Kolias, N.J., Compton, R.C., Fitch, J.P., Pozar, D.M. “Antennas” The Electrical Engineering Handbook Ed. Richard C. Dorf Boca Raton: CRC Press LLC, 2000 38 Antennas 38.1Wire Short Dipole?Directivity?Magnetic Dipole?Input Impedance?Arbitrary Wire Antennas?Resonant Half-Wavelength Antenna?End Loading?Arrays of Wire Antennas?Analysis of General Arrays?Arrays of Identical Elements?Equally Spaced Linear Arrays?Planar (2-D) Arrays?Yagi–Uda Arrays?Log- Periodic Dipole Arrays 38.2Aperture The Oscillator or Discrete Radiator?Synthetic Apertures?Geometric Designs?Continuous Current Distributions (Fourier Transform)?Antenna Parameters 38.3Microstrip Antennas Introduction?Basic Microstrip Antenna Element?Feeding Techniques for Microstrip Antennas?Microstrip Antenna Arrays?Computer-Aided Design for Microstrip Antennas 38.1 Wire N.J. Kolias and R.C. Compton Antennas have been widely used in communication systems since the early 1900s. Over this span of time scientists and engineers have developed a vast number of different antennas. The radiative properties of each of these antennas are described by an antenna pattern. This is a plot, as a function of direction, of the power P r per unit solid angle W radiated by the antenna. The antenna pattern, also called the radiation pattern, is usually plotted in spherical coordinates q and j. Often two orthogonal cross sections are plotted, one where the E-field lies in the plane of the slice (called the E-plane) and one where the H-field lies in the plane of the slice (called the H-plane). Short Dipole Antenna patterns for a short dipole are plotted in Fig. 38.1. In these plots the radial distance from the origin to the curve is proportional to the radiated power. Antenna plots are usually either on linear scales or decibel scales (10 log power). The antenna pattern for a short dipole may be determined by first calculating the vector potential A [Collin, 1985; Balanis, 1982; Harrington, 1961; Lorrain and Corson, 1970]. Using Collin’s notation, the vector potential in spherical coordinates is given by (38.1)Aaa=- - m p qq q0 0 4 Idl e r jkr r ( cos sin) N.J. Kolias Raytheon Company R.C. Compton Cornell University J. Patrick Fitch Lawrence Livermore Laboratory David M. Pozar University of Massachusetts at Amherst ? 2000 by CRC Press LLC where k 0 = 2p/l 0 , and I is the current, assumed uniform, in the short dipole of length dl (dl << l 0 ). Here the assumed time dependence e jw t has not been explicitly shown. The electric and magnetic fields may then be determined using (38.2) The radiated fields are obtained by calculating these fields in the so-called far-field region where r >> l. Doing this for the short dipole yields (38.3) where Z 0 = . The average radiated power per unit solid angle W can then be found to be (38.4) FIGURE 38.1 Radiation pattern for a short dipole of length dl (dl << l 0 ). These are plots of power density on linear scales. (a) E-plane; (b) H-plane; (c) three-dimensional view with cutout. EA A HA=- + ?? × =?′j j w wm e m 00 0 1 Ea Ha = = - - jZ Idlk e r jIdlk e r jk r jk r 00 0 0 0 4 4 sin sin q p q p q j m 0 e 0 ¤ D DW P rIZdlk r r (, ) {}() sinqj q p =?′×= 1 2 32 22 0 2 0 2 2 2 e EH*a ** ? 2000 by CRC Press LLC Directivity The directivity D(q,j) and gain G(q,j) of an antenna are defined as (38.5) Antenna efficiency, h, is given by (38.6) For many antennas h ?1 and so the words gain and directivity can be used interchangeably. For the short dipole (38.7) The maximum directivity of the short dipole is 3/2. This single number is often abbreviated as the antenna directivity. By comparison, for an imaginary isotropic antenna which radiates equally in all directions, D(q,j) = 1. The product of the maximum directivity with the total radiated power is called the effective isotropic radiated power (EIRP). It is the total radiated power that would be required for an isotropic radiator to produce the same signal as the original antenna in the direction of maximum directivity. Magnetic Dipole A small loop of current produces a magnetic dipole. The far fields for the magnetic dipole are dual to those of the electric dipole. They have the same angular dependence as the fields of the electric dipole, but the polar- ization orientations of E and H are interchanged. (38.8) where M = p r 0 2 I for a loop with radius r 0 and uniform current I. Input Impedance At a given frequency the impedance at the feedpoint of an antenna can be represented as Z a = R a + jX a . The real part of Z a (known as the input resistance) corresponds to radiated fields plus losses, while the imaginary part (known as the input reactance) arises from stored evanescent fields. The radiation resistance is obtained from R a = 2P r /|I| 2 where P r is the total radiated power and I is the input current at the antenna terminals. For electrically small electric and magnetic dipoles with uniform currents D P P G P P r r r (,) (,) (,) (,) qj p qj p qj p qj p == Radiated power per solid angle Total radiated power/4 / / Radiated power per solid angle Total input power/4 / / in DDW DDW 4 4 h qj qj o= P P G D r in (,) (,) D(,) sinqj q= 3 2 2 Ha Ea =- = - - Mk e r MZk e r jkr jkr 0 2 00 2 0 0 4 4 sin sin q p q p q j ? 2000 by CRC Press LLC (38.9) The reactive component of Z a can be determined from X a = 4w(W m -W e )/|I| 2 where W m is the average magnetic energy and W e is the average electric energy stored in the near-zone evanescent fields. The reflection coefficient, G, of the antenna is just (38.10) where Z 0 is the characteristic impedance of the system used to measure the reflection coefficient. Arbitrary Wire Antennas An arbitrary wire antenna can be considered as a sum of small current dipole elements. The vector potential for each of these elements can be determined in the same way as for the short dipole. The total vector potential is then the sum over all these infinitesimal contributions and the resulting E in the far field can be found to be (38.11) where the integral is over the contour C of the wire, a is a unit vector tangential to the wire, and r¢ is the radial vector to the infinitesimal current element. Resonant Half-Wavelength Antenna The resonant half-wavelength antenna (commonly called the half-wave dipole) is used widely in antenna systems. Factors contributing to its popularity are its well-understood radiation pattern, its simple construction, its high efficiency, and its capability for easy impedance matching. The electric and magnetic fields for the half-wave dipole can be calculated by substituting its current distribution, I = I 0 cos(k 0 z), into Eq. (38.11) to obtain (38.12) The total radiated power, P r , can be determined from the electric and magnetic fields by integrating the expression 1/2 Re {E ′ H* · a r } over a surface of radius r. Carrying out this integration yields P r = 36.565 |I 0 | 2 . The radiation resistance of the half-wave dipole can then be determined from R dl R r a a = ? è ? ? ? ÷ = ? è ? ? ? ÷ 80 320 2 0 2 6 0 0 4 p l p l electric dipole magnetic dipole G= - + ZZ ZZ a a 0 0 Eaa ar r () [( ) ]()rjkZ e r Il e dl jk r rr c jk =×-¢¢ - × ¢ ò 00 0 0 4p Ea Ha = ? è ? ? ? ÷ = ? è ? ? ? ÷ - - jZ I e r jI e r jk r jk r 00 0 2 2 2 2 0 0 cos cos sin cos cos sin p q qp p q qp q j ? 2000 by CRC Press LLC (38.13) This radiation resistance is considerably higher than the radiation resistance of a short dipole. For example, if we have a dipole of length 0.01l, its radiation resistance will be approximately 0.08 W (from Eq. 38.9). This resistance is probably comparable to the ohmic resistance of the dipole, thereby resulting in a low efficiency. The half-wave dipole, having a much higher radiation resistance, will have much higher efficiency. The higher resistance of the half-wave dipole also makes impedance matching easier. End Loading At many frequencies of interest, for example, the broadcast band, a half-wavelength becomes unreasonably long. Figure 38.2 shows a way of increasing the effective length of the dipole without making it longer. Here, additional wires have been added to the ends of the dipoles. These wires increase the end capacitance of the dipole, thereby increasing the effective electrical length. Arrays of Wire Antennas Often it is advantageous to have several antennas operating together in an array. Arrays of antennas can be made to produce highly directional radiation patterns. Also, small antennas can be used in an array to obtain the level of performance of a large antenna at a fraction of the area. The radiation pattern of an array depends on the number and type of antennas used, the spacing in the array, and the relative phase and magnitude of the excitation currents. The ability to control the phase of the exciting currents in each element of the array allows one to electronically scan the main radiated beam. An array that varies the phases of the exciting currents to scan the radiation pattern through space is called an electronically scanned phased array. Phased arrays are used extensively in radar applications. Analysis of General Arrays To obtain analytical expressions for the radiation fields due to an array one must first look at the fields produced by a single array element. For an isolated radiating element positioned as in Fig. 38.3, the electric field at a far- field point P is given by (38.14) where K i (q,j) is the electric field pattern of the individual element, a i e –jai is the excitation of the individual element, R i is the position vector from the phase reference point to the element, i p is a unit vector pointing toward the far-field point P, and k 0 is the free space wave vector. Now, for an array of N of these arbitrary radiating elements the total E-field at position P is given by the vector sum (38.15) This equation may be used to calculate the total field for an array of antennas where the mutual coupling between the array elements can be neglected. For most practical antennas, however, there is mutual coupling, R P I a r =? 2 73 0 2 ** W FIGURE 38.2Using end loading to increase the effective electrical length of an electric dipole. EK Ri iii jk ae ip i = ×- (,) [( )] qj a 0 EEK Ri tot == ×- = - = - ??ii jk ipi i N i N ae(,) [( )] qj a 0 0 1 0 1 ? 2000 by CRC Press LLC and the individual patterns will change when the element is placed in the array. Thus, Eq. (38.15) should be used with care. Arrays of Identical Elements If all the radiating elements of an array are identical, then K i (q,j) will be the same for each element and Eq. (38.15) can be rewritten as (38.16) This can also be written as (38.17) The function f(q,j) is normally called the array factor or the array polynomial. Thus, one can find E tot by just multiplying the individual element’s electric field pattern, K(q,j), by the array factor, f(q,j). This process is often referred to as pattern multiplication. The average radiated power per unit solid angle is proportional to the square of E tot . Thus, for an array of identical elements (38.18) Equally Spaced Linear Arrays An important special case occurs when the array elements are identical and are arranged on a straight line with equal element spacing, d, as shown in Fig. 38.4. If a linear phase progression, a, is assumed for the excitation currents of the elements, then the total field at position P in Fig. 38.4 will be FIGURE 38.3Diagram for determining the far field due to radiation from a single array element. (Source: Reference Data for Radio Engineers, Indianapolis: Howard W. Sams & Co., 1975, chap. 27–22. With permission.) EK Ri tot = ×- = - ? (,) [( )] qj a ae i jk i N ip i0 0 1 EK Ri tot where == ×- = - ? (,)(,) (,) [( )] qj qj qj a ffae i jk ipi i N 0 0 1 D DW P f r (,) ~(,)(,) qj qj qj** *K 22 ? 2000 by CRC Press LLC (38.19) where y = k 0 d cos q – a. Broadside Arrays Suppose that, in the linear array of Fig. 38.4, all the excitation currents are equal in magnitude and phase (a 0 = a 1 = . . . = a N – 1 and a = 0). The array factor, f(y), then becomes (38.20) This can be simplified to obtain the normalized form (38.21) Note that f'(y) is maximum when y = 0. For our case, with a = 0, we have y = k 0 d cosq. Thus f'(y) will be maximized when q = p/2. This direction is perpendicular to the axis of the array (see Fig. 38.4), and so the resulting array is called a broadside array. FIGURE 38.4 A linear array of equally spaced elements. EK KK tot = == - = - = - ? ? (, ) (, ) (, )( ) ( cos ) qj qj qj y qa y ae ae f n jn k d n N n jn n N 0 0 1 0 1 faea e e jn n N jN j ()y y y y == - - = - ?0 0 1 0 1 1 ¢ ==f f aN N N () () sin sin y y y y 0 2 2 ? 2000 by CRC Press LLC Phased Arrays By adjusting the phase of the elements of the array it is possible to vary the direction of the maximum of the array’s radiation pattern. For arrays where all the excitation currents are equal in magnitude but not necessarily phase, the array factor is a maximum when y = 0. From the definition of y, one can see that at the pattern maximum Thus, the direction of the array factor maximum is given by (38.21b) Note that if one is able to control the phase delay, a, the direction of the maximum can be scanned without physically moving the antenna. Planar (2-D) Arrays Suppose there are M linear arrays, all identical to the one pictured in Fig. 38.4, lying in the yz-plane with element spacing d in both the y and the z direction. Using the origin as the phase reference point, the array factor can be determined to be (38.22) where a y and a z are the phase differences between the adjacent elements in the y and z directions, respectively. The formula can be derived by considering the 2-D array to be a 1-D array of subarrays, where each subarray has an antenna pattern given by Eq. (38.19). If all the elements of the 2-D array have excitation currents equal in magnitude and phase (all the a mn are equal and a z = a y = 0), then the array will be a broadside array and will have a normalized array factor given by (38.23) Yagi–Uda Arrays The Yagi–Uda array can be found on rooftops all over the world—the standard TV antenna is a Yagi–Uda array. The Yagi–Uda array avoids the problem of needing to control the feeding currents to all of the array elements by driving only one element. The other elements in the Yagi–Uda array are excited by near-field coupling from the driven element. The basic three-element Yagi–Uda array is shown in Fig. 38.5. The array consists of a driven antenna of length l 1 , a reflector element of length l 2 , and a director element of length l 3 . Typically, the director element is shorter than the driven element by 5% or more, while the reflector element is longer than the driven element by 5% or more [Stutzman and Thiele, 1981]. The radiation pattern for the array in Fig. 38.5 will have a maximum in the +z direction. kd 0 cosqa= q a = ? è ? ? ? ÷ - cos 1 0 kd fae mn jnkd jmkd m M n N zy (,) [( cos ) ( sinsin )] qj qa qja = -+ - = - = - ?? 00 0 1 0 1 ¢ = ? è ? ? ? ÷ ? è ? ? ? ÷ ? è ? ? ? ÷ ? è ? ? ? ÷ f Nkd N kd Mkd M kd (,) sin cos sin cos sin sinsin sin sinsin qj q q qj qj 0 0 0 0 2 2 2 2 ? 2000 by CRC Press LLC One can increase the gain of the Yagi–Uda array by adding additional director elements. Adding additional reflector elements, however, has little effect because the field behind the first reflector element is small. Yagi–Uda arrays typically have directivities between 10 and 100, depending on the number of directors [Ramo et al., 1984]. TV antennas usually have several directors. Log-Periodic Dipole Arrays Another variation of wire antenna arrays is the log-peri- odic dipole array. The log-periodic is popular in applica- tions that require a broadband, frequency-independent antenna. An antenna will be independent of frequency if its dimensions, when measured in wavelengths, remain constant for all frequencies. If, however, an antenna is designed so that its characteristic dimensions are periodic with the logarithm of the frequency, and if the characteristic dimensions do not vary too much over a period of time, then the antenna will be essentially frequency independent. This is the basis for the log-periodic dipole array, shown in Fig. 38.6. In Fig 38.6, the ratio of successive element positions equals the ratio of successive dipole lengths. This ratio is often called the scaling factor of the log-periodic array and is denoted by (38.24) Also note that there is a mechanical phase reversal between successive elements in the array caused by the crossing over of the interconnecting feed lines. This phase reversal is necessary to obtain the proper phasing between adjacent array elements. To get an idea of the operating range of the log-periodic antenna, note that for a given frequency within the operating range of the antenna, there will be one dipole in the array that is half-wave resonant or is nearly so. This half-wave resonant dipole and its immediate neighbors are called the active region of the log-periodic array. As the operating frequency changes, the active region shifts to a different part of the log-periodic. Hence, the frequency range for the log-periodic array is roughly given by the frequencies at which the longest and shortest dipoles in the array are half-wave resonant (wavelengths such that 2L N < l < 2L 1 ) [Stutzman and Thiele, 1981]. FIGURE 38.6The log-periodic dipole array. (Source: D.G. Isbell, “Log periodic dipole arrays,” IRE Transactions on Antennas and Propagation, vol. AP-8, p. 262, 1960. With permission.) FIGURE 38.5 Three-element Yagi–Uda antenna. (Source: Shintaro Uda and Yasuto Mushiake, Yagi–Uda Antenna, Sendai, Japan: Sasaki Printing and Publishing Company, 1954, p. 100. With permission.) t= = ++ z z L L n n n n 11 ? 2000 by CRC Press LLC Defining Terms Antenna gain:The ratio of the actual radiated power per solid angle to the radiated power per solid angle that would result if the total input power were radiated isotropically. Array: Several antennas arranged together in space and interconnected to produce a desired radiation pattern. Directivity: The ratio of the actual radiated power per solid angle to the radiated power per solid angle that would result if the radiated power was radiated isotropically. Oftentimes the word directivity is used to refer to the maximum directivity. Phased array:An array in which the phases of the exciting currents are varied to scan the radiation pattern through space. Radiation pattern: A plot as a function of direction of the power per unit solid angle radiated in a given polarization by an antenna. The terms radiation pattern and antenna pattern can be used interchangeably. Related Topics 37.1 Space Propagation?69.2 Radio References C.A. Balanis, Antenna Theory Analysis and Design, New York: Harper and Row, 1982. R. Carrel, “The design of log-periodic dipole antennas,” IRE International Convention Record (part 1), 1961, pp. 61–75. R.E. Collin, Antennas and Radiowave Propagation, New York: McGraw-Hill, 1985. R.F. Harrington, Time Harmonic Electromagnetic Fields, New York: McGraw-Hill, 1961. D.E. Isbell, “Log periodic dipole arrays,” IRE Transactions on Antennas and Propagation, vol. AP-8, pp. 260–267, 1960. P. Lorrain and D.R. Corson, Electromagnetic Fields and Waves, San Francisco: W.H. Freeman, 1970. S. Ramo, J.R. Whinnery, and T. Van Duzer, Fields and Waves in Communication Electronics, New York: John Wiley & Sons, 1984. W.L. Stutzman and G.A. Thiele, Antenna Theory and Design, New York: John Wiley & Sons, 1981. S. Uda and Y. Mushiake, Yagi–Uda Antenna, Sendai, Japan: Sasaki Printing and Publishing Company, 1954. Further Information For general-interest articles on antennas the reader is directed to the IEEE Antennas and Propagation Magazine. In addition to providing up-to-date articles on current issues in the antenna field, this magazine also provides easy-to-read tutorials. For the latest research advances in the antenna field the reader is referred to the IEEE Transactions on Antennas and Propagation. In addition, a number of very good textbooks are devoted to antennas. The books by Collin and by Stutzman and Thiele were especially useful in the preparation of this section. 38.2 Aperture J. Patrick Fitch The main purpose of an antenna is to control a wave front at the boundary between two media: a source (or receiver) and the medium of propagation. The source can be a fiber, cable, waveguide, or other transmission line. The medium of propagation may be air, vacuum, water, concrete, metal, or tissue, depending on the application. Antenna aperture design is used in acoustic, optic, and electromagnetic systems for imaging, communications, radar, and spectroscopy applications. There are many classes of antennas: wire, horn, slot, notch, reflector, lens, and array, to name a few (see Fig. 38.7). Within each class is a variety of subclasses. For instance, the horn antenna can be pyramidal or conical. The horn can also have flaring in only one direction (sectoral horn), asymmetric components, shaped ? 2000 by CRC Press LLC HIGH-SPEED SPACE DATA COMMUNICATIONS ? 2000 by CRC Press LLC edges, or a compound design of sectoral and pyramidal combined. For all antennas, the relevant design and analysis will depend on antenna aperture size and shape, the center wavelength l, and the distance from the aperture to a point of interest (the range, R). This section covers discrete oscillators, arrays of oscillators, synthetic apertures, geometric design, Fourier analysis, and parameters of some typical antennas. The emphasis is on microwave-type designs. SI/TelSys Inc., Columbia, Maryland, is a company formed to commercialize NASA high-data- rate telemetry technology originally developed at Goddard Space Flight Center’s Microelectronic Systems Branch. Today, TSI/TelSys Inc. designs, manufactures, markets, and supports a broad range of commercial satellite telecommunications gateway products. These technologies and products support two-way, high-speed space data communications for telemetry, satellite remote sensing, and high-data-rate communications applications. The satellite antenna shown above is part of a system used for high-speed data transmissions. (Courtesy of National Aeronautics and Space Administration.) T The Oscillator or Discrete Radiator The basic building block for antenna analysis is a linear conductor. Movement of electrons (current) in the conductor induces an electromagnetic field. When the electron motion is oscillatory—e.g., a dipole with periodic electron motion, the induced electric field, E, is proportional to cos(wt – kx + f), where w is radian frequency of oscillation, t is time, k is wave number, x is distance from the oscillator, and f is the phase associated with this oscillator (relative to the time and spatial coordinate origins). When the analysis is restricted to a fixed position x, the electric field can be expressed as (38.25) where the phase term f now includes the kx term, and all of the constants of proportionality are included in the amplitude A. Basically, the assumption is that oscillating currents produce oscillating fields. The description of a receiving antenna is analogous: an oscillating field induces a periodic current in the conductor. The field from a pair of oscillators separated in phase by d radians is (38.26) Using phasor notation, ? E d , the cosines are converted to complex exponentials and the radial frequency term, wt, is suppressed, (38.27) The amplitude of the sinusoidal modulation E d (t) can be calculated as ê ? E d ?. The intensity is (38.28) When the oscillators are of the same amplitude, A = A 1 = A 2 , then (38.29) For a series of n equal amplitude oscillators with equal phase spacing (38.30) FIGURE 38.7 Examples of several types of antennas: (a) pyramidal horn, (b) conical horn, (c) axial slot on a cylinder, and (d) parabolic reflector. Et A t( ) cos( )=+wf Et A t A t d wf wfd( ) cos ( ) cos ( )=+++ 12 ? () () Et Ae Ae ii d ffd =+ + 12 IE A A AA==++** ** ** ? cos( ) d d 2 1 2 2 2 12 2 EtAt At At d wf wfd d wf d ( ) cos( ) cos( ) cos cos =++++ = ? è ? ? ? ÷ ++ ? è ? ? ? ÷ 2 22 Et A t j n j n d wf d( ) cos( )=++ = - ? 0 1 ? 2000 by CRC Press LLC By using phasor arithmetic the intensity is given as (38.31) where I 0 =n –2 to normalize the intensity pattern at d = 0. For an incoming plane wave which is tilted at an angle q from the normal, the relative phase difference between two oscillators is kd sinq, where d is the distance between oscillators and k is the wave number 2p/l (see Fig. 38.8). For three evenly spaced oscillators, the phase difference between the end oscillators is 2kd sinq. In general, the end-to-end phase difference for n evenly spaced oscillators is (n – 1)kd sinq. This formulation is identical to the phase representation in Eq. (38.30) with d = kd sinq. Therefore, the intensity as a function of incidence angle q for an evenly spaced array of n elements is (38.32) where L = nd corresponds to the physical dimension (length) of the aperture of oscillators. The zeros of this function occur at kL sinq = 2mp, for any nonzero integer m. Equivalently, the zeros occur when sinq = ml/L. When the element spacing d is less than a wavelength, the number of zeros for 0 < q < p/2 is given by the largest integer M such that M £ L/l. Therefore, the ratio of wavelength to largest dimension, l/L, determines both the location (in q space) and the number of zeros in the intensity pattern when d £ l. The number of oscillators controls the amplitude of the side lobes. For n = 1, the intensity is constant—i.e., independent of angle. For l > L, both the numerator and denominator of Eq. (38.32) have no zeros and as the length of an array shortens (relative to a wavelength), the intensity pattern converges to a constant (n = 1 case). As shown in Fig. 38.9, a separation of l/4 has an intensity rolloff less than 1 dB over p/2 radians (a l/2 separation rolls off 3 dB). This implies that placing antenna elements closer than l/4 does not significantly change the intensity pattern. Many microwave antennas exploit this and use a mesh or parallel wire (for polarization sensitivity) design rather than covering the entire aperture with conductor. This reduces both weight and sensitivity to wind loading. Note that the analysis has not accounted for phase variations from position errors in the element placement where the required accuracy is typically better than l/10. FIGURE 38.8A two-element and an n-element array with equal spacing between elements. The propagation length difference between elements is d sinq, which corresponds to a phase difference of kd sinq, where k is the wave number 2p/l. The length L corresponds to a continuous aperture of length nd with the sample positions beginning d/2 from the ends. ItE Ae e A e e I n I n nn iij j n in i dd fd d d d d d d () ? cos( ) cos() sin( ) sin( ) == = - - = - - = = - ? ** 2 0 1 2 2 2 0 0 2 2 1 1 1 1 2 2 / / II knd kd I kL n kL I L L n nL () sin sin sin sin sin sin sin sin sin sin sin sin q q q q q p l q p l q = ? è ? ? ? ÷ ? è ? ? ? ÷ = ? è ? ? ? ÷ ? è ? ? ? ÷ = ? è ? ? ? ÷ ? è ? ? ? ÷ 0 2 2 0 2 2 0 2 2 1 2 1 2 1 2 1 2 ? 2000 by CRC Press LLC For L > > l, sinq ? q, which implies that the first zero is at q = l/L. The location of the first zero is known as the Rayleigh resolution criteria. That is, two plane waves separated by at least l/L radians can be discriminated. For imaging applications, this corresponds roughly to the smallest detectable feature size. As shown in Fig. 38.10, the first zero occurs at approximately l/L = 0.25 radians (the Rayleigh resolution). Note that there is no side lobe suppression until d £ l, when the location of the zeros becomes fixed. Having more than eight array elements (separation of less than a quarter wavelength) only moderately reduces the height of the maximum side lobe. Synthetic Apertures In applications such as air- and space-based radar, size and weight constraints prohibit the use of very large antennas. For instance, if the L-band (23.5-cm wavelength) radar imaging system on the Seasat satellite (800-km altitude, launched in 1978) had a minimum resolution specification of 23.5 m, then, using the Rayleigh resolution criteria, the aperture would need to be 8 km long. In order to attain the desired resolution, an aperture is “synthesized” from data collected with a physically small (10 m) antenna traversing an 8-km flight FIGURE 38.9Normalized intensity pattern in decibels (10 log(I)) for a two-element antenna with spacing 4l, 2l, l, l/2, and l/4 between the elements. FIGURE 38.10Normalized intensity pattern in decibels (10 log(I)) for a length 4l array with 2, 3, 4, 5, and 8 elements. ? 2000 by CRC Press LLC path. Basically, by using a stable oscillator on the spacecraft, both amplitude and phase are recorded, which allows postprocessing algorithms to combine the individual echoes in a manner analogous to an antenna array. From an antenna perspective, an individual scattering element produces a different round trip propagation path based on the position of the physical antenna—a synthetic antenna array. Using the geometry described in Fig. 38.11, the phase is (38.33) It is convenient to assume a straight-line flight path along the x-axis, a planar earth (x, y plane), and a constant velocity, v, with range and cross-range components v r (x) and v c (x), respectively. In many radar applications the broad side distance to the center of the footprint, R, is much larger than the size of the footprint. This allows the distance R(x) to be expanded about R resulting in (38.34) The first term in Eq. (38.34) is a constant phase offset corresponding to the center of beam range bin and can be ignored from a resolution viewpoint. The second term, 2v r /l, is the Doppler frequency shift due to the relative (radial) velocity between antenna and scattering element. The third term represents a quadratic cor- rection of the linear flight path to approximate the constant range sphere from a scattering element. It is worth noting that synthetic aperture systems do not require the assumptions used here, but accurate position and motion compensation is required. For an antenna with cross range dimension D and a scattering element at range R, the largest synthetic aperture that can be formed is of dimension lR/D (the width of the footprint). Because this data collection scenario is for round trip propagation, the phase shift at each collecting location is twice the shift at the edges of a single physical antenna. Therefore at a range R, the synthetic aperture resolution is (38.35) The standard radar interpretation for synthetic apertures is that information coded in the Doppler frequency shift can be decoded to produce high-resolution images. It is worth noting that the synthetic aperture can be formed even with no motion (zero Doppler shift). For the no-motion case the antenna array interpretation is appropriate. This approach has been used for acoustic signal processing in nondestructive evaluation systems as FIGURE 38.11Synthetic aperture radar geometry and nearly orthogonal partitioning of the footprint by range (circular) and Doppler frequency (hyperbolic) contours. f p l p l () ()xRxxyz==++ 2 2 2 2 222 f p l p lll () ()tRvt Rv t v R t rc ==++ ì í ? ? ? ü y ? t ? 2 22 22 2 2 ll l R D R RD D SA == 22/ ? 2000 by CRC Press LLC well as wave migration codes for seismic signal processing. When there is motion, the Doppler term in the expansion of the range dominates the phase shift and therefore becomes the useful metric for predicting resolution. Geometric Designs The phase difference in a linear array was caused by the spatial separation and allowed the discrimination of plane waves arriving at different angles. Desired phase patterns can be determined by using analytic geometry to position the elements. For example, if coherent superposition across a wave front is desired, the wave front can be directed (reflected, refracted, or diffracted) to the receiver in phase. For a planar wave front, this corresponds to a constant path length from any point on the reference plane to the receiver. Using the geometry in Fig. 38.12, the sum of the two lengths (x, R + h) to (x, y) and (x, y) to (0, h) must be a constant independent of x—which is R + 2h for this geometry. This constraint on the length is (38.36) This is the equation for a parabola. Losses would be minimized if the wave front were specularly reflected to the transceiver. Specular reflection occurs when the angles between the normal vector N [or equivalently the tangent vector T = (x,f¢¢(x))] = (1,x/2h) and the vectors A = (0, –1) and B = (–x, h – y) are equal. This is the same as equality of the inner products of the normalized vectors, which is shown by The constant path length and high gain make the parabolic antenna popular at many wavelengths including microwave and visible. More than one reflecting surface is allowed in the design. The surfaces are typically conical sections and may be designed to reduce a particular distortion or to provide better functionality. Compound designs often allow the active elements to be more accessible and eliminate long transmission lines. A two-bounce reflector with a parabolic primary and a hyperbolic secondary is known as a Cassegrain system. In all reflector systems it is important to account for the blockage (“shadow” of the feed, secondary reflector, and support structures) as well as the spillover (radiation propagating past the intended reflecting surface). Continuous Current Distributions (Fourier Transform) Ideally, antennas would be designed using solutions to Maxwell’s equations. Unfortunately, in most cases exact analytic and numerical solutions to Maxwell’s equations are difficult to obtain. Under certain conditions, FIGURE 38.12Parabolic reflector systems: (a) geometry for determining the function with a constant path length and specular reflection, (b) single-bounce parabolic reflector, (c) two-bounce reflector with a parabolic primary and hyperbolic secondary (Cassegrain). Rhyxhy Rh x hy+-+ +-=+ = 22 24() or 2 ?? (,) (,) ?? (,) (, ) () () () TA TB ×= + ×-= - + ×= + × -- +- = -+ + = - + 2 4 01 4 2 4 4 4 4 22 22 222 2 22 2232 22 hx xh x xh hx xh xhy xhy xx h xh x xh / (38.37) (38.38) ? 2000 by CRC Press LLC approximations can be introduced that allow solution to the wave equations. Approximating spherical wave fronts as quadratics has been shown for the synthetic aperture application and is valid when the propagation distance is greater than (pL 2 /4l) 1/3 , where L is the aperture size. In general, this is known as the Fresnel or near-field approximation. When the propagation distance is at least 2L 2 /l, the angular radiation pattern can be approximated as independent of distance from the aperture. This pattern is known as the normalized far- field or Fraunhofer distribution, E(q), and is related to the normalized current distributed across an antenna aperture, i(x), by a Fourier transform: (38.39) where u = sinq and x¢ = x/l. Applying the Fraunhofer approximation to a line source of length L (38.40) which is Eq. (38.32) when n >> L/l. As with discrete arrays, the ratio L/l is the important design parameter: sinq = l/L is the first zero (no zeros for l > L) and the number of zeros is the largest integer M such that M £ L/l. In two dimensions, a rectangular aperture with uniform current distribution produces (38.41) The field and intensity given in Eq. (38.41) are normalized. In practice, the field is proportional to the aperture area and inversely proportional to the wavelength and propagation distance. The normalized far-field intensity distribution for a uniform current on a circular aperture is a circularly symmetric function given by (38.42) where J 1 is the Bessel function of the first kind, order one. This far-field intensity is called the Airy pattern. As with the rectangular aperture, the far-field intensity is proportional to the square of the area and inversely proportional to the square of the wavelength and the propagation distance. The first zero (Rayleigh resolution criteria) of the Airy pattern occurs for uL/l = 1.22 or sinq = 1.22l/L. As with linear and rectangular apertures, the resolution scales with l/L. Figure 38.13 shows a slice through the normalized far-field intensity of both a rectangular aperture and a circular aperture. The linearity of the Fourier transform allows apertures to be represented as the superposition of subapertures. The primary reflector, the obscurations from the support structures, and the secondary reflector Eu ixe dx iux () ()= ¢¢ ¢ ò 2p Eu e dx L u L u L L L iux L L ( sin) sin sin sin sin == ¢= ? è ? ? ? ÷ = ? è ? ? ? ÷ ¢ - ò q p l p l p l q p l q p l l 2 2 2 / / Euu uL uL uL uL Iuu Eu Eu RRL (,) sin sin (,) () () 12 11 11 22 22 12 1 2 2 2 = ? è ? ? ? ÷ ? è ? ? ? ÷ = p l p l p l p l and ** * Iu JuL uL C ()= ? è ? ? ? ÷ é ? ê ê ê ê ê ù ? ú ú ú ú ú 2 1 2 p l p l ? 2000 by CRC Press LLC of a Cassegrain-type antenna can be modeled. Numerical evaluation of the Fourier transform permits straight- forward calculation of the intensity patterns, even for nonuniform current distributions. Antenna Parameters Direct solutions to Maxwell’s equations or solutions dependent on approximations provide the analytic tools for designing antennas. Ultimately, the analysis must be confirmed with experiment. Increasingly sensitive radar and other antenna applications have resulted in much more attention to edge effects (from the primary aperture, secondary, and/or support structures). The geometric theory of diffraction as well as direct Maxwell solvers are making important contributions. With the diversity of possible antenna designs, a collection of design rules of thumb are useful. The directivity and gain for a few popular antenna designs are given in Table 38.1. Directivity is the ratio of the maximum to average radiation intensity. The gain is defined as the ratio of the maximum radiation intensity from the subject antenna to the maximum radiation intensity from a reference antenna with the same power input. The directivity, D, and gain, G, of an antenna can be expressed as (38.43) FIGURE 38.13Normalized intensity pattern in decibels (10 log[I(v = uL/l)]) for a rectangular and a circular antenna aperture with uniform current distributions. TABLE 38.1Directivity and Gain of Some Higher Frequency Antennas Antenna Type Directivity a Gain a Uniform rectangular aperture Large square aperture Large circular aperture (parabolic reflector) Pyramidal horn a Directivity and gain are relative to a half-wave dipole. 4 2 p l LL xy 4 2 p l LL xy 126 2 . L l ? è ? ? ? ÷ 77 2 . L l ? è ? ? ? ÷ 987 2 . D l ? è ? ? ? ÷ 7 2 D l ? è ? ? ? ÷ 4 2 p l ? è ? ? ? ÷ LL xy 05 4 2 . p l ? è ? ? ? ÷ LL xy DA GA em e = ? è ? ? ? ÷ ? è ? ? ? ÷ 44 22 p l p l and = ? 2000 by CRC Press LLC where A em is the maximum effective aperture and A e is the actual effective aperture of the antenna. Because of losses in the system, A e = kA em , where k is the radiation efficiency factor. The gain equals the directivity when there are no losses (k = 1), but is less than the directivity if there are any losses in the antenna (k < 1), that is, G = kD. As an example, consider the parabolic reflector antenna where efficiency degradation includes ? Ohmic losses are small (k = 1) ? Aperture taper efficiency (k = 0.975) ? Spillover (feed) efficiency (k = 0.8) ? Phase errors in aperture field (k = 0.996 to 1) ? Antenna blockage efficiency (k = 0.99) ? Spar blockage efficiency (k = 0.994) Each antenna system requires a customized analysis of the system losses in order to accurately model performance. Defining Terms Antenna: A physical device for transmitting or receiving propagating waves. Aperture antenna: An antenna with a physical opening, hole, or slit. Contrast with a wire antenna. Array antenna: An antenna system performing as a single aperture but composed of antenna subsystems. Directivity: The ratio of the maximum to average radiation intensity. Fraunhofer or far field: The propagation region where the normalized angular radiation pattern is indepen- dent of distance from the source. This typically occurs when the distance from the source is at least 2L 2 /l, where L is the largest dimension of the antenna. Fresnel or near field: The propagation region where the normalized radiation pattern can be calculated using quadratic approximations to the spherical Huygens’ wavelet surfaces. The pattern can depend on distance from the source and is usually valid for distances greater than (p/4l) 1/3 L 2/3 , where L is the largest dimension of the antenna. Gain: The ratio of the maximum radiation intensity from the subject antenna to the maximum radiation intensity from a reference antenna with the same power input. Typical references are a lossless isotropic source and a lossless half-wave dipole. Oscillator: A physical device that uses the periodic motion within the material to create propagating waves. In electromagnetics, an oscillator can be a conductor with a periodic current distribution. Reactive near field: The region close to an antenna where the reactive components of the electromagnetic fields from charges on the antenna structure are very large compared to the radiating fields. Considered negligible at distances greater than a wavelength from the source (decay as the square or cube of distance). Reactive field is important at antenna edges and for electrically small antennas. Related Topic 37.1 Space Propagation References R. Feynman, R.B. Leighton, and M.L. Sands, The Feynman Lectures on Physics, Reading, Mass.: Addison-Wesley, 1989. J.P. Fitch, Synthetic Aperture Radar, New York: Springer-Verlag, 1988. J.W. Goodman, Introduction to Fourier Optics, New York: McGraw-Hill, 1968. H. Jasik, Antenna Engineering Handbook, New York: McGraw-Hill, 1961. R.W.P. King and G.S. Smith, Antennas in Matter, Cambridge: MIT Press, 1981. J.D. Krause, Antennas, New York: McGraw-Hill, 1950. Y.T. Lo and S.W. Lee, Antenna Handbook, New York: Van Nostrand Reinhold, 1988. ? 2000 by CRC Press LLC A.W. Rudge, K. Milne, A.D. Olver, and P. Knight, The Handbook of Antenna Design, London: Peter Peregrinus, 1982. M. Skolnik, Radar Handbook, New York: McGraw-Hill, 1990. B.D. Steinberg, Principles of Aperture & Array System Design, New York: John Wiley & Sons, 1976. Further Information The monthly IEEE Transactions on Antennas and Propagation as well as the proceedings of the annual IEEE Antennas and Propagation International Symposium provide information about recent developments in this field. Other publications of interest include the IEEE Transactions on Microwave Theory and Techniques and the IEEE Transactions on Aerospace and Electronic Systems. Readers may also be interested in the “IEEE Standard Test Procedures for Antennas,” The Institute for Electrical and Electronics Engineers, Inc., ANSI IEEE Std. 149-1979, 1979. 38.3Microstrip Antennas David M. Pozar Introduction Microstrip antenna technology has been the most rapidly developing topic in the antenna field in the last 15 years, receiving the creative attentions of academic, industrial, and government engineers and researchers throughout the world. As a result, microstrip antennas have quickly evolved from a research novelty to com- mercial reality, with applications in a wide variety of microwave systems. Rapidly developing markets in personal communications systems (PCS), mobile satellite communications, direct broadcast television (DBS), wireless local-area networks (WLANs), and intelligent vehicle highway systems (IVHS) suggest that the demand for microstrip antennas and arrays will increase even further. Although microstrip antennas have proved to be a significant advance in the established field of antenna technology, it is interesting to note that it is usually their nonelectrical characteristics that make microstrip antennas preferred over other types of radiators. Microstrip antennas have a low profile and are light in weight, they can be made conformal, and they are well suited to integration with microwave integrated circuits (MICs). If the expense of materials and fabrication is not prohibitive, they can also be low in cost. When compared with traditional antenna elements such as wire or aperture antennas, however, the electrical performance of the basic microstrip antenna or array suffers from a number of serious drawbacks, including very narrow bandwidth, high feed network losses, poor cross polarization, and low power-handling capacity. Intensive research and development has demonstrated that most of these drawbacks can be avoided, or at least alleviated to some extent, with innovative variations and extensions to the basic microstrip element [James and Hall, 1989; Pozar and Schaubert, 1995]. Some of the basic features of microstrip antennas are listed below: ?Low profile form factor ?Potential light weight ?Potential low cost ?Potential conformability with mounting structure ?Easy integration with planar circuitry ?Capability for linear, dual, and circular polarizations ?Versatile feed geometries Basic Microstrip Antenna Element The basic microstrip antenna element is derived from a l g /2 microstrip transmission line resonator [Pozar, 1990]. It consists of a thin metallic conducting patch etched on a grounded dielectric substrate, as shown in Fig. 38.14. This example is shown with a coaxial probe feed, but other feeds are possible, as discussed below. ? 2000 by CRC Press LLC The patch has a length L along the x-axis, and width W along the y-axis. The dielectric substrate has a thickness d and a dielectric constant e r , and is backed with a conducting ground plane. With a coaxial probe feed, the outer conductor of the coaxial line is connected to the ground plane, and the inner conductor is attached to the patch element. The position of the feed point relative to the edge of the patch controls the input impedance level of the antenna. In operation, the length of the patch element is approximately l g /2, forming an open- circuit resonator. Because the patch is relatively wide, the patch edges at x = –L/2 and L/2 effectively form slot apertures which radiate in phase to form a broadside radiation pattern. Many analytical models have been developed for the impedance and radiation properties of microstrip antennas [James and Hall, 1989], but most of the qualitative behavior of the element can be demonstrated using the relatively simple transmission line model. As shown in Fig. 38.15, the patch element is modeled as a length, L, of microstrip transmission line of characteristic impedance Z 0 . The characteristic impedance of the line can be found using simple approximations [Pozar, 1990] and is a function of the width, W, of the line as well as the substrate thickness and dielectric constant. The ends of the transmission line are terminated in FIGURE 38.14 Geometry of rectangular coaxial probe-fed microstrip antenna. FIGURE 38.15 Transmission line circuit model for a rectangular microstrip antenna. The feed point is positioned a distance s from the radiating edge of the patch. ? 2000 by CRC Press LLC admittances, Y = G + jB, where the conductance serves to model the radiation from the ends of the line, and the susceptance serves to model the effective length extension of the line (due to fringing fields). Several approximations are available for calculating the end admittances [James and Hall, 1989], with a typical result for d << l 0 given as (38.44) where k 0 = 2p/l 0 and h = . The susceptance B is typically positive, implying a capacitive end effect. This means that the resonant length of the patch will be slightly less than l g /2. If the feed probe is located a distance s from the edge of the patch, the input impedance seen by the probe can be calculated using basic transmission line theory from the circuit of Fig. 38.15. Resonance is defined as the frequency at which the imaginary part of the input impedance is zero. As a result of the symmetry of the transmission line resonator, the voltage along the transmission line will have maxima at the ends and a null at the center of the line. This implies that the input impedance will be maximum when the feed point is at the edge of the patch, and will decrease to zero as the feed is moved to the center of the patch. Fig. 38.16 shows a Smith chart plot of the input impedance of a coaxial probe-fed microstrip antenna vs. frequency, for three different probe positions. Observe that the input impedance locus decreases as the feed point moves toward the center of the patch. Also, observe that the impedance locus becomes more inductive as the feed point moves toward the center of the patch. The far-field radiation patterns can also be derived from the transmission line model by treating the radiating edges at x = –L/2 and L/2 as equivalent slots. In the coordinate system of Fig. 38.14, the normalized far-zone fields of the rectangular patch can be expressed as FIGURE 38.16 Smith chart plot of the input impedance of a probe-fed rectangular microstrip antenna vs. frequency, for three different feed positions. Patch parameters are L = 2.5 cm, W = 3.0 cm, e r = 2.2, d = 0.79 cm. Frequency sweep runs from 3.6 to 4.25 GHz, in steps of 50 MHz. YGjB kW jkd=+= +- ( ) [] 0 0 0 2 11064 h .ln m 0 0 e ? 2000 by CRC Press LLC (38.45a) (38.45b) where and q and f are spherical coordinates. These patterns have maxima broadside (q = 0) to the patch, with 3-dB beamwidths typically in the range of 90° to 120°. Typical E-plane (f = 0) and H-plane (f = 90°) microstrip antenna radiation patterns are shown in Fig. 38.17. Microstrip antenna elements have a number of useful and inter- esting features, but probably the most serious limitation of this technology is the narrow bandwidth of the basic element. While antenna elements such as dipoles, slots, and waveguide horns have operating bandwidths ranging from 15 to 50%, the traditional microstrip patch element typically has an impedance bandwidth of only a few percent. Fig. 38.18 shows the impedance bandwidth vs. substrate thickness for a rectangular microstrip antenna with sub- strate permittivities of 2.2 and 10.2. Observe from the figure that bandwidth decreases as the substrate becomes thinner and as the dielectric constant increases. Both of these trends are explained as a result of the increased Q of the resonator, basically due to the fact that the patch current is in close proximity to its negative image in the substrate ground plane. In terms of bandwidth, it is preferable to use a thick antenna substrate, with a low dielectric constant. But because of inductive loading and possible spurious radiation from coplanar microstrip circuitry, the thickness of a microstrip antenna substrate is typically limited to 0.02l or less. This illustrates the essential compromise associated with the microstrip antenna concept, as it is not possible to obtain optimum performance from both a microstrip antenna and microstrip circuitry on a single dielectric substrate. These two functions are distinct electromagnetically, since the bound fields associated with nonra- diating circuitry obviate efficient radiation. While the bandwidth of the basic element is limited, considerable research and development during the last 15 years has led to a number of creative and novel techniques for the enhancement of microstrip antenna bandwidth, so that impedance bandwidths ranging from 10 to 40% can now be achieved [James and Hall, 1989; Pozar and Schaubert, 1995]. While there have been dozens of proposed techniques for the enhancement of microstrip antenna bandwidth, they can all be categorized according to three canonical approaches: ?Impedance matching using matching network ?Introducing dual resonance with stacked or parasitic elements ?Reducing efficiency by adding lossy elements The reader is referred to the literature for more details on specific techniques for bandwidth improvement. Figure 38.18 also shows the efficiency of the antenna, defined as EE q a a bf= 0 sin cos cos EE f a b bqf= 0 sin cos cos sin aqf bqf = = kW kL 0 0 2 2 sin sin sin cos FIGURE 38.17 E- and H-plane far-field radiation patterns of the rectangular micros- trip antenna of Fig. 38.16. ? 2000 by CRC Press LLC (38.46) where P rad is the radiated power and P loss is the power lost in the antenna. Losses in a microstrip antenna occur in three ways: conductor loss, dielectric loss, and surface wave loss. Unless the substrate is extremely thin, conductor loss is generally negligible. For quality microwave substrates (loss tangent £ 0.002), dielectric loss is also relatively small. Surface waves, which are fields bound to the dielectric substrate that propagate along its surface, often account for the dominant loss mechanism for microstrip antennas. As can be seen in Fig. 38.18, efficiency decreases with increasing substrate thickness and dielectric constant, again suggesting the use of low- dielectric-constant substrates. The overall radiation efficiency of a microstrip antenna on a low-dielectric substrate is typically 95%, or better. Besides rectangular patch elements, it is possible to use a variety of other patch shapes as resonant radiating elements. For purposes of polarization purity and analytical simplicity, however, it is usually preferable to use rectangular, square, or circular elements. Linear polarization is best obtained with rectangular elements, while dual linear or circular polarization can be obtained with square or circular patch elements [James and Hall, 1989; Pozar and Schaubert, 1995]. Feeding Techniques for Microstrip Antennas While Fig. 38.14 shows a coaxial probe-fed microstrip antenna element, it is also possible to feed the patch element by several other methods. Fig. 38.19a shows a rectangular patch element fed with a microstrip trans- mission line coplanar with the patch element. The amount of inset of the feed line controls the input impedance level at resonance, in a manner analogous to the positioning of the coax probe feed for impedance control. The equivalent circuit of the antenna near resonance is also shown in the figure. The patch appears as a parallel RLC resonant circuit, with a series inductance that represents the near-field effect of the microstrip feed line. (The same equivalent circuit applies to the probe-fed microstrip antenna.) Both the probe feed and the line feed excite the patch element through coupling between the equivalent J z electric current of the feed and the E z directed field of the patch resonator, which has a maximum below the center of the patch. The direct-contacting coax probe and inset microstrip line feeds have the advantage of simplicity, but suffer from some disadvantages. First, bandwidth is limited because of the requirement of a thin substrate, as discussed above. In addition, the inherent E-plane asymmetry of these feeds generates higher-order modes which lead FIGURE 38.18 Impedance bandwidth and radiation efficiency of a microstrip antenna vs. substrate thickness, for two values of substrate permittivity. e P PP = + rad rad loss ? 2000 by CRC Press LLC to cross polarization. And, in the case of the coax feed, the need for soldering can decrease reliability and increase cost if a large number of elements are involved. It is also possible to feed a microstrip antenna element using noncontacting feeds of various forms. Fig. 38.19b shows a proximity feed, where a two-layer substrate houses an embedded microstrip transmission feed line, with the radiating patch located on the top of a substrate layer placed over the microstrip feed line. The feed line is terminated in an open-circuited stub below the patch. Proximity coupling (often referred to in the literature by the less-descriptive term electromagnetic coupling) has the advantage of allowing the patch to reside on a relatively thick substrate, for enhanced bandwidth, while the feed line sees an effectively thinner substrate, FIGURE 38.19 Three types of feeding methods for rectangular microstrip antennas, and their associated equivalent circuits: (a) patch fed with an inset microstrip transmission line, (b) patch fed by proximity coupling to a microstrip transmission line, (c) patch fed by aperture coupling to a microstrip transmission line. ? 2000 by CRC Press LLC which is preferred to minimize spurious radiation and coupling. Fabrication is a bit more difficult than the single-layer coax or line feed, because of the requirement of bonding and aligning two substrates. The equivalent circuit of the proximity-coupled element is shown in Fig. 38.19b, where the series capacitor is indicative of the capacitive nature of the coupling between the open-ended microstrip line and the patch element. Another type of noncontacting feed is the aperture-coupled microstrip antenna shown in Fig. 38.19c [James and Hall, 1989; Pozar and Schaubert, 1995]. This configuration consists of two parallel substrates separated by a ground plane. A microstrip feed line on the bottom of the bottom substrate is coupled through a small aperture (typically a thin slot) in the ground plane to a microstrip patch element on the top of the top substrate. This arrangement allows a thin, high-dielectric-constant substrate to be used for the feed line, and a thicker, low-dielectric-constant substrate to be used for the radiating element. In this way, the two-layer design of the aperture-coupled element allows the substrates to be optimized for the distinct functions of circuit components and radiating elements. In addition, the ground plane provides isolation between the radiating aperture and possible spurious radiation or coupling from the feed network. An important aspect of the aperture-coupled approach is that the coupling aperture is below resonant size, so that the backlobe radiated by the slot is typically 15 to 20 dB below the main forward beam. The aperture-coupled geometry affords several degrees of freedom for control of the electrical properties of the antenna. The slot size (length) primarily determines the coupling level and, hence, the input impedance. Tightest coupling occurs when the slot is centered below the patch, with the input impedance decreasing as the slot size is decreased. As with any microstrip antenna, the resonant frequency is controlled primarily by the length of the patch. The feed line stub length can be used to adjust the reactance loading of the element, and the antenna substrate thickness and dielectric constant has a direct effect on the bandwidth of the element. It is also possible to use the slot to provide a double tuning effect to increase significantly impedance bandwidth. Such aperture-coupled antennas have been demonstrated with bandwidths up to 40% [Pozar and Schaubert, 1995]. Another unique feature of the aperture coupled patch is that the principal plane patterns have theoret- ically zero cross polarization because of the symmetry of the element. Microstrip Antenna Arrays One of the most useful features of microstrip antenna technology is the ease with which array antennas can be constructed, since the feed network can be fabricated with microstrip transmission lines and microstrip circuit components at the same time as the microstrip radiating elements. This eliminates the cumbersome and expensive coaxial or waveguide feed networks that are necessary for other types of arrays. In fact, microstrip technology offers such versatility in design that a wide variety of series-fed, corporate-fed, fixed-beam, scanning, multilayer, and polarization agile microstrip arrays have been demonstrated, many examples of which can be found in the literature [James and Hall, 1989; Pozar and Schaubert, 1995]. One of the most convenient architectures for microstrip arrays is the single-layer design where the microstrip feed lines are printed on the same substrate layer as the radiating patch elements. This results in a simple, low- profile, inexpensive, and easily fabricated antenna assembly. An example of a 2 ′ 4 microstrip array using this type of configuration is shown in Fig. 38.20. The microstrip feed network consists of a main feed line driving three levels of coplanar two-way power dividers, which in turn drive eight edge-fed patches. This is an example of a corporate feed network, in contrast to a series type of feed where array elements are tapped off of a single microstrip line. The corporate feed provides good bandwidth and allows precise control of element excitation, but requires considerable substrate area and can be lossy. A series feed can be very compact and efficient, but its bandwidth is typically limited to a few percent. Both types of feeds can be used for single and dual polarizations. More flexibility for feed network layout can be obtained by using a two-sided aperture-coupled patch geometry. This allows the feed network to be isolated from the radiating aperture by the ground plane, and the extra substrate area can be very useful for arrays that require dual polarization or dual-frequency operation. Similar features can be obtained by using feedthrough pins or vias, but the added fabricational complexity of solder connections can be formidable in a large array. A serious limitation of microstrip array technology is that array gain is limited by the relatively high losses of microstrip transmission lines. While array directivity increases with the area of the radiating aperture, the losses of the feed network increase exponentially with array size. This is especially serious at higher frequencies, ? 2000 by CRC Press LLC where it is usually impractical to design a microstrip array with a coplanar feed network for gains in excess of 28 to 32 dB. Off-board feed networks or multilayer designs can be used to partially circumvent this problem, but at the expense of simplicity and cost. Computer-Aided Design for Microstrip Antennas Ideally, antenna CAD software would combine a user-friendly interface with a computationally efficient set of accurate and versatile theoretical models. While software with such features has reached a fairly high level of refinement for the analysis of low-frequency and microwave circuit analysis and optimization, the development of microstrip antenna CAD software lags far behind. One reason is the economic reality that the market for antenna software is relatively small, which perhaps explains why there is very little commercially available antenna CAD software of any type. Microstrip antenna CAD software development has also been slow because of the fact that such antennas are relatively new, receiving serious attention only during the last 15 years. Furthermore, microstrip antenna geometries are relatively difficult to model because of the presence of dielectric inhomogeneities and a wide variety of feeding techniques and other geometric features. This last consideration makes the development of a general-purpose microstrip antenna analysis package extremely difficult. It may come as a surprise to the newcomer to practical antenna development, but it must be realized that many microstrip antenna designs have been successfully completed with little or no CAD support. There are, however, many situations where antennas and arrays can be designed more effectively, with better performance and less experimental iteration, when proper CAD software tools are available. And there are situations involving large arrays of microstrip elements which critically rely on the use of CAD software for successful design. Thus, CAD software is not absolutely necessary for all facets of microstrip antenna design work, but good software tools can be very useful for dealing with the more-complicated microstrip geometries. Another point that seems to be especially true for antenna design in general is that CAD software, no matter how versatile or accurate, cannot substitute for experience and understanding of the fundamentals of antenna operation. Further discus- sion of microstrip antenna CAD issues can be found in [Pozar and Schaubert, 1995]. Defining Terms Array antenna: A repetitive grouping of basic antenna elements functioning as a single antenna with improved gain or pattern characteristics. Bandwidth: The fractional frequency range over which the impedance match or pattern qualities of an antenna meet a required specification. Beamwidth: The angular width of the main beam of an antenna, typically measured at either –3 or –10 dB below beam maximum. Microstrip line: A planar transmission line consisting of a conducting strip printed (or etched) on a grounded dielectric substrate. Microwave integrated circuit (MIC): A microwave or RF subsystem formed by the monolithic or hybrid integration of active devices, transmission lines, and related components. FIGURE 38.20 A corporate-fed eight-element microstrip array antenna. ? 2000 by CRC Press LLC Related Topics 37.1 Space Propagation ? 37.2 Waveguides References J. R. James and P. S. Hall, Eds., Handbook of Microstrip Antennas, London: Peter Peregrinus (IEE), 1989. D. M. Pozar, Microwave Engineering, Reading, Mass: Addison-Wesley, 1990. D. M. Pozar and D. H. Schaubert, Microstrip Antennas: The Analysis and Design of Microstrip Antennas and Arrays, New York: IEEE Press, 1995. Further Information The most up-to-date information for developments in the field of microstrip antennas can be found in the technical journals. These include the IEEE Transactions on Antennas and Propagation, the IEE Proceedings, Part H, and Electronics Letters. There are also a large number of symposiums and conferences on antennas that usually emphasize practical microstrip antenna technology, such as the IEEE International Symposium on Antennas and Propagation, the International Symposium on Antennas and Propagation (ISAP–Japan), and the International Conference on Antennas and Propagation (ICAP–Great Britain). Coverage of basic antenna theory and design can be found in the previous sections in this chapter, as well as the references listed there. ? 2000 by CRC Press LLC