds to λ/16 wavelength. Upon defining the spatial frequency as ξ = sin(ϕ)/λ, (15.65) is recognized as the inverse Fourier transform so the weighting function can be determined by taking the Fourier transform of a specified electric field intensity with respect to the differential  = dsin(ϕ)/λ.

Evaluation of the electric field pattern for a uniformly illuminated aperture with constant phase, such that, W() = 1/La, results in the sin(x)/x electric field intensity pattern given by

The magnitude‐squared |E(ϕ)|2 of the electric field intensity results in the pattern of the radiated power that is used to characterize the antenna gain. The solid curve in Figure 15.9 is a plot of (15.66) in terms of |E(ϕ)|2 in decibels and is the familiar sinc2(x) response function.9 The antenna gain is normalized for unit amplitude at ϕ = 0 so the peak gain is 0 dB. The abscissa is plotted as the normalized angle images where Lλ = La/λ; however, the antenna beam angle is ϕ, evaluated as

(15.67) images

where the approximation applies when Lλ ≫ 1, that is, for large apertures.

Graph of normalized radiated power pattern for weighted linear aperture antenna illustrating three curve plots of solid, dashed, and dotted.

FIGURE 15.9 Normalized radiated power pattern for weighted linear (one‐dimensional) aperture antenna.

The cosine weighted aperture function is expressed as

where the factor images results in the same aperture power as in the uniformly weighted case. Using (15.68) in the evaluation of (15.65) and defining Lλ = λ/La results in the electric field intensity pattern

Equation (15.69) is normalized by images and plotted as the dashed curve in Figure 15.9. The response is identical to the previous applications of the cosine weighting function with a first spatial sidelobe level of −23 dB and null at Lλsin(ϕ) = 1.5.

The final antenna aperture to be considered is the triangular aperture with the weighting function given by

(15.70) images

where, in this case, the images results in the same aperture energy as the uniformly weighted aperture. Upon evaluation of (15.65) for the triangular aperture results in the electric field intensity given by

(15.71) images

This result is plotted in Figure 15.9 as the dotted curve.

For the uniformly illuminated aperture antenna, the first spatial sidelobes are 13 dB below the main lobe; the distance between the first nulls, on either side of the main lobe, is λ/La radians; and the 3 dB beamwidth is 0.89λ/La radians (51°). These characteristics are summarized in Table 15.2 along with those of the cosine and triangular weighted apertures.

TABLE 15.2 Antenna Pattern Characteristics for Several Aperture Weighting Functions

Aperture Aperture Weight images Angle Between First Nulls (rad) 3 dB Beamwidth θB (rad) First Sidelobe Level (dB) Gain Loss (dB)
Uniform 1.0 2.0/Lλ 0.89/Lλ −13.26 0.000
Cosine images 3.0/Lλ 1.19/Lλ −23.0 0.912
Triangular images 4.0/Lλ 1.27/Lλ −26.52 1.25

15.3.2 Circular Aperture Antenna Pattern

The circular aperture is evaluated by converting the aperture weighting function to polar coordinates with the result that the electric field intensity pattern is expressed as

Normally the aperture weighting is independent of the angle θ, that is, the illumination amplitude is constant around the aperture for 0 ≤ r ≤ D/2, and (15.72) simplifies to

(15.73) images

where Jo(x) is the zero‐order Bessel function of the first kind. Evaluation of this result for a uniformly illuminated aperture of diameter D results in the expression [18]

(15.74) images

where J1(x) is the first‐order Bessel function of the first kind and images. This result is normalized by πD2/4 to yield unit gain at E(ϕ = 0). The normalized radiated pattern for the uniformly weighted circular antenna is shown as the solid curve in Figure 15.10. This pattern is used in the link evaluation program to determine the ground noise dependence on the antenna elevation angle as discussed in Section 15.2.2. The 3‐dB beamwidth of this antenna is approximately 1.02/Dλ radians (58.4/Dλ degrees) and the first sidelobe is 17.6 dB below the peak gain. The aperture efficiency relative to the uniformly weighted linear aperture is 0.865 (−0.63 dB).

Graph of normalized radiated power pattern for several weighted circular aperture antennas illustrating three curve plots of solid, dashed, and dotted.

FIGURE 15.10 Normalized radiated power pattern for several weighted circular aperture antennas.

Silver [17] has developed expressions for the circular aperture electric field intensities using an aperture weighting function described by

(15.75) images

where p ≥ 0 with the result

(15.76) images

This result is normalized by πD2/4 so the value at E(ϕ = 0) is relative to the uniformly weighted aperture. Table 15.3 summarizes the approximate antenna characteristics for the circular aperture weighting functions considered.

TABLE 15.3 Approximate Antenna Pattern Characteristics for Several Circular Aperture Weighting Functions

Aperture Aperture Weight images Angle Between First Nulls (rad) 3 dB Beamwidth θB (rad) First Sidelobe Level (dB) Gain (dB)
Uniform 1.0 2.42/Dλ 1.02/Dλ −17.6 0.00
Silver, p = 1 images 3.28/Dλ 1.27/Dλ −24.6 −1.25
Silver, p = 2 images 4.02/Dλ 1.48/Dλ −30.6 −2.52

15.4 RAIN LOSS

Signal propagation through rain can result in a significant loss depending on the rain rate, usually measured in millimeter per hour; the system carrier frequency; and the propagation path through the rain region. The propagation path is influenced by the antenna elevation angle. This subject has received considerable attention over the years [19, 20] and has been reduced to several models that characterize the rain attenuation as a function of environmental and operational parameters. The Crane model [21] is the first comprehensive characterization of rain attenuation based on empirical worldwide rain statistics. Arnold and Kao [22] provide a summary of four models*: the Crane model, two CCIR models [23, 24], and the simple attenuation model (SAM) [25]. The link evaluation program discussed in Section 15.15 uses the SAM rain attenuation model with the communication link geometry shown in Figure 15.11. This geometry is used to compute the path length L(R) through the rain region, with rain rate R millimeter per hour, where L(R) is given by

(15.77) images
2 Schematics of communication link geometry used to evaluate rain attenuation with arrows depicting effective rain height, spherical earth radius, earth station, and comm link and labels satellite and expanded view.

FIGURE 15.11 Communication link geometry used to evaluate rain attenuation.

The slant ranges r1 and r2 are shown in Figure 15.11; all distances are in kilometer. This characterization assumes that the slant range between the communication platforms is greater than max(r1, r2) as encountered in satellite links; however, for terrestrial links, that may involve short ranges, L(R) must be computed as images where ro is the link range. The range r1 is the propagation range from the ground station through the rain region to the effective height of the rain and is evaluated using the law of sines as

(15.78) images

where θ is the ground antenna elevation angle, He is the effective rain height, related to the 0°C isothermal height (H) relative to the Earth’s surface, as

(15.79) images

and

(15.80) images

The isothermal height is a function of the latitude of the Earth station and is computed by Crane [26] as

(15.81) images

The range r2 is evaluated as images where E is the empirically derived horizontal extent of the rain region given by [27, 28]

In (15.82), E is in kilometers and p is the rain rate in millimeter per hour. Using the path length L, in kilometers, the attenuation is computed in decibels as

(15.83) images

where the constant λ = 1/14 results in the best fit to the data; R is the rain rate, in millimeter per hour, along the communication path L(R) in the selected rain region; and the constants a and b are frequency‐ and temperature‐dependent constants, approximated as [20]

(15.84) images

and

(15.85) images

The rain rate used by this model is associated with a world rain rate map and is identified by the geographic region of interest using a letter/number designation [29]. Table 15.4 shows the association of the letter designations of the geographic regions. For each geographic region, the selected rain rate is associated with a percent of time (PCT) that the attenuation will exceed the computed value. Table 15.4 also shows a similar correspondence between the PCT designations. For example, in the rain rate region A, a PCT of 0.001% will result in a computed rain attenuation that will be exceeded by only 0.00001T, where T is in units of time representing 1 year [20]. That is, in this example, the attenuation will exceed the computed value no more than 5.24 min in 1 year. Figure 15.12 identifies the rain rate regions in the continental United States, Alaska, and Hawaii. Figure 15.13 shows some examples, based on the SAM model, of the rain attenuation as a function of the elevation angle for an Earth–satellite communication link operating at 10 and 20 GHz. For these plots, the geographic region is D2 and the PCTs, 0.01, 0.1, and 0.5, correspond to respective rain rates of 49, 14.5, and 5.2 mm/h.

TABLE 15.4 Correspondence Between Rain‐Rate Regions and PCTa

PCT Climate Region Rain Rates (mm/h)
A B1 B2 C D1 D2 D3 E F G H
0.001 28.5 45.0 70.0 78.0 90.0 108.0 126.0 165.0 66.0 185.0 253.0
0.002 21.0 34.0 54.0 62.0 72.0 89.0 106.0 144.0 51.0 157.0 220.5
0.005 13.5 22.0 35.0 41.0 50.0 64.5 80.5 118.0 34.0 120.5 178.0
0.01 10.0 15.5 23.5 28.0 35.5 49.0 63.0 98.0 23.0 94.0 147.0
0.02 7.0 11.0 16.0 18.0 24.0 35.0 48.0 78.0 15.0 72.0 119.0
0.05 4.0 6.4 9.5 11.0 14.5 22.0 32.0 52.0 8.3 47.0 86.5
0.1 2.5 4.2 6.1 7.2 9.8 14.5 22.0 35.0 5.2 32.0 64.0
0.2 1.5 2.8 4.0 4.8 6.4 9.5 14.5 21.0 3.1 21.8 43.5
0.5 0.7 1.5 2.3 2.7 3.6 5.2 7.8 10.6 1.4 12.2 22.5
1.0 0.4 1.0 1.5 1.8 2.2 3.0 4.7 6.0 0.7 8.0 12.0
2.0 0.1 0.5 0.8 1.1 1.2 1.5 1.9 2.9 0.2 5.0 5.2
5.0 0.0 0.2 0.3 0.5 0.0 0.0 0.0 0.5 0.0 1.8 1.2

aIppolito [20]. Courtesy of the National Aeronautics and Space Administration (NASA).

Notes: Region B is the average of B1 and B2 rounded up to one decimal place and region D = D2.

Polar (A, dry; B, moderate), temperate (C, maritime; D, continental), subtropical (E, wet; F, arid), tropical (G, moderate; H, wet).

Map of the United States of America illustrating rain-rate climate regions with labels B1, B2, C, D1, D2, D3, E, and F.

FIGURE 15.12 U.S.A. rain‐rate climate regions.

Ippolito [20]. Courtesy of National Aeronautics and Space Administration (NASA).

Graph of elevation angle vs. one-way path attenuation illustrating solid, dashed, and dotted curve plots depicting 0.1, 0.01, and 0.5 PCT, respectively.

FIGURE 15.13 Examples of rain attenuation using the SAM model (ϕ = 45°N, H0 = 0, Region D2).

15.5 ELECTRIC FIELD WAVE POLARIZATION

Orthogonal antenna polarization provides for communication diversity without the need for bandwidth expansion. In general, the polarization of a wave in space is characterized as being elliptical in the plane normal to the direction propagation; linear polarization (LP) and circular polarization (CP) are special cases of elliptical polarization (EP) having widespread applications. Orthogonal linear polarized (LP) antennas transmit vertical polarized (VP) and horizontal polarized (HP) electric fields and CP antennas transmit either right‐hand circular (RHC) or left‐hand circular (LHC) electric fields depending upon parameter selections. The right‐ and left‐hand rules apply to the direction of the circular rotation of the electric field, that is, when the right‐ or left‐hand thumb points in the direction of the propagation, then the curl of the fingers, as viewed from the receiver, indicates either counterclockwise or clockwise rotation, respectively. Under ideal conditions, when the receiver antenna and electric field are polarized in the same manner the maximum signal power is received and the orthogonally polarized electric‐field results in zero cross‐polarization interference. Unfortunately, there are a number of practical issues that work against this ideal behavior, for example, in the case of LP it is difficult to keep the transmitter and receiver antennas exactly aligned and the channel introduces polarization rotation along the propagation path. The most deleterious propagation affects are multipath interference, Faraday rotation, especially at HF and VHF frequencies, and atmospheric polarization resulting from various forms of water content, for example, rain, hail and, to a lesser extent, snow.

The electric fields of a wave propagating in the z‐direction with Poynting vector* images are shown in Figure 15.14 with the time‐ and range‐dependent orthogonal electric fields given by [30]

and

where Mx and My are the magnitudes of the electric fields, ω is angular carrier frequency of the wave, β = 2π/λ is the phase‐per‐carrier frequency wavelength, and δ is the phase of the y‐electric field relative to the x‐electric field. The argument ωt − βz is derived from the delayed carrier ω(t − to) where to = z/c is the range delay and c = velocity of light. The instantaneous electric field intensity, expressed as functions of the parameters t and z, is given by

(15.88) images
XYZ coordinate system depicting wave electric fields along propagation path with labels Ey, Ex, and P.

FIGURE 15.14 Wave electric fields along propagation path.

Kraus [30] shows that (15.86) and (15.87), evaluated at z = 0, can be formulated into the expression of an ellipse given by

The orientation of the ellipse relative to the x, y axes is shown in Figure 15.15.

Diagram of the polarization ellipse, featuring lines and angles.

FIGURE 15.15 Polarization ellipse.

An important parameter in characterizing the polarization is the axial ratio (AR), defined as the ratio of the semimajor to semiminor axes of the ellipse. The polarization ellipse is completely defined by the tilt angle τ, the sense of the rotation of the electric field, and the AR. For the ellipse shown in Figure 15.15 the AR is

(15.90) images

and Kraus has characterized the tilt angle in terms of the parameters Mx, My, and δ as

(15.91) images

This expression for the tilt angle is relative to the x‐axis and applies to either the major or minor axis of the ellipse; the last condition occurs if the ellipse in Figure 15.15 is rotated by π/2 radians.

The parameters Mx, My, and δ are used to characterize LP and CP. For example, referring to (15.86) and (15.87), when δ = 0 and π the linear polarized waves are characterized as shown in Figure 15.16. The tilt angle represents the slope of the polarization. The vectors in Figure 15.16 are actually sinusoidal time and range varying functions as given by (15.86) and (15.87) with magnitudes Mx and My; these magnitudes vary with range based on the free space and other propagation losses. The double tipped vectors suggest the sinusoidal variations. The AR for VP and HP waves is infinite.

Three polarizations viewed toward Poynting vector illustrating linear (left), vertical (middle), and horizontal (right) polarizations.

FIGURE 15.16 Linear polarizations viewed toward the Poynting vector.

Circular polarization occurs when δ = ±π/2 and My = Mx = M in which case (15.86) and (15.87) become

(15.92) images

and

and (15.89) reduces to the equation of a circle, that is,

(15.94) images

The plus and minus sign in (15.93) represents the sense of the CP rotation when viewed toward the Poynting vector. For example, the phase of the composite sinusoidal E‐field wave rotates clockwise as time advances when δ = π/2 corresponding to the plus sign and counterclockwise when δ = −π/2 corresponding to the negative sign. These definitions correspond to those of conventional positive and negative frequencies with respective positive and negative phase advances with increasing time. Applying the right‐ and left‐hand thumb rules, described earlier, δ = π/2 corresponds to left‐hand CP (LHCP) and δ = −π/2 corresponds to right‐hand CP (RHCP). The AR for CP waves is unity.

15.5.1 Antenna Polarization Loss and Isolation

In this section, the analysis of the antenna power loss and the isolation of an orthogonally polarized interfering signal are examined based on the work of Ippolito [31]. In this analysis, the polarization state of the received wave at the input to the receiver antenna, that is, after propagation over the range distance zmax between the transmitter and receiver, is specified together with the receive antenna polarization state. The interaction of the received wave and the antenna is based on their polarization states* and the ARs as characterized by the polarization mismatch factor 0 ≤ mp ≤ 1. For arbitrary EP or LP states of the received wave and the receiver antenna, the polarization mismatch factor is given by

where rx is the AR and x = (w,a) refers to the received wave or antenna, respectively. The sign of the AR is negative for right‐hand polarization and positive for left‐hand polarization. For LP, the sign is always positive. The rotation of the polarization axis is π/2 radians for VP and zero radians for HP and right- and left-hand polarizations. The AR is expressed in decibels as

(15.96) images

The power at the output of the receive antenna is evaluated as

(15.97) images

where PD is the received power density and Ae is the effective area of the receive antenna normal to the Poynting vector. The antenna polarization power loss, in decibels relative to an ideally matched antenna, is expressed as

(15.98) images

The polarization mismatch factor for an elliptically polarized received wave with RHCP copolarization state and a vertically polarized receive antenna is shown in Figure 15.17 for wave ARs rw = −1, −1.5, and −5 (0, 3.52, 40 dB). This example is not representative of an ideal situation, in that, a vertically polarized received wave is required to match the vertically polarized receive antenna. Referring to Figure 15.17, it is seen that a RHCP received wave rw = −1 and |mp(w,a)| = 0.5 results in a power loss of 3 dB and, as rw becomes larger, the minimum loss approaches 0 dB and corresponds to the vertical y‐axis thus matching the vertically polarized receive antenna. An ideal circularly polarized antenna has an AR of ra = 1 (0 dB) and a well‐designed antenna will have an AR of ra = 1.19 (1.5 dB) and an ideal LP antenna has an AR of ra = ∞ (∞ dB) with practical values ranging from ra = 17.8 to 31.6 (25–30 dB).

Graph of the polarization mismatch factor for elliptical wave with RHCP copolarization state and vertical polarized antenna.

FIGURE 15.17 Polarization mismatch factor for elliptical wave with RHCP copolarization state and vertical polarized antenna.

The solid lines in Figure 15.18 show the maximum and minimum polarization losses as a function of the wave AR; the losses corresponding to the polarization mismatch factors in Figure 15.17 shown as the circled data points. These losses are orthogonal to each other with the minimum loss corresponding to the VP or the copolarized state and the maximum loss corresponding to the cross‐polarized state.

Graph of receive wave axial ratio vs. loss depicting polarization losses, with two lines (Lmax and Lmin) having both ends meet at 3 dB, represented by a hollow circle.

FIGURE 15.18 Polarization losses corresponding to Figure 15.17.

The antenna isolation is defined as the ratio of the power at the antenna copolarized output ac and the power at the cross‐polarized output ax. Given the total received signal power density images the isolation is expressed as

Because any wave can be resolved into two orthogonal polarized states, denoted as w and wo, (15.99) is rewritten as

where the second equality is obtained in a straightforward way with the constant xpd defined as the cross‐polarization discrimination, expressed as

Referring to (15.100) an ideal antenna occurs when all of the antenna co‐ and cross‐polarized outputs are equal to the respective co‐ and cross‐polarized inputs, such that, images; this also leads to the conditions images with the result that the isolation for an ideal antenna is

(15.102) images

The antenna mismatch factor, cross‐polarization discrimination, and isolation are examined in the following case study for an elliptical received wave with copolarization state LHCP and a receive antenna with a LHCP polarization state.

15.5.2 Case Study: Polarization Characteristics for a LHCP Antenna

This case study examines the antenna mismatch factor, the corresponding polarization loss, the cross‐polarization discrimination, and the isolation for a LHCP antenna with an elliptical received wave with a copolarization state that is matched to the receive antenna. The antenna is not ideal so the isolation asymptotically approaches the cross‐polarization discrimination function xfd for low values of xfd.

The antenna mismatch factor is evaluated using (15.95) for a specified ra and parametric values of rw with τa = 0 radians and variable angular displacements τw = θ*: 0 ≤ θ ≤ 2π radians. The results are plotted in Figures 15.19 and 15.20 corresponding to ra = 1 and 1.5, respectively. With ra = 1 the antenna mismatch factor appears as circles with radius ≤1. The case having unit radius corresponds the received wave being matched to the antenna with rw = ra = 1; the other cases result in a mismatched antenna and lower values of mp corresponding to higher polarization losses. The results in Figure 15.20 represent a nonideal antenna with ra = 1.5 and, although the mismatch factor for rw = 1 is a circle it does not have unit magnitude and results in a polarization loss. The case with rw = ra = 1.5 is matched in the horizontal direction with increasing loss in the vertical direction. As the parameter rw continues to increase the antenna mismatch factor decreases along both axes resulting in increasing polarization loss. The salient point is that the antenna mismatch factor results in zero loss when the received wave AR is equal to the antenna AR. The losses are shown in Figure 15.21 as a function of rw, expressed in decibels, for ra = 1 and 1.5. The circled data points represent the losses for the corresponding ra and rw conditions in the range of the abscissa.

Polar graph of antenna mismatch factor for LHCP antenna (ra= 1) and elliptical received wave with LHCP copolarization.

FIGURE 15.19 Antenna mismatch factor for LHCP antenna (ra = 1) and elliptical received wave with LHCP copolarization.

Polar graph of antenna mismatch factor for elliptical antenna (ra= 1.5) and elliptical received wave with LHCP copolarization.

FIGURE 15.20 Antenna mismatch factor for elliptical antenna (ra = 1.5) and elliptical received wave with LHCP copolarization.

Graph of wave axial ratio vs. minimum loss for minimum antenna mismatch factor losses, with solid, dashed, dotted and dash-dot curves for 0, 3.52, 14.0 and 26.0 dB and 1.0, 1.5, 5.0 and 20.0 linear, respectively.

FIGURE 15.21 Minimum antenna mismatch factor losses.

Evaluation of the isolation of an ideal LHCP antenna involves evaluating the cross‐polarization discrimination xpd as expressed in (15.101) as a function of the wave axial rw. The numerator of (15.101) is the copolarized antenna mismatch factor corresponding to wave AR ra = 1 and the denominator corresponds to the cross‐polarized or orthogonal antenna mismatch factor with ra = −1. In these evaluations, the antenna mismatch factors are independent of τa and τw and the resulting ideal antenna isolation is plotted in Figure 15.22 with the xpd(dB) = 10log10(xpd) and the abscissa plotted in terms of the common logarithm as 20log10(rw). From this result it is seen that the isolation increases as the wave AR approaches 0 dB or rw = 1. The value of xpd, plotted in Figure 15.22, is used to evaluate the isolation of a nonideal antenna using (15.100). In evaluating (15.100), the choice of the antenna copolarized and the cross‐polarized ARs rc and rx are based on the antenna design; the nonideal antenna isolation plotted in Figure 15.23 corresponds to rc = 1.122 (1 dB) and various values of rx in decibels. The selected values of the nonideal antenna ARs result in constant values of the antenna mismatch factors: mp(w, ac), mp(wo, ax), mp(w, ax), and mp(wo, ac). These factors influence the asymptotic convergence of the antenna isolation relative to that of the ideal antenna isolation. As rx approaches 0 dB in Figure 15.23, the isolation is ultimately limited by the selection of rc. The value of rc = 1.122 results in a copolarization loss of 0.014 dB and a cross‐polarization loss of 48.8 dB.

Graph of receive wave axial ratio vs. xpd depicting LHCP cross-polarization discrimination, with a negative slope line.

FIGURE 15.22 LHCP cross‐polarization discrimination.

Graph of cross-polarization isolation vs. isolation depicting LHCP polarization isolation, with 8 positive slope lines for rx (db) of 0, 0.1, 0.3, 0.5, 1, 1.5, 3, and 5.

FIGURE 15.23 LHCP polarization isolation (copolarized axial ratio rc = 1.0).

15.6 PHASE‐NOISE LOSS

Phase noise results from intrinsic noise in the transceiver and modem signal sources typically from oscillators and frequency synthesizers used for frequency translation. The phase noise is characterized in terms of the phase‐noise power spectral density (PSD) in dBc/Hz where the decibel level is relative to the signal carrier power. Typically the specification applies to all oscillators in a cascade of subsystems; for example, in a cascade of two subsystems with equal phase‐noise specifications the overall phase noise is increased by 3 dBc/Hz. For higher order modulations, like multiphase shift keying (MPSK) and quadrature amplitude modulation (QAM), the oscillator phase noise is a dominant source of performance loss over the AWGN channel and results in an irreducible error probability with increasing Eb/No.

There is a wealth of information [32–38] on the theory, design, and applications of oscillators with an emphasis on understanding and minimizing the impact of phase noise on communication systems. The phase‐noise variance is the principal performance measure and is defined as the integral of the phase‐noise PSD. In the following sections, the characteristics of the phase‐noise PSD are reviewed and, based on an acceptable phase‐noise variance specification, the subsystem phase‐noise density is established. In Section 15.6.2 the phase‐noise variance is determined from a specification of the spectral density and the resulting performance loss is given for the MPSK‐modulated waveform.

15.6.1 Phase‐Noise Characterization

Oscillator phase noise [37, 39, 40] is characterized in terms of a log–log plot of the relative phase‐noise* PSD (S(f)) as a function of the frequency deviation f from the oscillator frequency fo. The power density N is defined relative to the oscillator carrier frequency power C and plotted with the ordinate specified as 10log(N/C) dBc/Hz. The PSD is symmetrical about fo and, normally, only the positive frequency portion is shown. A generic plot of a phase‐noise PSD specification Ss(f) is shown in Figure 15.24 indicating distinct line segments with frequency dependence km/f m corresponding to amplitude roll‐offs of 10 m in decibels per decade. The parameter km is a scaling factor that determines the dBc/Hz level for each segment. The lowest frequency contained in Ss(f) is typically 1–10 Hz because accurate measurements below 1 Hz are difficult to obtain and the phase noise in these ranges is generally removed in the demodulator by the phaselock loop (PLL) filter. The upper frequency of interest in the phase‐noise PSD is determined by the intermediate frequency (IF) filter bandwidth and ultimately by the symbol matched filter bandwidth.

Graph of log-frequency vs. spectral density depicting generic phase-noise spectral density, with a negative slope curve labeled (from top to bottom) k3/f 3, k2/f 2, k1/f and k0; and dashed line labeled |T(f)|2.

FIGURE 15.24 Generic phase‐noise spectral density.

Barnes [33] characterizes a low frequency segment for m = 4; however, this is seldom considered in applications involving a demodulator PLL. The phase noise corresponding to m = 3 is referred to as flicker noise and is related to abrupt phase changes associated with the oscillators inherent feedback circuit as it attempts to control the phase of the oscillators carrier frequency. The phase noise associated with m = 2 results from white and flicker noise sources within the oscillator. The phase noise associated with m = 1 is high frequency flicker noise and influences the m = 3 and m = 2 phase‐noise segments. The phase noise associated with m = 0 results from the receiver white noise density No and is the underlying noise term associated with the demodulators Eb/No ratio.

Typically, a coherent communication demodulator applies the received carrier‐modulated waveform to a PLL for phase and frequency tracking [41] that also tracks and removes the lower frequency phase‐noise terms that are within the loop bandwidth BL. The PLL closed‐loop frequency response is characterized by the low‐pass function H(f) and the output phase noise results from the equivalent high‐pass response given by*

(15.103) images

In the following analysis, the closed‐loop response for a second‐order loop, with damping factor images, is used resulting in

Referring to (10.38) the natural resonant frequency of the closed‐loop response is related to the loop bandwidth BL as

The response |T(f)|2 is shown as the dotted curve in Figure 15.24 with a band‐reject frequency of BL ≪ Rs Hz. Therefore, for low symbol rate modems the phase noise possesses a more severe performance issue. Typical values of the loop bandwidth are BL = Rs/10 and Rs/100 for binary phase shift keying (BPSK) and quadrature phase shift keying (QPSK) modulations, respectively.

The variance of the phase noise is computed as

where the individual noise variances are computed as

The integral in (15.107) is evaluated for m = 3 and 2 using the integral formula given by Gradshteyn and Ryzhik [42] and using (15.105), with fn evaluated in terms of the symbol rate and the time‐bandwidth product images, the noise variances are evaluated as

(15.109) images

The integrals involving m = 1 and 0 with an infinite integration limit result in infinite variances so the integration must be performed over a finite bandwidth B ≥ Rs/2 Hz where the equality corresponds to the Nyquist bandwidth. In this evaluation the integral in (15.107) is defined as

with the integration limit B = Rs corresponding to the bandwidth of an integrate‐and‐dump (I&D) symbol detection filter. Therefore, for m = 1 and 0, (15.110) is evaluated using Mathcad symbolic processing and the results are expressed as

and

Using (15.105), the dependence on the normalized bandwidth B/Rs for images is obtained by substituting

(15.113) images

into Equations (15.111) and (15.112). The results are plotted in Figure 15.25 as a function of B/Rs for images and images. The numerical solutions to (15.110) for B/Rs = 0.5, 1, 4, and 8 are listed in Table 15.5.

Graph of normalized integration bandwidth vs. integral depicting integration results from m = 1 and 0 with ρL = 0.01 (dashed) and 0.1 (solid).

FIGURE 15.25 Integration results for m = 1 and 0.

TABLE 15.5 Selected Integration Values Im(B/fn) for m = 1 and 0

B/Rs ρL = 0.1 ρL = 0.01
m = 1 m = 0 m = 1 m = 0
0.5 2.8125 16.5915 5.1151 166.508
1 3.5056 33.2728 5.8082 333.025
4 4.8919 133.204 7.1945 1332.11
8 5.5851 266.418 7.8877 2664.22

Using these results, the phase‐noise variances for m = 1 and 0 are expressed as

and

The unknown parameters in these relationships are the scale factors km; however, with knowledge of the breakpoints and the additive constant phase‐noise PSD, the piecewise linear PSD shown in Figure 15.24 can be constructed. In the following section, these results are applied to a system specification of the phase noise and the results are used to evaluate the impact on the bit‐error probability of the communication link. The resulting performance loss is then used in establishing an overall link budget which is essentially the purpose of this chapter.

15.6.2 Phase‐Noise Evaluation Using System Specifications

The phase noise is frequently specified as part of an overall system requirements specification as in Figure 15.26.* The specification discussed in this section is intended to represent the demodulator heterodyning to baseband and is 10 dB lower than that specified for the receive terminal which has multiple heterodyning stages. Because the received and demodulator subsystems are typically developed by different contractors, separate subsystem phase‐noise specifications are required to provide for independent subsystem testing. The demodulator phase noise is intended not to impact the more sensitive phase noise of the receiver oscillators and frequency conversion stages.

Graph of log-frequency vs. spectral density depicting phase-noise spectral density specifications, with 2 negative slope solid and dashed curves labeled receiver and demodulator, respectively.

FIGURE 15.26 Phase‐noise spectral density specifications.

Defense Information Systems Agency (DISA) [43]. Courtesy of U.S.A. Department of Defense (DOD).

These specifications apply to the 1/f3 and 1/f frequency dependencies so the composite linear spectral density is expressed as

The constants km, required for the evaluation of (15.116), are determined from the specifications as follows. Considering the specification shown in Figure 15.26 and repeated in Figure 15.27 as the solid lines; each linear segment

(15.117) images

is plotted in terms of the common logarithm

where images and images are used as the abscissa in the spectral density plots. Therefore, using these results with m = 3, k3 is evaluated by solving for K3 in (15.118) under the condition F = 1. The corresponding values of L3 and K3 are −32 dBc/Hz and −2 dBc‐Hz2; in general Km has units of dBc‐Hzm−1. In a similar manner K1 is found to be −42 dBc and K0 = −112 dBc/Hz corresponding to the constant white noise spectral density. The corresponding values of km are as follows: images, images, and images. The composite phase‐noise PSD, shown in Figure 15.27 as the dashed curve, is a plot of (15.116) using the computed values of km. The dotted curves represent the computed line segments based on Km plotted over the range of abscissa values 1 ≤ F ≤ 8 with portions of these line segments corresponding to those of the specification.

Graph of log-frequency vs. spectral density depicting phase-noise spectral density characteristics for receiver subsystem, with negative slope curve labeled k3/f 3, k1/f and k0; and dashed curve 10log(S(f)).

FIGURE 15.27 Phase‐noise spectral density characteristics for receiver subsystem.

Referring to (15.116) and evaluating the phase‐noise variances images in (15.108), (15.114), and (15.115) with the values of km computed earlier for the receiver specification, the total untracked phase‐noise variance images is evaluated using (15.106) with images. The resulting untracked phase noise is plotted in Figure 15.28 with images and 0.01 for the indicated normalized bandwidth ratios B/Rs. Figure 15.29 shows the untracked phase‐noise variance when B = Rs corresponding to the bandwidth of the I&D symbol detection filter.

Graph of symbol rate vs. variance depicting total phase-noise variance for receiver subsystem, with solid and dashed curves representing ρL = 0.1 and 0.01, respectively.

FIGURE 15.28 Total phase‐noise variance for receiver subsystem.

Graph of symbol rate vs. variance depicting total phase-noise variance at I&D MF output for receiver subsystem (B/Rs = 1), with 2 curves labeled ρL = 0.1 and 0.01.

FIGURE 15.29 Total phase‐noise variance at I&D MF output for receiver subsystem (B/Rs = 1).

15.6.3 Case Study: BPSK and QPSK Performance with Phase Noise

The performance of BPSK and QPSK waveform modulations is examined using the phase‐noise characteristics developed in the preceding section for the receiver subsystem phase‐noise PSD. The modulated waveforms are generated using the rect(t/T) weighting functions and detected using an I&D matched filter. The upper integration limit in the evaluation of the phase‐noise variance images is B = Rs, where Rs corresponds to the noise bandwidth of the demodulator matched filter. The lower integration limit is determined by the high‐pass PLL transfer function T(f) expressed in (15.104). The low‐pass band‐reject bandwidth of T(f) is dependent on the loop time‐bandwidth product images where BL is the closed‐loop noise bandwidth. For BPSK and QPSK modulation, typical values* of ρL are 0.1 and 0.01, respectively; these values are used in the following performance evaluations.

In addition to the phase‐noise variance images resulting from the receiver heterodyning operations, the PLL introduces phase jitter that is characterized by the phase variance images expressed as [44]

(15.119) images

where γL is the signal‐to‐noise ratio in the PLL bandwidth and γb = Eb/No is the signal‐to‐noise ratio in the matched filter bandwidth of Rs hertz. From this discussion two factors are in play that influence the performance of the BPSK and QPSK modulations. The first is that, for a given symbol rate, the lower integration limit in determining the phase‐noise variance is decreased by ρL. For a given bit rate, this is exacerbated with QPSK modulation since Rs = Rb/2. These factors become less significant as the data rate is increased and the phase noise is eventually influenced solely by the oscillator white noise density corresponding to k0. The second factor influencing the performance is that the PLL phase noise is decreased by ρL with the advantage going to QPSK. In consideration of these issues, the total phase‐noise power measured in the matched filter bandwidth is given by

(15.120) images

Because the oscillator phase is influenced by several phase‐noise sources, as indicated by terms giving rise to the km/fm spectral density response, the phase‐noise random variable φ is generally characterized by the Gaussian distribution

This approximation and the limits require that images. The bit‐error probability performance conditioned on the phase error φ for BPSK and QPSK is expressed, respectively, as

and

Using (15.121) with either (15.122) or (15.123) the resulting bit‐error probability is expressed as

The bit‐error probability expressed in (15.124) is evaluated using a 96‐term Gauss‐quadrature integration and the results are depicted in Figure 15.30. The robustness of BPSK modulation is evidenced by a loss of less than 0.1 dB even for data rates as low as 150 bps. For the 1024 kbps data rate, the QPSK modulation has a maximum loss of about 0.1 dB; however, the sensitivity to the phase noise is evident at 150 bps with a maximum loss of about 0.3 dB for the range of signal‐to‐noise ratios considered.

Graph of signal-to-noise ratio vs. bit-error probability of BPSK and QPSK modulations using the receiver subsystem phase-noise specification, with dashed (BPSK), solid (QPSK) and dotted (antipodal) curves.

FIGURE 15.30 Performance of BPSK and QPSK modulations using the receiver subsystem phase‐noise specification.

Figure 15.31 depicts the performance sensitivity when the phase noise results from the cascade of three subsystems: a transmitter, satellite repeater, and receiver, each with a phase‐noise specification corresponding to receiver subsystem. In effect, the phase noise is increased by 4.77 dB from that used in Figure 15.30. In this case, the BPSK performance for the 1024 kbps data rate is essentially unchanged; however, the QPSK performance is degraded by about 0.2 dB at Pbe = 10−5 for 1024 kbps (512 ksps) and by about 0.5 dB for 150 bps (75 sps).

Graph of signal-to-noise ratio vs. bit-error probability of BPSK and QPSK modulations using a cascade of three subsystems, with dashed (BPSK), solid (QPSK), and dotted (antipodal) curves.

FIGURE 15.31 Performance of BPSK and QPSK modulations using a cascade of three subsystems corresponding to the receiver phase‐noise specification.

This procedure can be applied to higher order MPSK modulations (M > 4) with anticipated increases in the performance loss. In this regard, Baker [45] has analyzed the performance of 256‐ary and 1024‐ary QAM with phase noise in terms of the standard deviation σφ of the phase and, for images, the losses at images are about 0.6 and 5.5 dB, respectively.

15.7 SCINTILLATION LOSS

The losses due to signal scintillation are examined in detail in Chapter 20 and the results in a natural environment are summarized in Table 15.6 in terms of the loss factor Lf. The loss factor allows for determining the loss at any carrier frequency f > 2 MHz* according to

(15.125) images

where fMHz is carrier frequency expressed in MHz. The dependence of the loss on latitude results from the different electron density profiles in the three regions corresponding to equatorial (±15°), mid to low (15°–60°), and polar (>60°) latitudes. The loss factors correspond to the worst‐case mean electron densities and variations under turbulent conditions in the natural environment. For example, the loss corresponding to a 90% confidence level at 500 MHz in the equatorial region is determined as La(500) = 2.85e5/(500)2 = 1.14 dB. This example is shown in Figure 15.32 where the transitions between the latitude regions are plotted as 5° centered on the transitions at 15° and 60°. Figure 15.33 shows the scintillation losses at 250 MHz for the indicated confidence levels.

TABLE 15.6 Absorption Loss Factors Lf (dB) in Natural Environment

Confidence (%) Latitude
Equatorial Mid to Low Polar
50a 5.75e4 9.50e3 5.100e4
90 2.85e5 1.08e4 1.957e5
95 3.50e5 1.11e4 2.375e5
99 4.72e5 1.18e4 3.143e5

aMean or average electron densities.

Graph of latitude vs. loss displaying a curve along the equatorial (1.14 dB), mid-to-low (0.04 dB), and polar (0.78 dB).

FIGURE 15.32 Scintillation loss in natural environment (90% confidence at 500 MHz).

Graph of latitude vs. loss depicting scintillation loss in natural environment, illustrating 4 curves with confidence percentage of 99, 95, 90, and 50.

FIGURE 15.33 Scintillation loss in natural environment (250 MHz).

15.8 MULTIPATH LOSS

The maximum and average single‐reflection multipath loss is defined in terms of the multipath factor and the results are plotted in Figure 15.34 for a low Earth orbit (LEO) 200 km altitude satellite receiver and a ground station with antenna height of 30 m. The ground station and satellite use uniformly weighted antennas; the satellite antenna gain is 10 dB and the transmitter antenna gain is varied as indicated in Figure 15.34. The description of satellite link encounter with example multipath losses is given in Section 19.4. The impact of the antenna gains is evident, in that, multipath losses greater than 1 dB occur with transmit antenna elevation angles less than about 14°, 4°, and 1.5° for respective antenna gains of 20, 30, and 40 dB. These gains correspond to 3‐dB transmitter antenna beam widths of 16.2°, 5.2°, and 1.64°. It is assumed that the ground station and satellite beams are ideally tracking each other along the line of sight (LOS) or direct signal path. A word of caution is in order regarding the inclusion of the multipath losses shown in Figure 15.34 directly into the link budget. The reason is that waveform designs and signal processing techniques can effectively mitigate the multipath losses. These techniques include forward error control (FEC) coding and interleaving, optimal combining techniques with data repetition, data equalization, and adaptive antenna null steering.

Graph of Tx antenna elevation vs. loss illustrating three sets of curves for 20, 30, and 40 Gt(dB) with solid and dotted lines representing maximum and average loss, respectively.

FIGURE 15.34 Single‐reflection multipath loss.

15.9 INTERFACE MISMATCH LOSS

The losses associated with the termination of connecting cables and devices within a transmitter or receiver are evaluated in terms of the voltage standing wave ratio (VSWR) or the voltage reflection coefficient (ρr) as measured at the interface between the components. These losses must be added to the measured or specified device losses that are normally determined under ideal termination conditions.

The VSWR is defined as [46] the ratio of the maximum‐to‐minimum voltage of the standing wave expressed as

(15.126) images

The standing wave ratio (SWR) is simply the VSWR expressed in decibels, that is,

(15.127) images

The voltage reflection coefficient is defined as [46] the ratio of the reflected‐to‐incident voltages across the termination or load impedance ZL and is expressed as

where Zo is the characteristic impedance of the cable or the source impedance of the device. In general, ρr is a complex quantity, such that, images.

The relationships between VSWR and ρr are:

(15.129) images

and

(15.130) images

The return loss is defined as

Defining the incident power as images and the reflected power as images the power delivered to the load is images and, using these relationships together with (15.128) and (15.131), the loss at the termination is computed as

(15.132) images

The loss Lt is typically included in the device losses in the computation of the receiver noise figure discussed in Section 15.2.1.

15.10 MISCELLANEOUS SYSTEM LOSSES

If the component loss of a device at a frequency f1 is known to be L1 then the loss at another frequency can often be approximated simply by frequency scaling. For example, the loss at frequency f2 is approximately

(15.133) images

15.10.1 Antenna Shaping Loss

Antenna shaping loss is based on aperture weighting function and is included in antenna gain.

15.10.2 Antenna Scallop Loss

Antenna scallop loss is associated with antenna beam loss during spatial antenna acquisition. Site loss is the antenna gain loss at a site location removed from the center of the beam. For Earth coverage fixed satellite beams, the site loss is compensated through beam shaping to ensure equal returns from all locations.

15.10.3 Frequency Scallop Loss

Frequency scallop losses are associated with frequency acquisition using discrete frequency scanning or a Fourier transform.

15.10.4 Signal Processing Loss

Signal processing losses are associated with discrete amplitude and time sampled quantization losses. Discrete amplitude sampling losses are minimized by using the maximum number of bits to describe the signal amplitude at various points along the processing path and the discrete‐time sampling losses are minimized through filter design and sampling frequency selection. Other sources of signal processing loss are associated with the limitation of various algorithms that approximate theoretical models; for example, the computation of decision thresholds based on a limited sample size.

15.11 NONLINEAR POWER AMPLIFIER ANALYSIS AND SIMULATION

The following discussions and analysis refer to an amplitude‐modulated (AM) and/or phase‐modulated (PM) carrier signal. When a continuous wave (CW) carrier is applied to a nonlinear power amplifier (PA) the output will contain AM and possibly AM and PM distortion which is the subject of this section [47]. The intermodulation noise produced by a TWTA or SSPA nonlinearity is related to the input AM carrier signal level and the AM‐to‐AM (AM‐AM) and AM‐to‐PM (AM‐PM) characteristics of the amplifier. The operating point, or input drive level, of a TWTA is relative to the saturation level. For a SSPA, the operating point is relative to the 1 dB gain compression level. The input backoff (IBO) corresponds to the input power backoff from the saturation or the 1 dB gain compression powers of the respective devices required to achieve the specified system performance. As IBO is increased the devices approach the linear operating region resulting in decreasing levels of intermodulation noise and improved system performance. However, lowering the input power level also results in less efficient use of the overall power capabilities of the devices. Constant envelope‐modulated waveforms, like BPSK and the various forms of QPSK, are more tolerant to lower IBO levels than waveforms with inherently large peak‐to‐rms levels, like QAM, frequency division multiplex (FDM), and orthogonal frequency division multiplexing (OFDM).

The intermodulation noise is characterized in terms of the carrier‐to‐intermodulation noise (C/I) ratio or as the carrier‐to‐intermodulation noise density (C/Io) ratio. The intermodulation noise density is related to the intermodulation noise by the users channel bandwidth B Hz as Io = I/B W/Hz. The C/I performance parameter is discussed in the following sections in the context of the device AM‐AM and AM‐PM characteristics. In addition to the single‐channel intermodulation noise, the nonlinear PA output will include the signals of other users in frequency division multiple access (FDMA) applications. This situation occurs, for example, when the PA is operating as a multichannel satellite relay downlink transmitter, in which case, the receiver and intermodulation noise from all of the channels influences the amplifier output level. The desired signal will include some degree of signal suppression depending upon the severity of the nonlinearity.

As a consequence of the nonlinear device transfer function and the various noise terms, the output backoff (OBO) conditions are defined in terms of the level of the desired signal saturation point or the 1 dB gain compression point as described earlier. The following example focuses on the saturation point of the TWTA, for which, the OBO is characterized as

The parameter Ptot in (15.134) is the total power into the amplifier and Gtot is the corresponding gain. The gain Gtot is the input‐to‐output signal gain of the TWTA as determined by the input signal operating point defined by the IBO level. In the following sections, several methods of evaluating the intermodulation noise are examined and compared. These comparisons are characterized by C/I and OBO and their dependence on the input operating levels as defined by IBO.

15.11.1 Characterizing the TWTA Transfer Function

The characteristics of three TWTAs are examined with varying degrees of nonlinear signal distortion as identified by Saleh [48, 49]. The characteristics of interests are the AM‐PM transfer function in which the AM carrier input signal results in a nonlinear PM output signal; and the AM‐AM transfer function in which the AM carrier input signal results in a nonlinear AM output signal. These nonlinear transfer functions result in varying degrees of harmonic distortion at the output of the TWTA that cause co‐channel and adjacent channel interference (ACI). For clarity the three TWTAs, characterized by Saleh, are denoted as: TWTA No. 0 that is the least severe with zero AM‐PM phase distortion; TWTA No. 1 that exhibits a moderate AM‐PM phase distortion; and TWTA No. 2 that has severe AM‐PM phase distortion. All three result in nearly identical nonlinear AM‐AM distortion.

The comparison of the three traveling wave tube amplifiers is based on Saleh’s two‐parameter in‐phase and quadrature (I/Q) (or real and imaginary) curve‐fit functions expressed as

where αp, βp, αq, and βq are the curve fitting coefficients given in Table 15.7 for the three representative traveling wave tube (TWT) characteristics. The parameter r is the composite input signal level and determines the operating point on the TWT transfer function. In terms of the amplitude and phase responses, the corresponding I/Q components are expressed as images and images.

TABLE 15.7 Two‐Parameter Curve‐Fit Values from Saleha

TWT P(r) Q(r)
αp βp αq βq
0 2.0 1.0 0.0 0.0
1 1.90947 1.07469 4.35023 2.33525
2 2.11075 2.22764 7.33959 2.11475

aSaleh [49]. Reproduced by permission of the IEEE.

Plots of the curve fit results for the quadrature functions expressed in (15.135) are converted to the amplitude and phase functions and shown in Figure 15.35 using the parameter values listed in Table 15.7.

Graph of input vs. normalized output depicting normalized amplitude and phase characteristics illustrating solid and dashed curves representing volts and degrees, respectively.

FIGURE 15.35 Normalized amplitude and phase characteristics.

Saleh [49]. Reproduced by permission of the IEEE.

15.11.2 Evaluation of C/I and OBO

The evaluation of C/I and OBO discussed in this section is applied to a FDMA satellite communication system and follows the work of Saleh [49] using the I/Q responses for TWTs 0, 1, and 2 discussed earlier. The input signal is considered to be the sum of n independent phase‐modulated signals expressed as

(15.136) images

where ωi represents carriers in different FDMA bandwidths (B) occupying a total bandwidth of W = nB Hz. The output of the TWTA when operating in a FDMA network [50–52] is composed of distinct intermodulation tones with angular frequency images; the restriction images ensures that all of the intermodulation products (IMPs) fall in the first spectral zone of each frequency band. The order of the IMP is given by images, that is, only odd order IMPs are present. The value m = 1 represents the desired carrier term at the output of the nonlinear amplifier and all of the other terms, m = 3, 5, … are representative of output distortion tones in the users bandwidth. In the following analysis, the input carriers are assumed to have equal amplitudes with Vi = V :∀i and n ≫ m.

Based on this brief introduction, the C/I ratio at the output of the TWTA is given by

(15.137) images

where v = 0 corresponds to the C/I ratio at the center of the users band and images is the ratio at either edge of the users band. The parameter Nm,n(v) is the number of dominant m‐th order IMPs falling at location v and is given by [53]

Where images for images and images for images. The terms Pm,n and Qm,n represent the I/Q components of the m‐th order IMP, for which, Saleh developed the expressions

where

(15.141) images

and

(15.142) images
(15.143) images
(15.144) images
(15.145) images

with images for images and images. The parameter U is equal to the square root of the total average input power and the function E1(z) is the exponential integral [54]

(15.146) images

Saleh [49] provides an appendix that outlines a method for the numerical computation of (15.139) and (15.140) that avoids computational roundoff errors for large values of m. The solution involves rewriting the respective summations in the brackets { } of (15.139) and (15.140) as follows:

(15.147) images

and

(15.148) images

with the respective limit on m ≌ 82/X + 4.4 and 96/Y − 2.6. The F(z) is expressed as

(15.149) images

and can be evaluated numerically using Laguerre–Gauss quadrature integration.

The OBO is defined as

(15.150) images

where P(r) and Q(r) are the nonlinear output values corresponding to the operating point r of the input composite signal. These values are computed as given in Section 15.11.1 using the appropriate αp, βp, αq, and βq values from Table 15.7.

These relationships are evaluated for the TWTs 0, 1, and 2 and the results are shown in Figure 15.36 which relate the OBO dependence on C/I for the indicated IBO operating points. As the IBO increases the TWTs operate more in the linear region and away from saturation resulting in a higher C/I ratio. Figure 15.37 characterizes the OBO level as a function of the IBO for the three TWTs under consideration.

Graph of carrier-to-interference level vs. output backoff depicting C/I and OBO performance corresponding to center of the band (ν = 0), with 3 ascending curves with TWT of 2, 1, and 0.

FIGURE 15.36 C/I and OBO performance corresponding to center of the band (ν = 0).

Saleh [49]. Reproduced by permission of the IEEE.

Graph of input backoff vs. output backoff displaying 3 ascending curves with TWT of 1, 2, and 0.

FIGURE 15.37 OBO vs. IBO performance corresponding to the center of the band (v = 0).

15.12 COMPUTER MODELING OF TWTA AND SSPA NONLINEARITIES

The nonlinear amplifier is generally characterized in terms of the devices AM‐AM and AM‐PM transfer functions. These transfer functions are typically available from the manufacturer as the measured output power and phase vs. the input power or the measured output voltage and phase vs. the input voltage. The manufacturers may also provide information on the bandwidth and temperature sensitivities.

The following examples involve the modeling of two SSPAs having slightly different transfer characteristics. These models serve to highlight some significant performance differences concerning the specified output power and the OBO required to achieve a specified performance measure. The performance measure may be the spectral regrowth or the degradation in the bit‐error probability given as adjacent channel. This latter situation can also be characterized in terms of the ACI specification. Although these examples focus on the modeling of SSPAs, the techniques apply equally well to TWTAs with the primary exception being the definition of the maximum output power. The maximum output power of a SSPA is defined as the power corresponding to the 1‐dB gain compression point, referred to as P1, whereas, for the TWTA the maximum output power is the saturation power, referred to as Psat. SSPA and TWTA amplifiers are referred to as soft limiters and are contrasted with the hard limiters discussed in Section 10.2. Soft limiters are characterized as having an output that is differentiable with respect to the input and approaches a constant value or limit as the input increases. The soft limiter response is modeled by various forms of sigmoid functions of which the error and arctan functions are commonly used examples.

15.12.1 SSPA with Soft Saturation Response

In this example, the SSPA transfer function is characterized as having AM‐AM that is characterized in terms of the peak output voltage (Vpo) as a function of the peak input voltage (Vpi). Because the AM‐PM response is zero the SSPA is considered to be ideal. The dashed curve in Figure 15.38 shows the measured voltage transfer function and the solid curve is based on a seventh‐order polynomial that is curve fit to the Vpo vs. Vpi response. The modified response results in a constant gain over the linear range of the SSPA and a monotonically decreasing gain as the input signal level increases through the nonlinear range. The actual gain of the device is not important insofar as the performance simulation is concerned so the gain of the soft saturation response is purposely low to avoid dealing with large signal levels in the simulation program. The curve‐fit transfer function is expressed as

with the corresponding gain evaluated as 20log10(Vpo/Vpi). Using (15.151), the third‐order IMPs are characterized by the cubic term as

(15.152) images
Graph of peak input volts vs. peak output volts vs. gain depicting characteristics for soft saturation response SSPA, with descending curve and solid and dashed ascending curves.

FIGURE 15.38 Characteristics for soft saturation response SSPA.

With the input power expressed as images, Figure 15.39 shows the input–output power characteristics including the third‐order intermodulation power. The term images is the output power corresponding to the 1‐dB gain compression identified in Figure 15.38, images is the primary output power corresponding to the carrier power images of the input signal under test, IP3 is the third‐order intercept output power, and IM3 is the ratio of the test signal carrier power to the third‐order intercept power resulting from the nonlinearity. Using the linear approximations to these power transfer functions, the third‐order intermodulation power ratio is defined as ratio Po/Po3 and is expressed in decibels as

Graph of input power vs. output power displaying 2 ascending lines intersecting at IP3; 3 horizontal lines (from top to bottom) labeled P1, P0, and P03; and 2 arrows (OBO and IM3) between P1 and P0, and P0 and P03.

FIGURE 15.39 Third‐order intermodulation characterization for soft saturation SSPA (dashed curves are linear approximations).

Equation (15.153) is plotted in Figure 15.40 as the dotted theoretical curve for various OBO levels based on the SSPA characteristics described in Figures 15.38 and 15.39. The simulated intermodulation distortion discussed in Sections 15.13 and 15.14 is evaluated using a slightly different SSPA model and the theoretical and simulated intermodulation distortion are also shown in Figure 15.40 as the solid curve and the circled data points, respectively. Therefore, Figure 15.39 and Equation (15.153) serve as a general description for the evaluation of intermodulation distortion.

Graph of output backoff vs. intermodulation levels displaying an ascending line labeled IM3, which is parallel to a dashed line with circles, and a dashed curve with circles labeled Imcc5.

FIGURE 15.40 Intermodulation characteristics using soft saturation SSPA response (circled data is for SRRC‐QPSK with α = 0.4).

15.12.2 TWTA with Gain Compensated Response

The TWTA often use gain and phases compensation circuits to extend the linear range of the device. The resulting gain response rises to a slight peak before decreasing monotonically. The goal is to extend the 1 dB gain compression point and thereby extend the linear range. The gain compensated curve‐fit response for a TWTA is expressed as

(15.154) images

and the phase transfer function is approximated as

(15.155) images

It is left as an exercise (see Problem 13) to plot these responses and determine the peak input and output voltages that correspond to the 1 dB gain compression point.

15.13 ESTABLISHING SIGNAL LEVELS FOR SIMULATION MODELING

Consider the soft limiter modeled by curve fitting a fifth‐order polynomial to the measured AM‐AM gain characteristic of a nonlinear device. For this example, the polynomial coefficients are as follows: Cm = (0.0, 0.687834, 6.070865, −190.532776, 2729.8845, −8057.7788) and the peak output voltage (Vpo) is expressed in terms of the peak input voltage (Vpi) as

This voltage gain characteristic is shown in Figure 15.41. The magnitude of the voltage levels and the limiter gain are not important for the purpose simulation. However, it is important to map the voltage level of the composite input signal to the specified operating range of the nonlinear model and, in turn, to map the limiter output level back to the levels used in the simulation program. The importance of the input mapping is to ensure that the signal is operating at the input level corresponding to a specified IBO. Similarly the importance of the output mapping is to ensure that the limiter output is mapped back to the expected simulation level. This procedure allows for the addition of postlimiter noise based on the known signal power level. Throughout this section the focus is on the modeling of a HPA for use in a performance simulation program with an emphasis placed on the HPA nonlinearity. In this regard, the HPA gain is only used to establish the received signal‐to‐noise ratio specified in the simulation program.

Graph of Vpi vs. Vpi displaying 1 dashed and 2 horizontal arrow labeled Vpo(max), V'po, and Vpi(no); 1 dashed and 2 vertical arrows labeled Vpi(max), V'pi, and Vpi(no); and an arrow (OBO) between Vpo(max) and V'po.

FIGURE 15.41 Soft limiter voltage gain characteristic.

Before establishing the mapping, the operating point through the limiter must be identified. Referring to Figure 15.41, the maximum output is denoted Vpo(max) = 0.03055 mV and the corresponding maximum input is Vpi(max) = 0.08 mV. The output operating point (images) is defined in terms of the output power backoff relative to Vpo(max) and the corresponding input operating point is images. The OBO in decibels is: OBO(dB) = 20log(Vpo(max)/images) and images is determined as the inverse of the polynomial expression in (15.156). The voltages images and images apply to the single‐carrier operation discussed in the following section and the voltages images(no) and images(no) apply to the multiple‐carrier operation discussed in Section 15.13.2.

15.13.1 Single‐Carrier Simulations

With a single carrier per channel (SCPC) simulation the limiter may occur in the transmitter HPA of a single‐channel‐modulated waveform or with multiple data channels that are time division multiplexed (TDM) onto a single carrier. The SCPC waveform also applies to a satellite HPA; however, the satellite often combines many carriers to form a FDM downlink.

Consider the SCPC‐modulated waveform is expressed as

(15.157) images

The HPA or limiter operating points are characterized by (images, images) and the mapping from (and to) the simulation signal levels is determined as shown in Figure 15.42. The limiter operating points are determined using the maximum peak output voltage of the limiter Vpo(max) and a specified OBO level such that

(15.158) images
Schematic depicting flow of mapping of simulation levels to limiter operating point from either Vsi or Psi, to either Vso or Pso.

FIGURE 15.42 Mapping of simulation levels to limiter operating points.

The corresponding input images is determined using the inverse of the polynomial expression of the limiter transfer function as indicated in Figure 15.41.

With the modulation function m(t) normalized to result in a unit power level, the power of the simulated waveform is simply the power in the carrier signal with peak voltage Vsi, that is,

(15.159) images

The mapping to the limiter input is determined by the voltage gain, expressed as,

(15.160) images

Similarly, the output of the limiter is mapped back to the simulation signal level using the voltage gain

(15.161) images

It is convenient in the simulation to let the input and output signal levels be equal so that Pso = Psi. That is, when the limiter is not used, the HPA appears as an ideal unit gain amplifier. Neglecting the various propagation and channel distortion losses, the received signal‐to‐noise ratio (γb) is established at the limiter output and the received noise power is computed as

15.13.2 Multiple Carrier Simulations

Multiple carrier simulations will occur in the transmitter HPA when several modulated carriers are combined at the IF to form a FDM waveform; this is also typical of multichannel satellite repeater downlinks. In these cases, the simulation is generally focusing on the performance of a particular carrier or signal and the limiter operating point must be established for the signal of interest.

Consider the composite signal described as

(15.163) images

where

(15.164) images

Considering s(t;no) to be the desired signal with the modulation term m(t;no) normalized to result in unit power, the desired signal power is then the carrier power

(15.165) images

The limiter operating point images for the composite signal is established as described earlier with the input power given by

(15.166) images

However, the operating point for the desired signal is images so the input signal mapping to the limiter is determined by the voltage gain

(15.167) images

where the voltage gain is

Equation (15.168) maps the limiter composite input signal to the desired signal as indicated in Figure 15.41. The limiter output is mapped back to the simulation signal levels using the voltage gain

(15.169) images

where the desired output signal voltage images(no) is determined directly from the limiter transfer function using the input voltage images(no). In this case, the signal‐to‐noise ratio (γb) for the desired signal is established at the limiter output using the noise power

where

(15.171) images

The previous steps, required to setup and initialize the limiter processing, are summarized as follows:

  • Given Vpo(max) determine the composite signal operating point images at the limiter output using the specified OBO
  • Determine the output voltage gain Gvo that maps images to the simulation voltage Vso
  • For single‐carrier operation, compute the noise power images for the specified signal‐to‐noise ratio γb
  • Determine the corresponding composite signal operating point images at the limiter input
  • Determine the input voltage gain Gvi that maps the simulation voltage Vsi to the limiter operating point images

For multiple carriers, the following additional steps are required:

  • Determine the voltage gain Gvn between the signal of interest Vsi(no) and the composite signal Vsi
  • Using Gvn and images determine the limiter input operating point images for the desired signal
  • Determine the limiter output voltage images corresponding to the input images

The final step involves computing the noise power using either (15.162) or (15.170) for a desired signal‐to‐noise ratio γb.

15.14 CASE STUDY: PERFORMANCE SIMULATION OF SRRC‐QPSK WITH SSPA NONLINEARITY

In this case study, three major impacts of the HPA on the system performance are examined and related to the OBO of an SSPA. The three system performance measures are intermodulation (IM) noise, spectrum degradation, and ACI. The two‐tone third and fifth‐order intermodulation distortion terms are examined for various OBO conditions using computer simulation of the SSPA discussed in Section 15.12. The goal is to achieve the theoretical intermodulation IM performance shown as the solid curve in Figure 15.40. The circled data points in Figure 15.40 are based on the simulated IM performance and are used in this section. Following the description of the IM noise impact on the system performance, the spectral degradation of the spectral root‐raised‐cosine offset quadrature phase shift keying (SRRC‐QPSK) modulated waveform is examined for various amounts of OBO. In the concluding section of this case study, the bit‐error performance is examined for the output power backoff with and without adjacent channels. The results suggest that, to avoid significantly degrading the performance due to the ACI with Nyquist channel spacing, the OBO should be greater than 6 dB.

The mapping of the simulated signal levels into, and out of, the SSPA device is based on the descriptions given in Section 15.13. The waveforms used in examining the IM performance are based on SRRC‐QPSK modulation with an excess bandwidth of (1 + α) where α is the SRRC excess bandwidth parameter; α = 0.4 is used in these evaluations. The simulation program generates the desired modulated waveform at baseband with SRRC‐QPSK modulated upper and lower adjacent channels. The data rates, amplitude levels, carrier frequencies, and excess bandwidth for each of the channels are specified with the sampling frequency computed to satisfy the Nyquist criterion for the composite FDMA waveform. The spectrum plots in Figure 15.43 are examples of the three channels operating with symmetrical and asymmetrical adjacent channels. The asymmetrical adjacent channel spectrums, shown in Figure 15.43b, operate with symbol rates images four times the desired channel symbol rate of Rs. In this case, the adjacent channel levels are decreased by images = 6 dB; although the level can be arbitrarily specified in the simulation.

2 Graphs of normalized spectrum over normalized frequency for spectrum of SRRC-OQPSK waveforms, featuring 4.8 ksps channels (left) and 4.8 ksps with 19.2 ksps adjacent channels (right).

FIGURE 15.43 Spectrum of SRRC‐QPSK waveforms (linear PA).

The fundamentals of the signal processing are as follows: The desired (or center) channel waveform is centered fc = 0 Hz and is sampled using 128 samples per symbol, so the Nyquist sampling frequency is fs = 128Rs. The adjacent channel carrier frequencies are located at

and the maximum occupied bandwidth is

where images is the maximum of the upper and lower adjacent channel symbol rates and Δfmax is used, if necessary, to provide additional frequency separation to improve the system performance. Based on these conditions the Nyquist sampling criterion requires that images. The SRRC waveform is generated using a uniformly weighed span of 13 symbols with 6 leading and trailing symbols. On the other hand, the receiver matched filter is based on a total span of seven symbols; this results in a theoretical detection loss of less than 0.01 dB.

15.14.1 Simulation of Third‐ and Fifth‐Order Intermodulation Distortions Terms

The two‐tone intermodulation noise terms, resulting from the SSPA nonlinearity as characterized by (15.151), are evaluated as a function of OBO using the SSPA computer modeling described in Section 15.13. The third‐ and fifth‐order intermodulation noise terms are evaluated by applying two CW signals to the SSPA and examining the spectrum of the SSPA output. The simulation uses a received rms carrier signal level of 1 V that is mapped through the nonlinearity and back to the simulation level of 1 V rms.

The simulation uses a rectangular windowed SRRC frequency response with the QPSK‐modulated waveform operating at a baseband data rate of 4.8 K symbols per second with the adjacent channel at 19.2 Ksps. The adjacent channel is located at a positive carrier frequency in accordance with (15.172) using α = 0.4. The received CW tones for the intermodulation evaluation are obtained by using mark‐hold data on the I/Q rails. The intermodulation tones are clearly discernable in the inherently harmonic‐filled spectrum.

The spectrums are generated using an equivalent of Ns = 128 samples per symbol and averaging 20 spectrums with a cosine windowed fast Fourier transform (FFT) size of Nfft = 16,384 samples; the corresponding frequency resolution is 0.0375 Hz. The results depicted in Figure 15.44 are obtained using a linear PA and show the spectral content using the mark‐hold SRRC‐QPSK waveform to generate the two tones. Figure 15.45 shows the resulting two‐tone intermodulation performance for various OBO conditions; these results are plotted as the circled data points in Figure 15.40.

2 Spectrum plots of normalized frequency vs. normalized spectrum depicting performance with ideal linear PA (no SSPA) with carrier frequency tone (a) and two tones fc and fc+Δf (b).

FIGURE 15.44 Performance with ideal linear PA (no SSPA).

4 Spectrum plots of normalized frequency vs. normalized spectrum depicting SSPA IM performance with OBO level of 0 (a), 3 (b), 6 (c), and 9 (d) dB.

FIGURE 15.45 SSPA IM performance with various OBO levels.

15.14.2 Spectrum Degradation with SSPA OBO

The spectrum of the SRRC‐QPSK‐modulated waveform is shown in Figure 15.46 for various OBO levels; the linear channel spectrum corresponds to an essentially infinite OBO condition. The spectrum results apply to an arbitrary symbol rate as suggested by the normalized abscissa. The OBO is defined relative to the 1 dB gain compression point of the SSPA and the 0 dB backoff case results in the peak carrier voltage being compressed by 1 dB with the average carrier voltage being 3 dB below the peak level. Therefore, a −3 dB OBO results in the average carrier voltage being compressed by 1 dB; this causes considerable compressing or clipping of the peak voltage. The increase in the modulation one‐sided spectral bandwidth beyond the specified limit of 1.4Rs/2 (α = 0.4) is influenced by the SRRC modulator weighted window duration that spans 12 symbols as discussed in Section 4.4.4.1. The impact of the OBO level on the performance with the adjacent channels is examined in the following section; however, it appears as though the OBO should be greater than 6 dB to preserve the excellent Nyquist bandwidth properties of the SRRC‐modulated waveform.

Graph of normalized frequency vs. normalized spectrum depicting SRRC-OQPSK spectrum with SSPA OBO, displaying 6 different curves at -3, -1.5, 0, 6, 12 dB, and linear channel.

FIGURE 15.46 SRRC‐QPSK spectrum with SSPA OBO.

15.14.3 Bit‐Error Performance with OBO and Adjacent Channels

In the final section of this case study, the bit‐error performance of a SRRC‐BPSK‐modulated waveform is examined with and without adjacent channels. In this case the desired channel is centered at baseband with a symbol rate of Rs = 32 ksps and the adjacent channels are centered above and below the desired channel in accordance with (15.172) using α = 0.4. Two adjacent channel symbol rate conditions are examined, one with symmetrical adjacent channels with images and the other with asymmetrical adjacent channels with images. The Monte Carlo simulations are based on using 100K symbols for each signal‐to‐noise ratio ≤6 dB, otherwise, 1 M symbols for each signal‐to‐noise ratio is used. On occasions, when the Monte Carlo results appear inconsistent with the expected performance curve, 10 M symbols are used to establish the data point.

The performance in Figure 15.47 corresponds to the desired channel bit‐error performance without the adjacent channels. The dotted performance curve represents ideal antipodal signaling and is shown as a reference point when using a linear PA; the performance with 12 dB OBO is nearly the same as the ideal curve. The performance with OBO = 0 dB is only degraded by 0.2 dB at Pbe = 10−5. The severe limiting with OBO = −3 dB is very much like that of an ideal hard limiter, in that, the SRRC AM is essentially removed, resulting in a performance degradation of about 0.9 dB.

Graph of signal-to-noise ratio vs. bit-error probability of SRRC-BPSK single-channel performance with SSPA OBO, with markers for -3 (inverted triangle), 0 (triangle), and 12 (circle).

FIGURE 15.47 SRRC‐BPSK single‐channel performance with SSPA OBO.

In the performance evaluation with adjacent channels, each channel is generated using a separate transmitter with identical SSPAs and the channels are combined to form a FDMA channel; the SSPAs operate with identical OBO conditions. With this understanding, Figure 15.48 shows the performance with symmetrical adjacent channels with the 12 dB OBO condition having a negligible effect on the desired channel performance. The degradations with 0 and −3 dB OBO are about 0.5 and 1.8 dB, respectively.

Graph of signal to noise ratio vs. bit-error probability displaying 3 descending curves representing -3 (inverted triangle), 0 (triangle), and 12 (circle).

FIGURE 15.48 SRRC‐BPSK symmetrical adjacent channel performance with SSPA OBO.

The performance with asymmetrical adjacent channels with images is evaluated under the conditions of constant Eb/No and C/No. If the higher data rate channels are modulated on a carrier with the same power as the desired channel, then the channel spectrum will be as shown in Figure 15.43 with the adjacent channel spectral levels images lower than the desired channel spectral level. This corresponds to the constant C/No case that results in 6 dB less Eb/No than is available for the desired channel. This results in degradation of the bit‐error performance with low levels of OBO, although, the impact on the desired channel performance is less severe. On the other hand, if the adjacent channel carriers are increased by images, then the spectrums in Figure 15.43 all have the same level and correspond to the constant Eb/No case. The increase in the carrier power results in a greater loss in the desired channel performance with low levels of OBO. The performance results for these two cases are shown in Figure 15.49 as the dashed and solid curves corresponding to the constant Eb/No and C/No cases, respectively. The performance difference in the degradations is about 2 and 1 dB for OBO = −3 and 0 dB, respectively, and is negligible for an OBO of 12 dB. Furthermore, the absolute performance loss decreases with increasing OBO and is negligible for OBO = 12 dB. The best performance for all channels is obtained when operating in the linear range of the SSPA under the constant Eb/No condition. The performance degradation can also be reduced, for a given OBO, by increasing the channel separation by Δfmax as indicated in (15.173). Therefore, there is a trade‐off between power and bandwidth.

Graph of signal to noise ratio vs. bit-error probability displaying 5 descending curves representing -3 (inverted triangle), 0 (triangle), and 12 (circle).

FIGURE 15.49 SRRC‐BPSK asymmetrical adjacent channel performance with SSPA OBO.

15.15 LINK BUDGET ANALYSIS

In this concluding section of Chapter 15 the previous results are brought into focus by completing a link budget to determine if the modem and transceiver designs are capable of establishing or closing the link in view of the system parameters and the communication channel characteristics. The examples in this section involve the link budget for a communication link between a ground terminal and a satellite and Table 15.8 lists the ground station’s location, the satellite’s orbit, and location in the orbit at an arbitrary initial time. The ground station and the satellite geometry change throughout the satellite orbit period in consideration of the rotating Earth and orbit dynamics as discussed in Chapter 16. The examples are specialized for an uplink signal that is received by the satellite.

TABLE 15.8 Example Ground Terminal and Satellite Parameters

Parameter Value Units Comments
Satellite
Longitude ascending node 180 Degrees Input (east is positive)
Argument of perigee 0 Degrees Input (north is positive)
Satellite orbit angle Degrees Computed
Inclination 7 Degrees Input
Eccentricity 0.003 Input
Ground Terminal Location
Latitude 33 Degrees Input (north is positive)
Longitude −120 Degrees Input (east is positive)
Elevation 0 Meters Input

The various parameters that makeup the link budget focus on the communication range equation as discussed in the introduction to this chapter. Since an ideal demodulator is used Table 15.9 focuses on the transceiver and channel characteristics; however, to satisfy the requirements of the entire communication system, the modem characteristics shown in Tables 15.10 and 15.11 must be included as part of the link budget. Identifying the modem characteristics separately from those of the transceiver and channel is convenient because the interface between the receiver and demodulator is often at an IF, as discussed in Chapter 2, and these subsystems are generally designed, built, and tested independently before the final system integration. In this regard, Tables 15.10 and 15.11 form the basis of an independent modem specification. The principal modem parameters needed to establish the link budget are the required signal‐to‐noise ratio (Eb/No) and the user data rate (Rb). Eb/No is the required signal‐to‐noise ratio, measured in a bandwidth equal to Rb and is selected to achieve a specified bit‐error probability based on the waveform modulation. The required or specified carrier‐to‐noise density ratio is computed as

(15.174) images

Where LM ≥ 1 is the selected link margin and is a significant specification. The link is established, or closed, when the received C/No satisfies the condition

(15.175) images

TABLE 15.9 Satellite System Link Budget (Ideal Demodulator)

Parameter Value Units Comments
Data rate Rb 19.2 kbps Input
Uplink frequency 2.0 GHz Input
Link margin 2.0 dB Input
HPA power 4.6 dBW Input
Antenna feed loss 0.0 dB Input
Antenna gain 10.0 dB Input
   Reflector loss 0.0 dB Input
   Radome loss 0.0 dB Input
   Wet antenna loss 0.0 dB Input
Polarization Input (LVP:AR = ∞)
EIRP 14.6 dBW Computed
Channel
Path loss 0.214 dB Computed (less Fs loss)
   Atmosphere 0.014 dB Input
   Rain 0.2 dB Input
   Antenna tracking 0.0 dB Input
   Multipath 0.0 dB Inputa
   Scintillation 0.0 dB Inputa
Polarization rotation 0.0 Degrees Input
Sky temperature 20.0 °K Inputb
Receiver
Antenna gain 30.0 dBi Input
   Radome loss 0.0 dB Input
      Temperature 290 °K Inputb
   Reflector loss 0.2 dB Input
      Temperature 290 °K Inputb
Polarization RHC Input
   Axial ratio 0.0 dB Input
Polarization loss 3.0 dB Input RHC: AR = 0 dB
Antenna system temperature 506.8 °K Computed
Receiver G/T 3.75 dB/°K Input
Feed loss 0.0 dB Input
   Temperature 290 °K Inputb
Rx noise figure dB Inputc
Receiver system temperature 506.8 °K Computed
Receiver C/No 50.8 dB‐Hz Computed
   Rx phase noise loss 0.2 dB Input
   Rx ISI/IF filter loss 0.1 dB Input
Channel bandwidth W 4 MHz Input
Receiver/user bandwidth 30.0 kHz Input
Received Eb/No (req’d) 8.0 dB Computed
Specified Eb/No 6.0 dB Input
Margin 2.0 dB Computed

aBased on waveform and demodulator algorithm analysis and simulation.

bThese parameters are used to compute the system noise temperatures when antenna gain and G/T is not provided as input.

cThe receiver noise figure is not needed when antenna gain and G/T is provided, otherwise, it is used as the starting point for computing the system noise temperatures.

TABLE 15.10 Modem Detection Budget

Parameter Value Units Comments
Modulation 1 Bits/symbol BPSK
Data rate Rb 19.2 kbps Specified
Theoretical Eb/Noa,b 2 dB Input
Bandwidth Bcbb 57.6 kHz Baseband
Link margin 1 dB Input
Implementation loss 1.4 dB Computed
   Demod phase noise 0.2 dB Input
   PLL tracking 0.4 dB Input
   Symbol tracking 0.2 dB Input
   ACI/CCI 0.2 dB Input
   MF detection 0.1 dB Input
   Quantization 0.1 dB Input
   Signal processing 0.2 dB Input
Modem spec. Eb/No 4.4 dB Computed
Modem spec. C/No|reqd 47.2 dB‐Hz Computed

aAt Pbe = 10−5.

bRate 1/3 turbo code.

TABLE 15.11 Modem Acquisition Budget

Parameter Value Units Comments
Data rate Rb 19.2 kbps Specified
Theoretical Ecb/Noa −2.8 dB Input
Bandwidth Bcbb 57.6 kHz Baseband
Link margin 0 dB Input
Implementation lossc 2.6 dB Computed
   Channel fading 0.4 dB Input
   ACI/CCI 0.2 dB Input
   Interference 0.2 dB Input
   Antenna scalloping
   Frequency scalloping 0.25 dB Input
   Time scalloping 0.25 dB Input
   NonCoh integration 1.0 dB Input
   Quantization 0.1 dB Input
   Signal processing 0.2 dB Input
Acquisition spec. Pacq 0.95 Probability
Modem spec. Ecb/No −0.2 dB Computed
Modem spec. C/No 47.4 dB‐Hz Computed

aThis is 0.8 dB lower than detection budget.

bRate 1/3 turbo code.

cNoncoherent integration losses are overcome with increased acquisition time.

The channel bandwidth W must support the transmitted symbol rate which may be significantly wider than the data rate with FEC coding and spread‐spectrum processing. During acquisition the channel bandwidth may be selected to support the frequency uncertainty range of the received signal.

The modem link budget typically focuses on the message detection requirements as shown in Table 15.10. However, an equally important consideration is the modem acquisition link budget shown in Table 15.11. In both of these modem link budgets, the selection of the link margins is established; however, the theoretical Eb/No specification for acquisition may be selected one or two decibels lower than for data detection; an alternate method is to choose the acquisition link margin somewhat less than that selected for the modem detection budget.* The acquisition processing must detect the presence of the received signal and estimate the necessary parameters with sufficient accuracy to provide for the message detection. The link acquisition budget requirements include scallop losses associated with antenna scanning, symbol time and frequency uncertainties, and other modem‐related losses as shown in the acquisition loss budget. During the initial search for signal detection, the carrier frequency and phase are unknown and noncoherent integration may be required to establish an adequate signal‐to‐noise ratio to meet the specified correct acquisition probability.

The channel scintillation and fading losses are influenced by the waveform design and the demodulator algorithms used to mitigate these losses. Therefore, although these losses are included as channel loss, they are established by analysis and simulation of the degradation in the bit‐error performance for the selected waveform. For example, the channel may result in a scintillation or multipath fading loss of 10–20 dB; however, properly designed and analyzed waveforms, using various combinations of FEC coding, data interleaving, and frequency and spatial diversity, will very likely result in a loss of 1–2 dB. Similarly, the receiver intersymbol interference (ISI) loss is evaluated based on the waveform modulation and is typically determined by evaluation of the sensitivity of the modulated waveform to the filter bandwidth, amplitude ripple, and phase linearity. The performance loss of ACI and co‐channel interference (CCI) is usually determined by analysis and verified by simulation of the improvements attainable through the use of adaptive canceling algorithms.

ACRONYMS

ACI
Adjacent channel interference
ADC
Analog‐to‐digital converter
AM
Amplitude modulation
AM‐AM
Amplitude modulation to amplitude modulation (conversion)
AM‐PM
Amplitude modulation to phase modulation (conversion)
AR
Axial ratio
BPSK
Binary phase shift keying
C/I
Carrier‐to‐intermodulation noise (ratio)
C/Io
Carrier‐to‐intermodulation noise density (ratio)
CCI
Co‐channel interference
CCIR
Consultative Committee on International Radio
CP
Circular polarized (polarization)
CW
Continuous wave
EIRP
Effective isotropic radiated power
EP
Elliptical polarization
ERP
Effective radiated power
FDM
Frequency division multiplex
FDMA
Frequency division multiple access
FEC
Forward error correction
FFT
Fast Fourier transform
G/T
Gain‐temperature ratio (receive antenna)
HP
Horizontal polarized (polarization)
HPA
High‐power amplifier
I&D
Integrate‐and‐dump
I/Q
In‐phase and quadrature (rails)
IBO
Input backoff
IF
Intermediate frequency
IM
Intermodulation
IMP
Intermodulation product
ISI
Intersymbol interference
LEO
Low Earth orbit (satellite)
LHC
Left‐hand circular
LHCP
Left‐hand circular polarized (or polarization)
LNA
Low‐noise amplifier
LOS
Line of sight
LP
Linear polarization
MPSK
Multiphase shift keying
OBO
Output backoff
OFDM
Orthogonal frequency division multiplex
OQPSK
Offset quadrature phase shift keying
PA
Power amplifier
PCT
Percent‐of‐time
PLL
Phaselock loop
PM
Phase modulation
PSD
Power spectral density
QAM
Quadrature amplitude modulation
QPSK
Quadrature phase shift keying
RF
Radio frequency
RHC
Right‐hand circular
RHCP
Right‐hand circular polarized (polarization)
SAM
Simple attenuation model
SCPC
Single carrier per channel
SRRC
Spectral root‐raised‐cosine
SSPA
Solid‐state power amplifier
SWR
Standing wave ratio
TDM
Time division multiplex
TWT
Traveling wave tube
TWTA
Traveling wave tube amplifier
VP
Vertical polarized
VSWR
Voltage standing wave ratio

PROBLEMS

  1. Given a receiver antenna with a physical temperature of Ta = 15°C, Tain = 0, and a gain of G = 50 dB. The antenna is connected to a receiver (with a noise figure Fn = 6 dB) through a lossy 1 dB cable at a physical temperature of 15°C. Determine the antenna G/T ratio in dB/°K. Repeat this problem using Tain = 1000°K and, using (15.13), determine the increase in the received signal‐to‐noise ratio required to maintain system performance.
  2. Derive the expression of the output temperature Tout for a cascade of two noisy filters shown in the following figure. Use the indicated losses and physical temperatures of the noisy devices; all of the temperatures are in °K. With Tin = 0, the effective input temperature of the two cascaded devices is defined as images. Evaluate the effective input temperature of the cascaded devices and express the result in terms of the effective temperatures images and images of the two equivalent noiseless devices.
    Flow diagram of cascade of two noisy filters starting from Tin to Filter 1 to Filter 2 to Tout.
  3. Write a general simulation program for the example in Section 15.2.3.1 to evaluate the antenna G/T as expressed in (15.56). The program is to provide for inputting all of the parameter values and outputting data files of Ta, Trs, Fns, and G/T(dB) for 0 ≤ TSun ≤ 100,000°K in steps of 500°K; create the plots parametrically with Ga and θB. The decreasing Sun temperatures reflect the percentage of the Sun’s area in the antenna beamwidth for a given ratio ΩSuna.
  4. Referring to the antenna gain (G) as expressed in Equation 15.3, derive expressions for: (A) the diameter D of a circular (dish) antenna as a function of the gain G; (B) the 3‐dB beamwidth θB for a uniformly weighted circular aperture as a function of G; (C) plot D in meters and θB in degrees as a function of the gain G in dB for carrier frequencies of 100 and 500 MHz.
  5. Using the general expression for the elliptically polarized wave given by (15.89), derive the expression for Ex as a function of Ey and the parameters Mx and My for the LP, that is, for δ = 0 and π.
  6. For ideally polarized received waves and antennas with the indicated polarizations, determine the polarization mismatch factor and the polarization loss, expressed in decibels, for the following cases:
    1. LHCP received wave and RHCP receiver antenna.
    2. RHCP received wave and RHCP receiver antenna.
    3. LHCP received wave and LP receiver antenna.
    4. LP received wave and LP receiver antenna.
    5. Elliptically polarized received wave and LP receiver antenna.
  7. Develop a computer program and evaluate the cross‐polarization discrimination, xpd, and the isolation Iisol for a nonideal VP antenna using a linearly polarized received wave with a VP copolarization state. This involves evaluating (15.100) using (15.101) and (15.95). Start by generating the ideal xpd vs. images curve using ARs of r = 104 for the received wave and ideal VP antenna. Then evaluate the nonideal antenna isolation under the four conditions: rax = rac = 300, 100, 30, and 10 for an ideal received wave using images; these ARs are linear values, not in dB. Plot the Iisol vs. xpd in decibels with each axis ranging from 0 to 50 dB. Hints: For VP polarization the antenna and the VP copolarized wave the tilt angles are τw = τac = 90° so τwo = τax = 0°. Also, in generating the ideal xpd curve use tilt angles images = 0(0.125)180°, that is, 0° to 180° in increments of 0.125°; this will result in a good quality plot.
  8. Develop a computer program and evaluate the cross‐polarization discrimination, xpd, for an ideal VP antenna using an ideal linearly polarized received wave with a VP copolarization state under the conditions described in Problem 7. (This part is complete if Problem 7 was examined.) Now develop the expression for the theoretical xpd for the same VP wave and antenna and compare with the earlier computer‐generated xpd by plotting both curves on the same graph.
  9. Derive the expression for the noise figure of a cascade of N linear amplifiers with individual noise figures Fni, gains Gi, and effective temperatures images, i = 1, …, N.
  10. Derive the expression for the effective noise temperature of a cascade of N lossless attenuators each having physical temperature Tpi and loss Li, i = 1, …, N.
  11. An antenna with effective temperature images is covered by a radome with a loss Lrad and effective temperature images. The antenna output is connected to an antenna feed with loss Lrfd and effective temperature images. Express the effective temperature of a single equivalent device, with loss L = Lrad Lrfd, in terms of the temperatures: images, images, and images. Hint: Compute images in the following figure.
    Two flow diagrams illustrating antenna configuration (left) and equivalent configuration (right).
  12. Consider that the antenna feed, described in Problem 11, is connected to the input of an LNA with a receiver system noise temperature Trs. If the antenna gain is 60 dB derive the expression for the antenna system temperature and the antenna G/T ratio and indicate the units for each result.
  13. For the gain compensated TWTA described in Section 15.12.2, plot the principal signal linear response and the third‐ and fifth‐order intermodulation responses. Then identify the intercept points IP3 and IP5 and determine the intermodulation distortion terms IM3 and IM5 for two values of the OBO. Express all results in decibels.

REFERENCES

  1. 1. R.S. Berkowitz, Editor, Modern Radar: Analysis, Evaluation, and System Design, John Wiley & Sons, Inc., New York, 1965.
  2. 2. R.S. Berkowitz, Editor, Modern Radar: Analysis, Evaluation, and System Design, PART V, Radio Frequency Considerations, Chapter 3, S. Charton, “Radar System Sensitivity,” pp. 415–430, John Wiley & Sons, Inc., New York, 1965.
  3. 3. M.I. Skolnik, Introduction to Radar Systems, Chapter 7, “Antennas,” McGraw‐Hill, New York, 1962.
  4. 4. R.S. Berkowitz, Editor, Modern Radar: Analysis, Evaluation, and System Design, PART V, Radio Frequency Considerations, Chapter 3, S. Charton, “Radar System Sensitivity,” pp. 415–420, John Wiley & Sons, Inc., New York, 1965.
  5. 5. L.V. Blake, Radar Range‐Performance Analysis, Munro Publishing Co., Silver Spring, MD, 1991.
  6. 6. G.P. Kuiper, Editor, The Solar System, Vol. 1, Chapter 7, “The Sun,” p. 485, University of Chicago Press, Chicago, 1953.
  7. 7. J.D. Kraus, H.C. Ko, “Celestial Radio Radiation,” Ohio State University Report 1 of Radio Frequency (RF) Project 673, U.S. Air Force Contract 19(604)‐1591, Columbus, OH, May 1957.
  8. 8. L.V. Blake, Radar Range‐Performance Analysis, pp. 160–162, Munro Publishing Co., Silver Spring, MD, 1991.
  9. 9. R.H. Brown, C. Hazard, “A Model of the Radio Frequency Radiation from the Galaxy,” Philosophical Magazine (Philos. Mag.), Vol. 44, pp. 939–963, September 1953.
  10. 10. D.C. Hogg, W.W. Mumford, “The Effective Noise Temperature of the Sky,” Microwave Journal, Vol. 3, No. 3, pp. 80–84, March 1960.
  11. 11. H.C. Ko, “The Distribution of Cosmic Radio Background Radiation,” Proceeding of the IRE, Vol. 46, pp. 208–215, January 1958.
  12. 12. S.F. Smerd, “Radio‐Frequency Radiation from the Quiet Sun,” Australian Journal of Scientific Research, Vol. A3, p. 34, 1950.
  13. 13. P.D. Strum, “Considerations in High‐Sensitivity Microwave Radiometry,” Proceedings of the IRE, Vol. 46, Issue 1, pp. 43–53, January 1958.
  14. 14. R.S. Berkowitz, Editor, Modern Radar: Analysis, Evaluation, and System Design, Part V, Chapter 3, S. Charton, “Radar System Sensitivity,” pp. 412–414, John Wiley & Sons, Inc., New York, 1965.
  15. 15. R. Gardner, “Final Engineering Report on Antenna Noise Temperature Study,” AIL Report 3304‐11, Contract DA‐49‐170‐SC‐1547, Airborne Instruments Laboratory, Mineola, NY, November 1957.
  16. 16. J.A. Stratton, Electromagnetic Theory, McGraw‐Hill, New York, 1941.
  17. 17. S. Silver, Editor, Microwave Antenna Theory and Design, MIT Radiation Laboratory Series, Vol. 12, McGraw‐Hill, New York, 1949.
  18. 18. J.D. Kraus, Antennas, McGraw‐Hill, New York, 1949.
  19. 19. L.J. Ippolito, Radiowave Propagation in Satellite Communications, Van Nostrand Reinhold Co., Inc., New York, 1986.
  20. 20. L.J. Ippolito, “Radio‐Wave Propagation for Space Communications Systems,” NASA Technical Paper 1770, p. 29, Figure 12 and Table 9, NASA Office of Space and Terrestrial Applications, Scientific and Technical Information Branch, NASA National Aeronautics and Space Administration, Washington, DC, 1981.
  21. 21. R.K. Crane, “Prediction of Attenuation by Rain,” IEEE Transactions on Communications, Vol. COM‐28, No. 9, pp. 1717–1733, September 1980.
  22. 22. G.A. Arnold, T.W. Kao, “Models Describe Rain Attenuation in EHF Systems,” Microwave and RF, Part 1, pp. 74–83, April 1996.
  23. 23. International Telecommunications Union‐Radio (ITU‐R), “Calculation of Long‐Term Attenuation Statistics for Point Rainfall Rate,” Report 564‐4: Recommendations and Reports of the CCIR, Vol. V, Formerly: Consultative Committee on International Radio (CCIR), Geneva, 1986.
  24. 24. A.W. Dissanayake, “Proposed Amendment to Consultative Committee on International Radio (CCIR), Recommendation P618: Propagation Data and Prediction Methods Required for the Design of Earth‐Space Telecommunication Systems, Prediction of Rain Attenuation,” Document USSG5C/1T, International Telecommunications Union, Geneva, Switzerland, 1992.
  25. 25. W.L. Stutzman, K.M. Yon, “A Simple Rain Attenuation Model for Earth‐Space Radio Links Operating at 10–35 GHz,” Radio Science, Vol. 21, No. 1, pp. 65–72, January–February 1986.
  26. 26. R. K. Crane, “A Global Model for Rain Attenuation Prediction,” EASCON’78 Record, IEEE Aerospace and Electronic Systems Society, Arlington, VA, pp. 391–395, September 25–27, 1978.
  27. 27. W.J. James, “The Effect of the Weather in Eastern England on the Performance of X‐Band Ground Radars,” Technical Note No. 655, Royal Radar Establishment, Malvern, England.
  28. 28. W. Holzer, “Atmospheric Attenuation in Satellite Communications,” Microwave Journal, Vol. 8, pp. 119–125, March 1965.
  29. 29. L.J. Ippolito, Propagation Effects Handbook for Satellite System Design, Fourth Edition, NASA Reference Publication 1082(04), Section 3.4.2, Westinghouse Electric Corporation, Baltimore, MD, 1989.
  30. 30. J.D. Kraus, Antennas, Chapter 15, “Antenna Measurements,” pp. 444–492, McGraw‐Hill, New York, 1950.
  31. 31. L.J. Ippolito, Propagation Effects Handbook for Satellite Systems Design, Fourth Edition, NASA Reference Publication 1082(04), Chapter IV, “Polarization on Earth‐Space Path,” pp. 4‐1–4‐54, Westinghouse Electric Corporation, Baltimore, MD, 1989.
  32. 32. D.B. Leeson, “A Simple Model of Feedback Oscillators,” Proceedings of the IEEE, Vol. 54, pp. 329–330, February 1966.
  33. 33. J.A. Barnes, “Models for the Interpretation of Frequency Stability Measurements,” National Bureau of Standards, NBS Technical Note 683, U.S. Department of Commerce, Washington, DC, 1976.
  34. 34. G. Sauvage, “Phase Noise in Oscillators: A Mathematical Analysis of Leeson’s Model,” IEEE Transactions on Instrumentation and Measurement, Vol. IM‐24, pp. 408–410, December 1977.
  35. 35. A. Hajimiri, T.H. Lee, The Design of Low Noise Oscillators, Kluwer Academic, Norwell, MA, 1999.
  36. 36. T.H. Lee, A. Hajimiri, “Oscillator Phase Noise: A Tutorial,” IEEE Journal of Solid‐State Circuits, Vol. 35, pp. 326–336, March 2000.
  37. 37. W.P. Robins, Phase Noise in Signal Sources (Theory and Applications), Peter Peregrinus, Ltd., London, UK, 1996.
  38. 38. F.M. Gardner, Phaselock Techniques, Third Edition, Chapter 7, “Effects of Phase Noise,” pp. 143–182, Chapter 9, “Oscillators,” pp. 209–236, Wiley‐Interscience, John Wiley and Sons, Inc., Hoboken, NJ, 2005.
  39. 39. A.R. Chi, Guest Editor, “Special Issue on Frequency Stability,” Proceeding of the IEEE, Vol. 54, No. 2, pp. 101–338, February 1966.
  40. 40. J. Rutman, “Characterization of Phase and Frequency Instabilities in Precision Frequency Sources: Fifteen Years of Progress,” Proceeding of the IEEE, Vol. 66, No. 9, September 1978.
  41. 41. J. Rutman, G. Sauvage, “Measurement of Frequency Stability in Time and Frequency Domains via Filtering of Phase Noise,” IEEE Transactions on Instrumentation and Measurement, Vol. IM‐23, pp. 515–518, December 1974.
  42. 42. I.S. Gradshteyn, I.M. Ryzhik, Table of Integrals, Series and Products,” p. 292, Sect. 3.241, no. 2, Academic Press, New York, 1980.
  43. 43. U.S.A. Department of Defense (DOD), Interface Standard, MIL‐STD‐188‐164B, Interoperability of SHF Satellite Communications Terminals, p. 19, Figure 1, Terminal Phase Noise, Defense Information Systems Agency (DISA), Fort Meade, MD, March 23, 2012.
  44. 44. F.M. Gardner, Phaselock Techniques, John Wiley & Sons, Inc., New York, 1967.
  45. 45. D. Baker, “The Effects of Phase Noise on High‐Order QAM Systems,” Communication Systems Design, Vol. 5, No. 10, pp. 35–42, October 1999.
  46. 46. H.J. Reich, J.G. Skalnik, P.F. Ordung, H.L. Krauss, Microwave Principles, D. Van Nostrand Company, Inc., Princeton, NJ, 1959.
  47. 47. A.R. Kaye, D.A. George, M.J. Eric, “Analysis and Compensation of Band‐Pass Nonlinearities for Communications,” IEEE Transactions on Communications, Vol. COM‐20, pp. 965–972, October 1972.
  48. 48. A.A.M. Saleh, “Frequency‐Independent and Frequency‐Dependent Nonlinear Models of TWT Amplifiers,” IEEE Transactions on Communications, Vol. COM‐29, Issue 11, pp. 1715–1720, November 1981.
  49. 49. A.A.M. Saleh, “Intermodulation Analysis of FDMA Satellite Systems Employing Compensated and Uncompensated TWT’s,” IEEE Transactions on Communications, Vol. COM‐30, Issue 5, pp. 1233–1242, May 1982.
  50. 50. O. Shimbo, “Effects of Intermodulation, AM‐PM Conversion and Additive Noise in Multi‐Carrier TWT Systems,” Proceeding of the IEEE, Vol. 59, pp. 230–238, February 1971.
  51. 51. J.C. Fuenzalida, O. Shimbo, W.L. Cook, “Time‐Domain Analysis of Intermodulation Effects Caused by Nonlinear Amplifiers,” COMSAT Technical Review, Vol. 3, pp. 89–143, Spring 1973.
  52. 52. E. Imboldi, G.R. Stette, “AM‐to‐PM Conversion and Inter‐modulation in Nonlinear Devices,” Proceedings of the IEEE, Vol. 61, pp. 796–797, June 1973.
  53. 53. F.E. Bond, H.F. Meyer, “Intermodulation Effects in Limiter Amplifier Repeaters,” IEEE Transactions on Communications, COM‐22, pp. 127–135, April 1970.
  54. 54. M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions With Formulas, Graphs, and Mathematical Tables, Second Printing, p. 228, eq. 5.1.1, November 1964, National Bureau of Standards Applied Mathematics Series, 55, U.S. Department of Commerce, Washington, DC, June 1964.

NOTES

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset