Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

IEEE TRANSACTIONS ON ANTENNAS AND PROPAGATION, VOL. 49, NO.

3, MARCH 2001

393

Greens Function Calculations on Circular Microstrip Patch Antennas


Hoton How, Carmine Vittoria, Fellow, IEEE, Leo C. Kempel, Senior Member, IEEE, and Keith D. Trott, Senior Member, IEEE

AbstractGreens function calculations have been applied to solve the radiation problem of circular microstrip patch antennas. The patch is fed by either a microstrip line or a coax line to be approximated as delta-function-like sources. The resultant Galerkin elements involve only one-fold Sommerfeld-type integrals. Resonant frequency, impedance, and far-field pattern have thus been calculated, which compared nicely with measurements. Index TermsAntenna impedance, antenna radiation pattern, dyadic Greens function, microstrip patch antenna, Sommerfeld integral.

I. INTRODUCTION OR microstrip antennas, analytical models may be formulated in terms of coupled integral equations [1], or dyadic Greens functions [2]. For the integral equation method, the circular patch antenna problem has been formulated using vector Hankel transforms in the regions of thin and thick dielectric substrates [3], [4]. These analyses agree with the cavity-model approximation when the dielectric substrate layer is thin, and it deviates from the cavity-model approximation when the substrate layer is thick, indicating that magnetic wall boundary conditions are inappropriate for a patch antenna fabricated on top of a thick dielectric substrate [5]. Dyadic Greens function calculations have been applied to rectangular patch antennas [6]. For circular and elliptical microstrip patches, the Greens function method has been utilized assuming the dominant-mode approximation [7], [8]. In this paper, we consider the Greens function calculations for microstrip antennas possessing a circular geometry without assuming such an approximation. This allows us to calculate the engineering parameters of the antennas to be excited away from the resonance conditions, for example. In the past, we have modified the Greens functions to include dielectric and conductor losses [9]. Dielectric loss can be included in a straightforward manner by allowing the dielectric constant to be a complex number containing the dielectric loss tangent in the imaginary part. Conductor loss is added to the Greens function by introducing finite surface impedance to the metal
Manuscript received January 8, 1999; revised March 6, 2000. This work was supported by AFOSR/NM. H. How is with ElectroMagnetic Applications, Inc., Boston, MA 02109 USA (e-mail: hhow@neu.edu). C. Vittoria is with the ECE Department, Northeastern University, Boston, MA 02115 USA. L. C. Kempel is with the ECE Department, Michigan State University, East Lansing, MI 48824 USA (e-mail: kempel@egr.msu.edu). K. D. Trott is with Mission Research Corporation, Valparaiso, FL 32580 USA. Publisher Item Identifier S 0018-926X(01)00546-4.

patch and Wheelers incremental impedance to the ground plane [8]. As such, surface-pole singularities are removed from the real axis and the Sommerfeld-type integrals can be processed as proper integrals. Also, we have shown in the past that a convenient basis for currents on the metal patch can be derived from current potentials involving normal modes of the patch geometry [9]. By using current potentials, the symmetry of the antenna is reserved, and the resultant Galerkin elements reduce to forms involving one-fold Sommerfeld-type integrals. As such, the model facilitates greatly numerical computation upon which many important engineering parameters of the antenna can be readily calculated. In this paper, we consider the circular microstrip patch antenna to be fed by two kinds of feeder lines, a microstrip line located at the edge of the metal patch or a coax line deployed directly under the patch. Both of these current sources are assumed to have delta-function-like distributions. Radiation patterns, input impedance, and radiation frequencies have thus been calculated, which compared nicely with measurements. Under coax excitation, the coupling between the antenna circuit and the feeder-line inductance is discussed to illustrate a critical feeder-line position beyond which no more resonant radiation from the antenna can be made possible. II. FORMULATION We consider the radius of the circular metal patch to be , which is deposited on top of a dielectric substrate with a thickness and a dielectric constant . The lateral dimension of the substrate is considered to be infinitely wide so that the previously derived Greens functions can be used [9], [10]. The metal patch and the ground plane are of the same conductivity , and . In this paper, the loss tangent of the substrate is denoted as we assume copper is used as the conductor whose thickness is negligible comparing to the substrate thickness . The electric field at position produced by a unit current dipole source located at position is referred to as the dyadic . The total electric field arising from a Greens function is, therefore current distribution

(1) where given by , and is

(2)

0018926X/01$10.00 2001 IEEE

394

IEEE TRANSACTIONS ON ANTENNAS AND PROPAGATION, VOL. 49, NO. 3, MARCH 2001

Here

and (12)(14) determine the normalization constants

as

(3) (4) (5) (6) and and are given by (7) In (5) and (6), is the rf penetration skin depth in the ground is the surface impedance of the metal patch, given plane and by, respectively

and (15) For this employed current basis, denotes the angular modes and denotes the radial modes. We note that the current basis functions given by (12) are similar to the waveguide vector modal functions used in [7]. Using the Galerkin method, we obtain, from (10)

(8) The complex dielectric constant of the substrate is (9) which characterizes the lossy property of the substrate material. The finite conductivity of the metal patch gives rise to finite in (8), and the finite conductivity of the surface impedance ground plane results in shift of the transverse-magnetic (TM) surface-wave poles in the integrand of (1), as shown in (5) and (6) above, [9], [10]. Under external driving, we derive (10) where is the electric field generated by the excitation current . Let the patch current be expressed in terms of a curas rent basis (11) s are derived from the normal-mode current and potentials [10] (12) where (19) denotes the electric field generated by the In (18), . That is, is derived from (1) basis current as the source current in the integrand. using Input impedance can be calculated from (20) (13) Note that the current basis has been normalized so that (14) which can be written in the following form: (21) (16) In deriving (16), the reciprocity theorem of electrodynamics has been used [2], [11]. A detailed derivation of (16), together with a discussion of the singular properties arising from metal patch thickness, can be found in [5]. This implies the following matrix equation: (17) where (18)

and denotes the th root of the derivative of Bessel function : of order ,

HOW et al.: GREENS FUNCTION ON CIRCULAR MICROSTRIP PATCH ANTENNAS

395

provided that unit excitation current is used (22) Again, the reciprocal theorem has been used in deriving (21) is the patch current induced [2], [11]. In (20) and (22), is the electric field by the excitation current, , and denotes the cross-sectional area of generated by pointing the feeder line at the patch plane with area element along a unit normal direction . Up to first-order approximation, we consider the following coax drive current: (23) whose Here, we have assumed the feeder is located at thickness is infinitesimally narrow. For the microstrip feeder, the drive current is (24) We assume the current is uniformly distributed across the strip width of half suspension-angle ; the current distribution funcused in (24) is, therefore tion if otherwise
Fig. 1. Schematic drawing of a circular patch antenna shows the polar coordinates  and . Microstrip feeder current and its associated Huygens currents are also shown.

(25)

Both drive currents given in (23) and (24) are of unit magnitude. matrix (18) can For both forms of excitation, the resultant be cast in the following form:

(26)
Fig. 2. Input impedance loci of the cited antenna. Solid line is from calculation, and small circles represent measurements from [17].

, and the Here, it is understood that for the coax feeder, takes the limiting value of one in (26). For factor . the microstrip feeder, We note that (10) is complete in the sense that it includes the externally applied field, an Ohmic field, and a geometry-induced or image field in the presence of a dielectric substrate backed by a metal ground plane. By using the Greens function expression (1), (10) is compatible with Maxwell equations, and by taking divergence on both sides, it implies continuity of current. This is true because continuity of current results from Coulomb and Ampere equations. By using the reciprocity theorem of electrodynamics, the source driving current is included in a volume integral (16) appearing as the inhomogeneous term in the Galerkin equation (17). Thus, (17) is exact, subject to no approximation, and it is valid under all kinds of excitation currents reinforcing continuity of current at the surface of the metal patch.

In this paper, we have, for simplicity, assumed delta-function-like source currents (23) and (24). The singular part of the feeder-line currents, as well as the surface currents on the metal patch, can all be integrated out analytically, resulting in no difficulty at all for the remaining numerical calculations. In [7], a fictitious attachment current surrounding the input feeder-line filament was used as the source current. Attachment currents of this kind resemble the Huygens currents viewed from outside of the coax filament, as discussed in [5]. If Huygens currents are used as the effective input currents, the original driving current should not be included in the surface integral evaluating the corresponding Galerkin source-elements, and vice versa. Although the authors of [7] used the Huygens currents to represent a coax feeder, we have, in this paper, chosen the Huygens currents to represent a microstrip feeder (24), but they used the direct filament current as the source current for a coax feeder

396

IEEE TRANSACTIONS ON ANTENNAS AND PROPAGATION, VOL. 49, NO. 3, MARCH 2001

(23). For these reasons, both the coax and the microstrip feeders give rise to similar mathematical expressions for the Galerkin source-elements, as described in a universal form of (26). Microstrip feeder current and its associated Huygens current are shown in Fig. 1. Our calculations compared nicely with experiments, as shown in Fig. 2. The radiation field associated with a point (Hertzian) dipole located at the air-dielectric interface of a microstrip structure is [12]

zero, giving rise to floating-point errors. However, we notice that these integrals always appear in the following form: (31) even when , because which remains finite at is also a zero of of order one. Therefore, Sommerfeld-type integrals near these quasi-singular points are physically well behaved, and Taylor expansions need to be used . for (31) near The second difficulty is attributed to the simple poles in and (4), appearing as the TM and transverse-electric (TE) surface modes, respectively [9], [11]. However, by taking account of dielectric loss and conductor loss, (4) is modified as (5) and (6), whose corresponding poles are pushed away from the real axis toward the lower half-plane. As such, integration can be physically performed along the real axis. However, due to sharp cancellation of the integrands near these quasisingular points, care needs to be taken. In the following calculations, we have adopted the implicit fifth-order RungeKutta method, which is able to monitor local truncation errors, to adaptively adjust stepsize to ensure the overall accuracy [14]. Accuracy in one part per million has thus been achieved in this type of integral. The last difficulty concerns the improper integral containing Bessel functions at infinity. We may simply terminate a Sommerfeld-type integral with a large cutoff and omit the contribution of the integral from cutoff to infinity. This approach is legitimate for simple cases in which mixing of modes is not exceeding [7], [8], [10]. However, when inverting a large matrix excited away from resonance containing coupling arising from many patch antenna elements, for example, truncation errors accumulate quickly so as to overturn the general convergence if we are not able to control the accuracy of Sommerfeld-type integrals. There is a need to evaluate Sommerfeld-type integrals accurately at infinity. Bessel functions oscillate indefinitely when infinity is approached. Although the amplitude of the oscillations decreases monotonically with the argument, the oscillation period of Bessel function is not strictly , rendering the conventional extrapolation scheme very much inaccurate. This difficulty can be circumvented by considering asymptotical expansion of Bessel functions [15]:

(27) where

(28) The radiation pattern associated with a given patch current disis, therefore tribution

(29) where is the wavevector propagating along the radial direcgiven by (11), we have tion. For current distribution

(30) A derivation of the dyadic Greens function (2), together with a discussion of its physical meaning and numerical applications, can be found in [5]. III. NUMERICAL EVALUATION In evaluating Sommerfeld-type integrals [13] described in (19) and (26), we encounter three kinds of quasi-singularities corresponding to the patch resonant modes, the surface modes, and oscillations at infinity, respectively. Although these integrals are well behaved and single valued, cares need to be taken to avoid large truncation errors. The first difficulty arises from . Near the patch quasiresonant modes occurring at these points, some of the denominators in (19) and (26) go to (32) where the s denote gamma functions. As such, these integrals are converted into series containing terms of the following form: (33)

HOW et al.: GREENS FUNCTION ON CIRCULAR MICROSTRIP PATCH ANTENNAS

397

which can be readily evaluated exploiting sine and cosine integrals and their derivatives. In (33), denotes a positive integer. By keeping eight terms in (32), we have found that the improper Bessel integrals can be very accurately calculated up to 12 significant digits for a moderate choice of the cutoff value expressed in (33). can be asserted by using the conIn (33), the coefficient ventional residue calculations applied to the integrands at the such that point of infinity. Or, equivalently, when , we can use the following asymptotic expan, and , given by (2)(5), as sions for

knowledge on the integrands. It can also achieve the same degree of accuracy. However, it requires for each coefficient one extra integration along the contour around infinity. It, therefore, consumes much more computer time and, hence, is not favored. Sommerfeld-type integrals, (19) and (26), have thus been evaluated up to six significant digits. IV. RESULTS In this paper, we consider a circular microstrip patch antenna . that is excited near the fundamental mode acquire the largest amplitudes compared with As such, the other multipole fields. We consider only the dipole terms , because the other angular modes do not couple to the dipole ones through the Greens function (19). From the following calculations, we will see that contribution from highorder angular modes is not important if the patch is excited near dipole resonance, although they can be induced through feeder excitation (26). The radial modes are coupled together through the Greens function (19) responsible for the formation of a (fictitious) boundary between the resonant electromagnetic field directly under the metal patch and the surrounding surface waves and fringing fields at periphery [10]. We have used 20 radial . By doubling terms in the following calculations: the radial terms to 40, it brings about corrections in the order of a few parts per thousand if the excitation frequency is not too much different than assumed by a resonant mode. We have applied Greens function calculations to two circular microstrip patch antennas. The first antenna was cited from [16], which involves the following parameters: cm, cm, and . The half rad, suspension-angle of the microstrip feeder is resulting in 50- impedance in the present geometry. Fig. 1 shows the polar coordinates used for the patch antenna calculations. Microstrip feeder current and its associated Huygens currents are also shown in Fig. 1. The calculated resonance frequency of the fundamental mode is 0.7936 GHz, which compares exactly to the measured value of 0.794 GHz. In contrast, the cavity model calculation predicted a resonant frequency of 0.805 GHz [16]. We, therefore, conclude that at resonance surface waves and fringing fields have effectively increased the radius of the metal patch by 1.37%. The calculated input impedance of the antenna is shown in Fig. 2. It is seen in Fig. 2 that the calculated impedance loci in Smith chart compare very well with measurements that are shown as small circles cited from [16]. It is also seen in Fig. 2 that slightly larger discrepancies occur near the two ends of the impedance loci, indicating that more multipole terms need to be kept in the expansion of (17) when frequencies are shifted away from resonance. In Fig. 3, Curves (1) and (2) show the calculated radiation and in the and pattern of the antenna, planes, respectively. We note that only the copolarized radiation can be generated from the fundamental mode excitation. and The cross-polarized field cancels out for the two modes in the and plane, as can be explicitly checked from (30). We note that cross-polarized radiations can exist in a physical circular patch antenna, due to the finite

(34)

(35)

(36) As a summary, we have applied the above three techniques to facilitate the evaluation of Sommerfeld-type integrals (19), , Region I, Gaussian quadratures can (26). For be safely applied, and this region presents no difficulty at all , surfor numerical integration. In Region II, face poles appear, and high-order adaptive algorithm incorporating variable stepsize may be employed. Integration in this region consumes most of the computer time, and it also limits the overall accuracy of the integrals. Region III is defined for cutoff, which contains the quasiresonant modes of the metal patch. Again, we can apply Gaussian quadratures in this region, provided that Taylor expansion is used to evaluate (31) to avoid truncation errors when falls in the vicinities and or . of The integration cutoff can be chosen to be the largest number . For cutoff , Region IV, of asymptotic expansion of the integrands may be applied using (32), (34), (35), and (36). The expanded series are then evaluated term-by-term analytically using sine and cosine integrals and their derivatives in the form of (33). We have kept the ex. Although it is most tedious pansion terms up to the order of to perform integration in this region, the computer time here is not much compared with that consumed in evaluating integrals in Region II. Alternatively, we have also tried the other residue calculations in this region to evaluate the expansion coefficients in (33). The resultant routine is simpler, which requires less

398

IEEE TRANSACTIONS ON ANTENNAS AND PROPAGATION, VOL. 49, NO. 3, MARCH 2001

Fig. 4. Calculated radiation frequency as a function of substrate thickness (fed by microstrip line). Fig. 3. Radiation profile of the cited antenna E in the  (1), and E in the  = 90 plane, Curve (2).

=0

plane, Curve

size in ground plane/substrate, metal thickness, and so on, although their intensity is much weaker than the copolarized radiations. Our calculations shown in Fig. 3 compare nicely with measurements included in [16]. The second patch antenna considers a microstrip disk of racm, which is fed either by a coax line or a midius , crostrip line. The substrate is of a dielectric constant , whose thickness is subject to loss tangent vary. Fig. 4 shows the calculated resonant frequency of the fundamental mode as a function of the substrate thickness. In Fig. 4, as well as in Fig. 5, we have considered the patch to be fed by a rad. It is seen microstrip line of half suspension angle in Fig. 4 that the radiation frequency decreases monotonically with substrate thickness. Two linear decrease rates may be assumed in Fig. 4, associated with the initial decrease at small values and the final decrease at large values. The initial decrease of the resonant frequency can be attributed to the generation of (quasi-)static fringing field around the metal patch, and the final decrease is mainly due to the generation of surface waves. From Fig. 4, it is seen that the final decrease is slightly slower than is the initial decrease. Similar to Fig. 2, input impedance loci can be plotted for each substrate thickness. They are characterized by two parameters, the radiation resistance and the radiation linewidth. Resonance is defined when the input impedance becomes a real number, which occurs when the impedance loci intersect the real axis. Radiation resistance is referred to as the input impedance at resonance. We also define the radiation linewidth to be the difference between two frequencies at the two sides of resonance at which positions the reactance exhibits largest absolute values but opposite in sign. The calculated radiation resistance and linewidth are shown as solid circles and squares, respectively, in Fig. 5, as a function of substrate thickness, . In Fig. 5, it is seen that the radiation resistance stays relatively constant for large values of , which increases rapidly when approaches cm, the zero. Therefore, when is not too small, say, antenna can be possibly impedance matched to the feeder line by varying the port suspension angle . This is clear from the

Fig. 5. Calculated radiation resistance (solid circles) and radiation linewidth (solid squares) as a function of substrate thickness (fed by microstrip line).

fact that the calculated impedance is not very sensitive to the (21), (26). When value of , up to a factor of is varied from 0 to 0.2 rad, changes from 1 to 0.987, whereas the feeder line impedance is considerably changed from infinity to 41.9 [17]. Impedance match, thererad. fore, occurs at In practice, a microstrip patch antenna is preferred to be fed by an input line with impedance close to 50 . For this purpose, the antenna may be directly connected with a microstrip quarter-wave transformer feeder to considerably reduce the inputimpedance requirement. For example, a quarterwave feeder of impedance about 100 enables the patch antenna, which possesses an input impedance slightly above 200 , as shown in Fig. 5, to be fed with a 50- microwave source. Because a patch antenna is a narrowband device, radiation bandwidth can be sufficiently covered by the quarterwave transformer device. Alternatively, a microstrip feeder can be attached to an antenna in a manner described in [18], intruding into the interior of the antenna patch by cutting thin slots along its sides. Due to the existence of a common expression for both the microstrip and the coax feeders (26), excitation of the antenna via these two kinds of feeders are very much the same, resulting in similar functional dependence of the input impedance, for example.

HOW et al.: GREENS FUNCTION ON CIRCULAR MICROSTRIP PATCH ANTENNAS

399

Fig. 7. Calculated radiation frequency (solid circles) and radiation resistance (solid squares) as a function of coax feeder position.

Fig. 6. Input impedance loci of the antenna for several coax feeder positions.

Thus, from discussions following Fig. 7, the impedance associated with an intruded microstrip feeder is smaller than a normal feeder discussed here in Fig. 1. The condition ensuring the operation of an intruded microstrip feeder is that the slot cuts accompanying the feeder run parallel to the rf current lines on the metal patch so as not to significantly interfere with the antenna operation. In other words, an intruded microstrip feeder can only be used to excite certain radiation modes of the antenna, but not all of the modes. Fig. 5 also plots the calculated radiation linewidth as a function of the substrate thickness . It is seen in Fig. 5 that the linewidth, shown as solid squares, increases monotonically with . This increase can also be approximated by two linear rates: the initial rate is characteristic of the static fringing effect of the patch antenna, and the final rate is for the dynamic generation of surface waves. Because it is understood that the linewidth is a measure of the amount of loss occurring during resonance, the calculated slower initial rate of Fig. 5 indicates that the antenna more resembles a closed cavity when it is surrounded by fringing fields than by surface waves. The rest of the paper discusses coax-feeder excitation. The , and the parameter is defined as feeder is located at . Fig. 6 shows the calculated input impedance loci , and , Curves (1) to of the antenna for (4), respectively. For each feeder location, the two resonance frequencies shown in Fig. 6 are the resonance frequency of the probe-inductance in series with the detuned patch-resonator forming a parallel-resonant circuit, and the resonance frequency of the probe-inductance in series with the detuned patch-resonator forming a series-resonant circuit. The field distribution of the patch at these two resonance frequencies shows very little difference, and the patch is operated at the same resonance

Fig. 8.

=0

Radiation profile of the lower frequency fed at r 0:6 for E in the plane, Curve (1), and E in the  = 90 plane, Curve (2).

mode; yet more or less detuned from its resonance frequency, which occurs between these two frequencies. It is seen in Fig. 6 . that these resonance frequencies appear only if , no patch resonance can be possibly made. For In Fig. 7, the solid circles show the calculated radiation frequency as a function of or . It is seen from Fig. 7 that the , radiation frequency remains roughly constant for is further reduced. In Fig. 7, which increases rapidly when the solid squares show the calculated radiation resistance as a or . It is seen in Fig. 7 that the radiation refunction of sistance increases monotonically with . The radiation pattern and in and planes, respectively, for the is shown in Fig. 8, Curves (1) and feeder location of (2). We have also calculated the radiation linewidth as a function of . The calculated radiation linewidth remains a constant , indicating that the loss value of 1.40 GHz for characteristics of the radiator is not a sensitive function of the feeder-line location. shown in Fig. 6 The two resonant frequencies for are, respectively, 5.570 and 5.954 GHz. The lower frequency is normally referred to as the radiation frequency of the antenna,

400

IEEE TRANSACTIONS ON ANTENNAS AND PROPAGATION, VOL. 49, NO. 3, MARCH 2001

V. CONCLUSION We have formulated the microstrip patch antenna problem in a circular geometry containing microstrip or coax feeder lines. The formulation uses the modified Greens function, in which the dielectric and conductor losses have been explicitly included. The patch currents are expressed in terms of a current basis comprising normal modes of the patch. As such, the use of Galerkin method results in one-fold integrals of the Sommerfeld type. Three kinds of difficulties have been recognized and resolved during the integration of these integrals, and numerical accuracy has been achieved to a satisfactory degree. Input impedance and radiation patterns of patch antennas have thus been calculated, which compared nicely with measurements. The functional dependence of antenna performance on substrate thickness and the influence of feeder-line location have also been investigated in the paper.
Fig. 9.

=0

Radiation profile of the upper frequency fed at r 0:6 for E in the plane, Curve (1), and E in the  = 90 plane, Curve (2).

REFERENCES
[1] M. C. Bailey and M. D. Deshpande, Integral equation formulation of microstrip antennas, IEEE Trans. Antennas Propagat., vol. AP-30, p. 651, 1982. [2] J. R. James and P. S. Hall, Handbook of microstrip antennas, in Numerical Analysis of Microstrip Patch Antennas, J. R. Mosig, R. C. Hall, and F. E. Gardiol, Eds. London, U.K.: Peregrinus, 1989, ch. 8. [3] W. C. Chew and J. A. Kong, Analysis of a circular microstrip disk antenna with a thick dielectric substrate, IEEE Trans. Antennas Propagat., vol. AP-29, p. 68, 1981. [4] K. Araki and T. Itoh, Hankel transform domain analysis of open circular microstrip radiating structures, IEEE Trans. Antennas Propagat., vol. AP-29, p. 84, 1981. [5] H. How and C. Vittoria, Microstrip antennas, in Encyclopedia of Electrical and Electronics Engineering (Supplementary Volume), J. G. Webster, Ed. New York: Wiley, Mar. 2000. [6] D. M. Pozar, Input impedance and mutual coupling of rectangular microstrip antennas, IEEE Trans. Antennas Propagat., vol. AP-30, p. 1191, 1982. [7] M. C. Bailey and M. D. Deshpande, Analysis of elliptical and circular microstrip antennas using moment method, IEEE Trans. Antennas Propagat., vol. AP-33, p. 954, 1985. [8] , Analysis of finite phased arrays of circular microstrip patches, IEEE Trans. Antennas Propagat., vol. 37, p. 1355, 1989. [9] H. How and C. Vittoria, New formulation of dyadic Greens function: Applied to a microstrip line, IEEE Trans. Microwave Theory Tech., vol. 42, p. 1580, Aug. 1994. , Radiation modes in dielectric circular patch antennas, IEEE [10] Trans. Microwave Theory Tech., vol. 42, p. 1939, Oct. 1994. [11] R. F. Harrington, Time Harmonic Electromagnetic Fields. New York: McGraw-Hill, 1961. [12] N. K. Uzunoglu, N. G. Alexopoulos, and J. G. Filioris, Radiation properties of microstrip dipoles, IEEE Trans. Antennas Propagat., vol. AP-27, p. 853, 1979. [13] A. Sommerfeld, Partial Differential Equations. New York: Academic, 1962. [14] G. E. Forsythe, M. A. Malcolm, and C. B. Moler, Computer Methods for Mathematical Computations. Englewood Cliffs, NJ: Prentice-Hall, 1977. [15] M. Abramowitz and I. A. Stegun, Eds., Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. ser. Applied Mathematics. Washington, DC: National Bureau of Standards, 1964, vol. 55. [16] Y. T. Lo, D. Solomon, and W. F. Richards, Theory and experiment on microstrip antennas, IEEE Trans. Antennas Propagat., vol. AP-27, p. 137, 1979. [17] T. C. Edwards, Foundations for Microstrip Circuit Design. New York: Wiley, 1987. [18] K. R. Carver and J. W. Mink, Microstrip patch antenna technology, IEEE Trans. Antennas Propagat., vol. AP-29, p. 2, 1981.

Fig. 10. Radiation profile of the degenerate mode fed at r = 0:416 for E in the  = 0 plane, Curve (1), and E in the  = 90 plane, Curve (2).

whose radiation pattern is shown in Fig. 8. The higher resonant frequency occurs at very low resistance value, and hence, it is of very little practical use. The radiation pattern for this and in the higher resonant mode is shown in Fig. 9 for and planes, Curves (1) and (2), respectively. In comparing the radiation fields at these two resonant frequenacquires larger amplitudes for the higher cies, we find that plane near the end directions of frequency mode in the is further reduced beyond 0.6, the lower the substrate. As resonant frequency increases rapidly, as shown in Fig. 6, and , it coincides with the higher frequency mode to at form the degenerate case. The radiation pattern of this degenand in the erate mode is shown in Fig. 10 for and planes, Curves (1) and (2), respectively. Again, only copolarized radiations exist for the excitation of the present patch antenna.

HOW et al.: GREENS FUNCTION ON CIRCULAR MICROSTRIP PATCH ANTENNAS

401

[19] L. C. Shen and J. A. Kong, Applied Electromagnetism, 2nd ed. MA: PWS Engineering, 1987.

Boston,

Hoton How, photograph and biography not available at the time of publication.

Carmine Vittoria (S62M63SM83F90) photograph and biography not available at the time of publication.

Leo C. Kempel (S89M94SM99) was born in Akron, OH, in October 1965. He received the B.S.E.E. degree from the University of Cincinnati, Cincinnati, OH, in 1989 and participated in the cooperative education program at General Dynamics/Fort Worth Division. He received the M.S.E.E. and Ph.D. degrees from the University of Michigan, Ann Arbor, in 1990 and 1994, respectively. After a brief postdoctoral appointment at the University of Michigan, he joined Mission Research Corporation, Valparaiso, FL, in 1994 as a Senior Research Engineer. He led several projects involving the design of conformal antennas, computational electromagnetics, scattering analysis, and high power/ultrawideband microwaves. He joined Michigan State University as an Assistant Professor in 1998, where he is conducting research in computational electromagnetics, teaching undergraduate and graduate courses in electromagnetics, and supervising the research of several M.S. and Ph.D. students. Dr. Kempel is a member of Tau Beta Pi, Eta Kappa Nu, and Commission B of URSI. He is Chapter IV Vice-Chair for the Southeast Michigan chapter of the IEEE as well as Vendor Chair for the 2000 ACES Conference. He has organized several sessions at recent URSI and ACES meetings. He is an active reviewer for several IEEE publications as well as JEWA and Radio Science. He recently coauthored The Finite Element Method for Electromagnetics (New York: IEEE Press).

Keith D. Trott (S84M86SM91) was born in Boston, MA, on November 17, 1952. He received the B.S. degree in mathematics from State University of New York, Plattsburgh, in 1977, the M.S.E.E. degree from Syracuse University, Syracuse, NY, in 1981, and the Ph.D. degree from The Ohio State University, Columbus, in 1986, specializing in electromagnetic scattering and antennas. He is currently Group Leader and Senior Research Engineer with MRCs Electromagnetic Applications Group in Valparaiso, FL, where he conducts research in electromagnetic scattering, radar cross section, and computational electromagnetic methods to solve radiation and scattering problems. He is the Principal Investigator (PI) on several antenna and scattering projects. He has more than 30 publications. He had nearly 20 years experience in Air Force laboratories conducting research in radar cross-section phenomenology and electromagnetic scattering as it applies to the exploitation and modeling of basic radar target scattering physics. Before joining MRC, he was Chief of the Munition Sensor Technology Branch at Wright Laboratory, Armament Directorate, Eglin Air Force Base, FL, and had also been the Senior Scientist for that branch. His current research interests include: phenomenology and modeling of the sensor/target interaction for both active and passive RF; bistatic RCS prediction techniques for conducting and material bodies; material parameter modeling and characterization; and computational electromagnetics to solve radiation and scattering problems. Dr. Trott is a member of Sigma Xi and Eta Kappa Nu, and has held an appointment as Adjunct Assistant Professor at AFIT. His awards include: Air Force Materiel Command Achievement Award for Science and Technology, Armament Directorate Senior Engineer of the Year, U.S. Air Force Research and Development Award, three Rome Lab Scientific Achievement Awards, Air Force Office of Scientific Research Star Team (member), and Rome Lab Research Team Award (member). He is listed in Whos Who in Science and Engineering.

You might also like