Sunteți pe pagina 1din 69

Antenna Design Notes

Part I
Mario Orece
Dipartimento di Elettronica
Politecnico di Torino

c Copyright 2013 by the Author. All rights reserved.




Contents
1 INTRODUCTION
1.1 Summary on Maxwells equations
1.2 Denitions . . . . . . . . . . . . .
1.2.1 Antenna . . . . . . . . . .
1.2.2 Antenna matching . . . .
1.2.3 Ohmic eciency . . . . .
1.2.4 Gain . . . . . . . . . . . .
1.2.5 Directivity . . . . . . . . .
1.2.6 Radiation pattern . . . . .
1.2.7 Radiated eld . . . . . . .
1.2.8 Eective area . . . . . . .
1.2.9 Eective height . . . . . .
1.2.10 Polarization . . . . . . . .
1.2.11 Bandwidth . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

2 APERTURE ANTENNAS
2.1 General expression of the radiation . . . . . . . .
2.2 Equivalence Theorem . . . . . . . . . . . . . . . .
2.2.1 Equivalence Theorem and planar apertures
2.3 Radiation zones from an aperture . . . . . . . . .
2.4 Gain ed eciency of an aperture . . . . . . . . . .
2.5 Rectangular aperture . . . . . . . . . . . . . . . .
2.5.1 Uniform amplitude and phase . . . . . . .
2.5.2 Separable eld distributions . . . . . . . .
2.5.3 Eects of the phase error . . . . . . . . . .
2.6 Circular aperture . . . . . . . . . . . . . . . . . .
2.6.1 Uniform amplitude and phase . . . . . . .
2.6.2 Non uniform illumination . . . . . . . . .
2.7 Phase center . . . . . . . . . . . . . . . . . . . . .
2.8 Apertures with arbitrary eld distribution . . . .
2.9 Horn antennas . . . . . . . . . . . . . . . . . . . .
2.9.1 Sectoral and pyramidal horns . . . . . . .
2.9.2 Smooth circular horns . . . . . . . . . . .
1

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

1
1
3
3
4
4
5
6
6
9
10
11
12
15

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

16
16
17
19
21
23
24
24
25
28
30
31
32
36
39
41
41
48

M. Orece: Antenna design notes (2013)

2.10

2.11
2.12

2.13
2.14

2.15
2.16

2.17
2.18
2.19
2.20

2.21

2.22

2.23

2.9.3 Beam symmetry and cross polarization . . . . . . .


2.9.4 Bimodal horn . . . . . . . . . . . . . . . . . . . . .
2.9.5 Corrugated horns . . . . . . . . . . . . . . . . . . .
Geometrical optics . . . . . . . . . . . . . . . . . . . . . .
2.10.1 Introduction . . . . . . . . . . . . . . . . . . . . . .
2.10.2 Eikonal Equation . . . . . . . . . . . . . . . . . . .
2.10.3 Flux tubes . . . . . . . . . . . . . . . . . . . . . . .
2.10.4 Fermats principle . . . . . . . . . . . . . . . . . . .
2.10.5 Stationary phase method . . . . . . . . . . . . . . .
2.10.6 Reection from a surface . . . . . . . . . . . . . . .
Surface Currents Method and Physical Optics . . . . . . .
Reector antennas . . . . . . . . . . . . . . . . . . . . . .
2.12.1 Diraction from a revolution surface . . . . . . . .
2.12.2 Integration methods . . . . . . . . . . . . . . . . .
Paraboloidal antenna . . . . . . . . . . . . . . . . . . . . .
2.13.1 Paraboloid design . . . . . . . . . . . . . . . . . . .
Blocking . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.14.1 Null eld hypothesis . . . . . . . . . . . . . . . . .
2.14.2 Central blocking . . . . . . . . . . . . . . . . . . .
2.14.3 Support structures blocking . . . . . . . . . . . . .
2.14.4 Reaction on the feed . . . . . . . . . . . . . . . . .
Oset paraboloid . . . . . . . . . . . . . . . . . . . . . . .
Defocusing errors . . . . . . . . . . . . . . . . . . . . . . .
2.16.1 Axial defocusing . . . . . . . . . . . . . . . . . . .
2.16.2 Transverse defocusing . . . . . . . . . . . . . . . . .
Eects of the surface tolerances . . . . . . . . . . . . . . .
Total eciency of a reector antenna . . . . . . . . . . . .
Reector antenna feeds . . . . . . . . . . . . . . . . . . . .
Geometrical Theory of Diraction . . . . . . . . . . . . . .
2.20.1 Diracted eld at shadow and reection boundaries
2.20.2 Equivalent currents methods . . . . . . . . . . . . .
Dual reector antennas . . . . . . . . . . . . . . . . . . . .
2.21.1 Cassegrain antenna . . . . . . . . . . . . . . . . . .
2.21.2 Simplied design of a Cassegrain antenna . . . . . .
2.21.3 Twistreector . . . . . . . . . . . . . . . . . . . . .
2.21.4 Other types of reector antennas . . . . . . . . . .
Shaped beam reectors . . . . . . . . . . . . . . . . . . . .
2.22.1 Search radar pattern . . . . . . . . . . . . . . . . .
2.22.2 Empirical design of shaped beam antennas . . . . .
2.22.3 Cylindrical reector antenna . . . . . . . . . . . . .
2.22.4 Reector synthesis with Geometrical Optics . . . .
Lens antennas . . . . . . . . . . . . . . . . . . . . . . . . .
2.23.1 General concepts . . . . . . . . . . . . . . . . . . .

2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

50
54
56
66
66
67
70
72
73
76
80
82
85
86
89
93
97
97
98
100
103
106
107
107
108
111
113
114
117
124
125
126
126
130
131
134
136
137
139
142
145
149
149

M. Orece: Antenna design notes (2013)


2.23.2 Analysis of a lens antenna . . . . . . . . .
2.23.3 Lenses with refractive index < 1 . . . . . .
2.23.4 Aperture amplitude distribution . . . . . .
2.24 Travelling wave antennas . . . . . . . . . . . . . .
2.24.1 Surface wave antenna: radiation conditions
2.24.2 Leaky wave antennas: radiation conditions

3
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

149
153
153
156
157
161

Chapter 1
INTRODUCTION
1.1

Summary on Maxwells equations

Electromagnetic phenomena, of which the radiation from antennas is an important


example, are ruled by the Maxwell equations, of which we will recall briey the
fundamental aspects.
In their most general expression, Maxwells equations are given by
E =

B
M
t

(1.1)

D
+J
H=
t
with the divergence equations:

B = m
(1.2)
D = e

where the vectors E, H, J , M, D, B and the scalars e and m are quantities function
of time and space - e.g. E = E(P, t) - and represent, in the order, the following
quantities: electric eld, magnetic eld, electric current density, magnetic current
density, electric displacement, magnetic induction, electric charge density, magnetic
charge density.
Actually, some of these (magnetic current and charge density) are not real physical quantities, but ctitious quantities which can be derived by others (in particular
electric eld, charge and current) but which may signicantly simplify the calculations.
Very often antennas are placed in a linear, homogeneous and isotropic medium,
at least in their neighborhood, and with quasi-stationary signals. Actually, there are
also some important exceptions, like the antennas in plasmas, the EMP (ElectroMagnetic Pulse) simulators, the antennas made by non- homogeneous non-isotropic
non-linear material (e.g. antennas made by carbon bre), and others. However, in
1

M. Orece: Antennas 2013, Introduction

the most part of the cases, it is possible to use the assumptions of homogeneity,
linearity and isotropy, and harmonic regime, and re-write eqs. (1.1-1.2) as:
E = jH M
H = jE + J
(1.3)
H = m /
E = e /
where vectors and scalars indicate the phasors, i.e. quantities depending on time
with ejt (this dependence will be omitted hereafter); and  are respectively the
magnetic permeability and the dielectric constant, or permittivity, of the medium.
Note also that dierent fonts have been used for the quantities in eqs.(1.1-1.2) (arbitrary dependence on time) and those in eqs.(1.3), because the former represent
the instantaneous values, the latter the phasors.
Is is also known from the basic concepts of Electromagnetic Waves that, manipulating eqs.(1.1-1.2-1.3), we can derive the Wave Equations; in fact, from eqs.(1.3)
we get:
E k 2 E = jJ M
(1.4)
H k 2 H = jM J
with k 2 = 2 , where k is the propagation constant of the medium. If this is
lossless, k is real and related to the wavelength by the equation:
k = 2/

(1.5)

Considering, as an example, the rst of eqs.(1.4), this can be rewritten as


L(, j) E(P, ) = jJ eq (P, )

(1.6)

where J eq is the equivalent current


J eq = J

j
M

(1.7)

and the operator L is given by:


L(, j) = I 2 I

(1.8)

Eq.(1.8) represents an important operator relationship between the electric eld and
the current density (see g.1.1).
We recall also the boundary conditions necessary to solve the Wave Equations
in presence of dierent media:
n
(H 2 H 1 ) = J s
(E 2 E 1 ) n
= Ms

(1.9)

M. Orece: Antennas 2013, Introduction

J
L

Figure 1.1: Operator relationship between eld and current density.


where J s and M s are the electric and magnetic surface currents at the separation
surface between two homogenous media 1 e 2, and H 1,2 , E 1,2 are the magnetic and
electric eld on the separation surface, in each medium. This boundary condition is
also known as continuity of the tangential components of the electric and magnetic
elds.

1.2

Denitions

.
We give in the following a number of denitions of the most important parameters
for the electrical characterization of the antennas 1

1.2.1

Antenna

The antenna is that part of a radiofrequency (RF) electronic circuit which eects
the conversion, in both directions, between transmitted or received electromagnetic
waves, and the RF voltage or current (or power) guided along a transmission link
connected to it.
The antenna radiates (or receives) electromagnetic waves in a more or less free
space, and the presence of the near objects may modify its radiation characteristics.
On the contrary, far objects may inuence anyway the propagation characteristics
and therefore the far radiated eld.
For the description of the antenna characteristics, the preferred coordinate system is usually the spherical system (R, , ), because at large distance (which is the
typical antenna operating range) the eld radiated from an antenna can be expressed
as spherical wave, with constant amplitude and phase on the sphere R=const.
Many types of antennas can be fount in the literature or in the market, but they
can be roughly grouped in a few categories:
the aperture antennas, which are extended in two dimensions, usually large
with respect to the wavelength (up to thousands of wavelengths); for this
reasons, they are used especially at higher frequency (approximately, above
1

The antenna, being a structure often of large dimensions, and subject to various types of
environmental conditions, has also important mechanical characteristics (wind load, mechanical
resonant frequency, etc.) which, however, are in general out of the scope of this text.

M. Orece: Antennas 2013, Introduction

around 1 GHz): typical examples of antennas of this type are the horn antennas, the parabolic antennas and others;
the wire antennas, which are extended essentially in one dimension, usually of
small size, very often less than one wavelengths, so that they are used mainly
at low frequencies (approximately, below about a few GHz): typical examples
of antennas of this type are the dipoles on others derived from it;
the array antennas, which are sets of radiating elements (of one of the previous
two categories) arranged in the space to obtain particular radiation characteristics.

1.2.2

Antenna matching

From the circuit point of view, the transmitting antenna is a load, placed at the end
of a transmission line (coaxial cable, microstrip, waveguide, etc.), which absorbs a
certain amount of power. It is therefore characterized by an input impedance Z,
whose knowledge is essential for a correct design of the transmission line and of the
matching network, if any.
The value of the input impedance is strictly dependent on the position of its input
terminals. These can be physically found in case of a coaxial or a two-conductors
feeding; in case of a waveguide feeding, we shall refer to the equivalent line and to
an input (or reference) section.
Generally, however, rather than the input impedance, it is more interesting the
matching to the transmission line. Consequently, very often the antenna impedance
characteristics are given in terms of its reection coecient or of its modulus;
alternatively they can be expressed by the standing wave ratio or the return loss.
This latter is very used in practice, because it allows the immediate calculation of
the amount of power lost because of the mismatch. These quantities are related
each other by the known relationships:
S=

1 ||
1 + ||

(1.10)

RL = 20 log10 ||

1.2.3

Ohmic eciency

It is dened as the ratio between the power radiated by the antenna and the power
accepted by the antenna at its input terminals:
= Pr /Pa

(1.11)

From the circuit point of view, eq.(1.11) can be represented (neglecting possible
reactances) with a series circuit of the type shown in g.(1.2), where RL and Rr are
the resistances associated respectively to the losses and to the radiation. We obtain:

M. Orece: Antennas 2013, Introduction

Rr

Pd

Pr

Figure 1.2: Equivalent circuit for radiation resistance and loss resistance.

Rr
(1.12)
RL + Rr
The ohmic eciency is usually less than 1 when the antenna size is much smaller
than the wavelength, the radiation resistance is low (a few ohms or even less) and
comparable to (or lower than) that associated to the losses 2 . Antennas at microwave
frequency, and also at lowwer frequencies (above a few MHz) have generally  1,
except when they contain dissipative materials (e.g. lossy dielectrics). For higher
frequencies (millimeter waves ad above) the losses in conductors can be signicant,
even if with high conductivity (e.g. copper) because of the skin eect.
=

1.2.4

Gain

The gain is a function of the observation angles , around the antenna and it can
be dened for both receiving and transmitting antennas.
For a transmitting antenna the gain is dened as the ratio between the power
density generated by the antenna in a point P at large distance, with angular coordinates , , and that generated in the same point by a lossless isotropic radiator,
fed with the same power delivered to the considered antenna. This denition can
therefore be formulated as:
S(, , r)
(1.13)
G(, ) =
Pd /4r2
where S is the real part of the Poynting vector in point P . At large distance, the
power density is inversely proportional to r2 , and the gain can be rewritten, by using
the angular power density dP/d, as:
G(, ) = 4

dP (, )/d
Pd

(1.14)

2
The resistance associated to the losses can be subdivided according to the various phenomena
that generate them, i.e., essentially, Joule eect in the conductors, dissipation in dielectrics, nite
conductivity of the ground.

M. Orece: Antennas 2013, Introduction

so that

G(, )d = 4

(1.15)

In reception, the gain is dened as the ratio between the available power at the
output terminals of the antenna, when illuminated by a plane wave from a direction
, , and that available at the output terminals of an isotropic radiator, illuminated
by the same plane wave.
If the antenna, as often happens, is reciprocal, the gain in reception and transmission are the same.
The term gain without other specications indicates usually the maximum value
of the function G(, ), and it is usually expressed in decibels3 .
This denition is clearly independent on the polarization and it refers to the total
power (or power density). In practice, often there is a distinction between the power
associated to the two orthogonal components of the electric eld. Depending on the
polarization characteristics and requirements for the antenna, it can be distinguished
between direct polarization and cross polarization, as it will be explained in one of
the next paragraphs.

1.2.5

Directivity

The same denition holds as for the gain, except that, instead of the power delivered
to the antenna (Pe ), the power radiated from the antenna (Pr ) is considered:
D(, ) = 4

dP (, )/d
Pr

(1.16)

From (1.11) it follows:


G(, ) = D(, )

1.2.6

(1.17)

Radiation pattern

This term is very common in practice, and indicates the graphical (or numerical)
representation of the gain or of the directivity functions (or of a part of them), typically a particular cut for a xed value of one of the variables, in case normalized
to its maximum value; one of the most common cases is to represent the gain (directivity) functions for =cost., for varying . For example, for an antenna with a
spherical reference system as shown in g.(1.3), the most signicant diagrams are
those of G(, ) for = 0 and = /2 (see g.1.4). Such planes, in case of linear
3
The denition of gain compares essentially an antenna with the isotropic source. Another
denition of gain (a little obsolete but sometimes still used, especially for terrestrial broadcasting
antennas) refers, for historical practical reasons, to the half-wave dipole, whose gain is about 2 dB.
As a consequence, the gain of the same antenna may be expressed by two dierent numbers: the
one referred to the isotropic radiator is greater than that referred to the dipole by 2 dB. To avoid
confusion, the notation dBi e dBd (not much appreciated by the metrologists...) is used,
respectively. As a standard, however, the reference is the isotropic radiator.

M. Orece: Antennas 2013, Introduction

polarization and with a suitable choice of the reference system, are coincident with
the principal planes of the radiation pattern, which are also called E and H planes,
because they are dened respectively by the the electric (or magnetic) eld vector
and the propagation unit vector.

dS

P k

H
H
H plane

E plane

Figure 1.3: Reference system for an aperture antenna.


The radiation pattern may also have a 3-D representation, as in g.(1.5), or at
level curves as in g.(1.6). This latter is particularly used in the case of contoured
beam antennas.
In order to have a rough characterization of a radiation pattern, just a few
characteristic parameters are enough, without the need of providing, or computing,
the full pattern. The most signicant of such parameters are: the sidelobe level and
the half power beamwidth (HPBW), also called -3 dB angle since the half power
corresponds to a -3 dB level.
The sidelobe level is the level of the highest of the side lobes (usually the one
adjacent to the main beam), and it is expressed in decibels with respect to the main
beam. For more accurate characterizations, sometimes it is given as an envelope
function: as an example, 29 25 log is the envelope of all the sidelobes of some
types of antennas for earth stations in satellite communications: in this case the
reference is the isotropic radiator.
The half power beamwidth, often abbreviated with HPBW, is the full angle where
the gain is maintained within -3 dB below the maximum. Sometimes other angles
and levels are used to indicate the width of a beam (e.g. -6 dB, or -10 dB, or the
rst minimum near the main beam) but in this case this must be clearly specied.
There is also an approximate relationship between the HPBW (which may also
be dierent in the principal planes) and the gain of a lossless antenna; this is given

M. Orece: Antennas 2013, Introduction

G (dB)

10

12

14

16

18

20
50

40

30

20

10

0
(gradi)

10

20

30

40

50

Figure 1.4: Cuts in the principal planes of a radiation pattern.


by
G=

K
E H

(1.18)

where G is the maximum gain (expressed as a number, not in dB), E,H are the
HPBWs in the principal planes (e.g. the E- and H-plane or anyway in the two
symmetry planes of the main lobe), and K is a suitable constant. If the angles are
in degrees, the value of K is about 3 104 . 4
Eq.(1.18) is not valid for antennas directive in one plane only and omnidirectional
in the other plane (as a dipole or a linear array). It can be veried that for such
omnidirectional antennas the following relationship applies:
G=

Ko
0

(1.19)

with Ko 100, where 0 is the HPBW.


Exercise: Assuming that all the radiated power is all concentrated in the main beam, and this
is symmetric about the axis and suciently narrow that in the region of interest it is  sin ),
demonstrate eq.(1.18) and compute the constant K in the three following cases of dependance on
of the radiation pattern: a) quadratic in eld; b) quadratic in power; c) cosine to a power.
4

For low directivity antennas (e.g. up to 10-15 dB) the constant K is slightly higher, and it
goes to about 4 104 .

M. Orece: Antennas 2013, Introduction

10
0
10
20
30
40
10
10

5
5

0
0

5
10

10

Figure 1.5: Tri-dimensional representation of a radiation pattern (rectangular horn


without phase error).
The maximum value of the directivity, DM ax (and as a consequence of the gain,
if the ohmic eciency is known) can be computed more precisely than with eq.(1.18)
if the behavior of the directivity function is known. In fact, assuming
D = DM ax d(, )

(1.20)

where obviously the maximum value of d(, ) is 1. Integrating both sides of eq.(1.16)
on the whole solid angle we obtain
DM ax = 

(1.21)

d(, )d

Exercise: Compute the directivity of an antenna whose radiated power density is zero in the
negative z half space, and for positive z is expressed by the function d = cos , symmetric about
the z axis.

1.2.7

Radiated eld

The electromagnetic eld radiated from an antenna is computed, in general, by


using very complex formulas, obtained by inverting the operator (1.8). However, at
distances suciently large from the antenna, it can be written in formally simple
manner as:
ejkr
p(, )
(1.22)
E = V0 F (, )
r

M. Orece: Antennas 2013, Introduction

10

0.2
30
0.15
30

20

0.1
30
0.05

40

40 40
0

40

20
20

10

3
30

30
40

30
40

0.05
40

40

20

30

0.1
40

20

30
0.15
30
0.2
0.2

0.15

0.1

0.05

40
0
u

0.05

0.1

0.15

0.2

Figure 1.6: Representation of the radiation pattern of g.(1.5) through level curves.
where:
V0 is a normalization constant, having the dimensions of a voltage;
F is a complex scalar function called radiation force, which depends only on
the angular coordinates. By choosing


V0 =
we have

Z0 P e
4

|F |2 = F F = G

(1.23)

(1.24)

p(, ) is the polarization unit vector of the antenna, and in general it depends
on the observation angles.

1.2.8

Eective area

This parameter, as well as the following one (the eective height) has a particular
practical importance because it allows to transform the electromagnetic quantities
(elds, power density) in electrical circuits quantities (voltage, power).

M. Orece: Antennas 2013, Introduction

11

In reception, the eective area Ae is the quantity which, multiplied by the incident power density onto the antenna, gives the available power at the output
terminals of the antenna. In formula:
Pav = Ae S

(1.25)

under the assumption of matching between the polarization unit vectors of the
incident wave and of the receiving antenna. If this is not, the polarization mismatch
is taken into account by the dot product between such unit vectors, (
pinc pant )2 .
Moreover, it is known the relationship between the eective area and the gain,
which may be derived through the Reciprocity Theorem:
G=

4
Aeq
2

(1.26)

The eective area has the physical dimensions of square meters, which explains
the name. For aperture antennas it exists also a relationship between the eective
area and the geometric area of the aperture A, given by:
Aeq = A

(1.27)

where is the aperture eciency5 , and it has, in general, values ranging between
0.5 and 0.85, depending on the types of antennas. In some particular cases (superdirective antennas) the value of 1 may be exceeded, but in a narrow bandwidth.

1.2.9

Eective height

In reception, it is the (vector) quantity which, multiplied (with a dot product) by the
electric eld incident on the antenna, gives the open-circuit voltage at the antenna
output terminals. In formula:
V = he E i

(1.28)

The eective height has the physical dimensions of meters, which explains the
name. For simple wire antennas, there is a simple relationship between the eective
height and the length of the antenna (see Tab.1.1).
These values can be easily derived from the denition of the eective height in
transmission, which relates it to the generalized electric moment P e through the
equation
(1.29)
P e = Ihe
where I is the antenna input current, and P e is related to the radiated eld by
E=

jZ0 jkr
e
Pe
2r

(1.30)

5
Actually, for many aperture antennas, the aperture eciency is only one of the gain reduction
factors, together with others (spillover, blockage, etc.); in practice, it is more common to dene
as the antenna eciency, including all factors.

M. Orece: Antennas 2013, Introduction

12

Antenna type
he
(lenght= )
hertzian dipole

short dipole
/2
resonant dipole /2 2 /
Table 1.1: Eective height for a few simple types of wire radiators.

By comparing the available power from this equivalent source with eq.(1.25) we
may derive the relationship between the eective area and the eective height, in
case of polarization matching:
h2 Z 0
(1.31)
Aeq = e
4Rirr

1.2.10

Polarization

In the most general case, the polarization of an electromagnetic wave radiated by an


antenna (usually intended as the polarization of the electric eld) has an elliptical
polarization. In practice, however, all antennas have a nominal polarization, which
is the ideal polarization the antenna should have, depending on the design and on
the service for which it is used: usual nominal polarizations are linear (vertical,
horizontal, 45 ) or circular (right- or left-hand circular, often abbreviated with
RHC, LHC, also called respectively clockwise or counterclockwise (CW or CCW)).
However, since the actual polarization is not exactly as the nominal one, there is
also a spurious polarization component which may be non negligible, especially out
of the intended coverage region of the antenna.
We may therefore dene two eld components, the direct polarization and the
cross polarization (or also wanted and unwanted), which are the projections on the
two polarization unit vectors, indicated respectively with p and q, which correspond
respectively to the desired (or nominal) polarization, and to the spurious polarization, orthogonal to the previous one. As it will be seen in the following, their
denition can be given in dierent ways, and it is therefore important to make a few
distinctions6 .
Possible denitions of polarization
The direct and cross polarization unit vectors may be dened in dierent ways,
so that the various couples of unit vectors, although coincident in the direction of
maximum radiation, dier each other signicantly for dierent directions.
6

This subject has been widely discussed by A.C.Ludwig in The denition of cross polarization,
in IEEE Transactions on Antennas and Propagation, Jan.1973, pp.116-119, and to this paper the
reader is addressed for a thorough examination of this subject.

M. Orece: Antennas 2013, Introduction

13

Figure 1.7: First and second denition of cross polarization.


Lets consider a directive antenna (as an example, an aperture with arbitrary
shape in the x, y plane) with the maximum radiation direction along z. If the eld
on the aperture is mainly directed in one direction, e.g. along y, the radiated eld
along the axis will have essentially the same direction, as shown in g.(1.7). As a
consequence it may seem reasonable to assume:7
p = y

q = x

(1.32)

which in spherical coordinates become:


p = sin sin r + cos sin + cos
q = sin cos r + cos cos sin

(1.33)

If this denition is correct on the z axis, for other directions it is no longer acceptable because the polarization unit vectors would have radial components that the
actual radiated eld has not. As a consequence this rst denition is not applicable,
if we want a denition which can be used out of the z axis.
Considering that the radiated far eld is always transverse (perpendicular to
:
but this is not possible,
r) one could think of using as polarization vectors ,
because on the z axis (the polar axis, = 0 or ) the variable is undetermined,
and so are the unit vectors.
A second possible denition is to associate to the aperture a spherical reference
system with axis directed along y. In this case we may assume8
p = 
;
q = 
(1.34)
7
S.Silver:Microwave antennas theory and design, New York: Mc.Graw-Hill, 1949, pp.424 and
557-564.
8
V.P.Narbut, N.S.Khmelnitskaya: Polarization structure of radiation from axisymmetric reector antennas, Radio Eng. Electron. Phys., vol.15, pp.1786-1796, 1970.

M. Orece: Antennas 2013, Introduction

14

where primes indicate a spherical reference system with polar axis y (see g.1.7).
Such unit vectors, transformed in the main reference system (polar axis z), are
expressed as:
sin cos + cos
p = 
1 sin2 sin2
(1.35)
cos sin cos
q = 
1 sin2 sin2
Again, this denition is insucient in the xy plane for = /2, where the unit
vectors are not dened.
A third denition considers as direct polarization of the antenna that dened by
the usual method of measurement of the antenna radiation pattern9 .

=0
x

=0

z
p
q
p
0 q
Figure 1.8: Cuts for the radiation pattern measurement and third denition of cross
polarization.
As it is known, this latter is obtained by rotating a probe, inclined according the
direct (or cross) polarization, about the antenna under test (AUT) in various planes
at =const. (v.g.1.8); this technique is perfectly equivalent to the most usual one of
keeping xed the probe (or the probes if both polarizations are taken simultaneously)
and rotating the AUT about y for each cut, by successively increasing by (around
9

564.

See J.D.Kraus:Antennas, New York: Mc.Graw-Hill, 1950, Cap.15, or in Silver, ibid., pp.557-

M. Orece: Antennas 2013, Introduction

15

z) both the AUT and the probe. It follows that the denition of the unit polarization
vectors is:
p = cos + sin
(1.36)
q = sin + cos
A source of this type is coincident with the Huygens source.
The eld lines for the direct and cross polarizations associated to the unit vectors
of the three denitions are shown in g.(1.9).

Figure 1.9: Field lines for the three denitions of direct and cross polarization:
(above, the reference polarization, below, the cross polarization).
Exercise: Show that the eld lines of the 3rd denition are obtained by intersecting a sphere
with a bundle of planes with axis parallel to y with origin in the point 0,0,-1.

Also in the third denition it exists an indetermination point, corresponding to


= . In fact, while the singularity for = 0 can be eliminated by continuity
y for = /2, and to
(lim p = y for any ), for = we have that p tends to
0
+
y for = 0 o = .

1.2.11

Bandwidth

It is the frequency range where a certain parameter (e.g. gain, or impedance, or


sidelobe level, etc.) is maintained within the prescribed limits: consequently, there
are as many bandwidths as the parameters of interest. Particularly important is the
operating bandwidth, which is the intersection of all the bandwidths of interest for
a given antenna.

Chapter 2
APERTURE ANTENNAS
2.1

General expression of the radiation

As already mentioned in the introduction, the operator expression bounding the


electric eld and the current density, expressed by
L(, j) E(P, ) = jJ eq (P, )

(2.1)

where J eq is the equivalent current


J eq = J

j
M

(2.2)

and the operator L is given by:


L(, j) = I k 2 I

(2.3)

may be inverted in order to derive the electric eld from the currents. In eq.(2.3)
k 2 = 2 ; k is the propagation constant in the medium (which may be also complex,
with negative imaginary part)1 . The inverse relationship is clearly given by:
E(P, ) = jL1 (, j) J eq

(2.4)

It is known that eq.(2.4) may be expressed as:


E(P ) = j


V

G(P P  ) J eq (P  )dV

(2.5)

where P is the observation point, V is any volume which encloses all the sources,
P  is a generic point in V , and G is the Greens dyadic function given by:


G = I + 2 (P )
k

(2.6)

1
In practice, since the most common case is when antennas radiate in air, we will always assume

(except where otherwise indicated) k = k0 = 0 0

16

M. Orece: Antennas 2013, Aperture antennas

17

where

1 jkr
e
(2.7)
4r
is the scalar Greens function and I is the identity dyadic.
If P is at distance such that the condition |P P  | >> is satised for any P  ,
eq.(2.6) can be simplied to:
(P ) =

G(P ) I t,r (P )

(2.8)

where I t,r is the transverse identity dyadic with respect to r, given by:
I t,r = +

(2.9)

and eq.(2.5) becomes:


E(P ) = j


V

I t,r J eq (P  )

ejkr
dV
4r

(2.10)

Then, if P is at distance such that all the segments from P to any P  can be
considered as parallel, I t,r can be extracted from the integral, and, moreover, can
be assumed as practically constant in amplitude. We can therefore apply the far
eld approximation

ejkr ejk(RR)
(2.11)
=
r
R
which may be considered valid if |P P  | RF , where RF = 2D2 / is the Fraunhofer distance, and D is the maximum dimension of the volume in a direction transverse to to r. Eq.(2.5) becomes

ejkR

I t,r J eq (P  )ejkRr dV
(2.12)
4R
V
where R is the (xed)distance of the observation point P from an origin O placed
in the region of the sources (although arbitrary), and r = OP  .
Eq.(2.12) is theRadiation Integral in far eld.

E(P ) = j

2.2

Equivalence Theorem

This theorem allows to simplify the Radiation Integral, transforming a volume integral into a surface integral.
Lets consider a distribution of sources, all enclosed in a surface S, which denes
two volumes V1 e V2 (inner and outer to S) respectively with and without sources,
as in g. 2.1. Here the normal to S is directed outwards. The currents radiate a
eld distribution E 1 , H 1 in V1 and E 2 , H 2 in V2 .
If we want to replace the eld E 1 , H 1 in V1 with an arbitrary distribution (e.g.
E, H), it will be necessary, to the external eects in region 2, the introduction of

M. Orece: Antennas 2013, Aperture antennas

18

n^
Distribuzione

S
Sorgenti (J_ , H)
_
e distribuzione

Distribuzione

di campo

di campo
_E2 , H
_2

_E2 , H
_2

V2

Campo
nullo

di campo _E1, H
_1

Js= n^ x H
_

V1

V2
V1

^
M s = E_ x n

Figure 2.1: Equivalence Theorem


surface currents on S to make the the distribution E, H compatible with E 2 , H 2 .
The continuity conditions of the tangent component on S imply the presence on the
surface of a distribution of electric and magnetic surface currents given by:
Js = n
(H 2 H)S
M s =
n (E 2 E)S

(2.13)

where the values of the electric and magnetic eld are those on S. Such currents
may be interpreted as those which, together with the sources internal to V1 , radiate
a eld E, H in V1 and E 2 , H 2 in V2 .
The eld distribution E, H can be chosen arbitrarily. If we choose, for example,
E = 0, H = 0 inside S, the currents on S will be:
Js = n
(H 2 )S
M s =
n (E 2 )S

(2.14)

This theorem may be formulated in other two dierent forms:


1. Since the eld in V1 is zero, we can put in V1 , and immediately below S, a
perfect electrically conducting (PEC) surface.
The electric currents will ten be short-circuited and will not radiate; conversely, the magnetic current components will be doubled. By consequence,
the external led can be computed by using only the magnetic current, as
(omitting the sux S for the eld):
n E2
M s = 2

(2.15)

2. Dually, inserting a perfect magnetic conductor (PMC), we can obtain the eld
from the electric currents:
n H2
J s = 2

(2.16)

M. Orece: Antennas 2013, Aperture antennas

19

The equivalence theorem has, for the radiation, a similar signicance as the
Thevenin theorem in Electric Circuits, because it allows, given an arbitrarily complex structure (network), to determine electric and magnetic currents (voltage and
internal resistance for Thevenin) equivalent to it, only relatively to the external problem.

2.2.1

Equivalence Theorem and planar apertures

Despite of the advantage of replacing a volume integral with one surface integral (or
two, if the formulation with both magnetic and electric currents is chosen) a few
problems are connected with the use of the Equivalence Theorem. For example, the
choice of the arbitrary surface enclosing the sources must be done in appropriate
manner, in order to simplify the computation of the elds and of the equivalent
surface currents on it, and of the integral.
To this purpose, one of the simplest surfaces to be chosen is that consisting of a
plane and of the half sphere at innity enclosing all sources.
Thanks to the radiation conditions at the innity, the eld on the half sphere
at innity vanishes, so that the sources on one side of the plane are equivalent, for
the purpose of computing the eld on the other side, to the surface currents on the
plane, computed by means of eqs.(2.14), or equivalent forms.
Lets call E a , H a the transverse eld components on the plane. If we dene the
double Fourier Transforms of such vectors 2 as:
 

f t (kx , ky ) = fx x + fy y =

 

g t (kx , ky ) = gx x + gy y =

E a (x, y)ej(kx x+ky y) dx dy

H a (x, y)ej(kx x+ky y) dx dy

(2.17)

kx = k sin cos
ky = k sin sin
where the scalar functions fx,y , gx,y are the Fourier Transforms of the transverse
eld components of the aperture Ea,x,y , Ha,x,y .
Carrying out the computations, from (2.12) we get for the three formulations,
respectively:
JS, M S:
E =
E =
2

jkejkR
[fx
4R

cos + fy sin + Z0 cos (gy cos gx sin )]

jkejkR
[cos (fy
4R

(2.18)
cos fx sin ) Z0 (gy sin + gx cos )]

See R.E. Collin, F.J. Zucker: Antenna Theory, Part I; New York: Mc. Graw-Hill, 1969, pp.
69-74.

M. Orece: Antennas 2013, Aperture antennas


2J S :
E =

jkZ0 cos ejkR


(gy
2R

E =

jkR
0e
jkZ2R
(gy

2M S :
E =
E =

jkejkR
(fx
2R

20

cos gx sin )
(2.19)

sin + gx cos )

cos + fy sin )

jk cos ejkR
(fy
2R

(2.20)
cos fx sin )

The radiated eld from a volume of sources is therefore expressed through the
double Fourier Transforms of the tangential eld components on a plane which has
all the sources on the same side. It his however apparent that this eld distribution
is non-zero only in a limited region of the plane, in order to be, sooner or later, in
far eld conditions and therefore use eq.(2.12).
Lets see now the relationship between the eld components on the aperture,
directed along x, y, and those of direct and cross polarization, along p, q, of the
radiated eld. The polarization unit vectors, according to the 3rd Ludwigs denition, are given, for p y on the normal to the aperture, by eq.(1.36), and the
components, in matrix form, are:

Ep (, )
Eq (, )

sin

cos

cos sin

(2.21)

Replacing eq.(2.20) in eq.(2.21) we obtain:

Ep (, )
Eq (, )

2
= K cos /2

1 t2 cos 2
t2 sin 2

t2 sin 2
1 + t2 cos 2

fy (, )

(2.22)

fx (, )

where K = jejkR /R, and t = tan /2.


Eq.(2.22) is an exact expression for the far eld radiated from an aperture in its
front region, and it allows to note that, for t small, the direct and cross polarization
components are directly proportional to the Fourier transforms of the scalar components along y and x of the eld. Since the cross polarization is of particular interest
mainly in correspondence of the main beam, this means that for a large aperture
(whose main beam is narrow) there will be a low cross polarization if the eld on
the aperture is linearly polarized (for example, a rectangular horn). On the contrary, if the aperture is small and the main beam is wide (as it is, for example, for a
reector antenna feed) the relationship providing the polarization is more complex,
and consequently to minimize the cross polarization the eld on the aperture shall
not be linearly polarized.

M. Orece: Antennas 2013, Aperture antennas

2.3

21

Radiation zones from an aperture

From the approximation of eq.(2.6) to eq.(2.8) it appears that, for distances larger
than a few wavelengths from the source volume (about 10 for a good accuracy, or
even less if a lower accuracy is accepted) the radiation integral is strongly simplied:
we may therefore dene,for lower distances, a near eld zone, where the expression
of the Greens dyadic must be complete, given by eq.(2.6), and another for the eld
far from the sources.
y
P

P
r ^r

^
RR

Figure 2.2: Coordinates for the radiation from an aperture.


To analyze the possible approximations in this latter region, lets consider a
planar aperture on the xy plane as in g.(2.2). By applying the equivalence theorem,
eq.(2.5) reduces to the surface integral
E(P ) = j


S

I t,r J s,eq (P  )

ejkr
dS
4r

(2.23)

where S is the surface of the whole plane, and J s,eq (P  ) is the equivalent electric
surface current, which is obtained from eq.(2.2) using the electric and/or magnetic
currents deriving from the Equivalence Theorem. We will assume that the equivalent
current is non-zero within a nite area A of the plane.
Let x , y  be the coordinates of a point on the aperture, and x, y, z (or the corresponding r, , ) those of the observation point P .
Eq.(2.23) becomes:
jkr
Z0 
e
dS
( + ) (J (x , y  ) + J(x , y  ))
E(P ) =
2 A
r

(2.24)

The zones in which the eld can be subdivided are mathematically determined,
although without precisely dened sharp boundaries, by the approximations that
can be made in the integral (2.24) and by the structure of the eld.
In the intermediate eld zone, or Fresnel zone, we may consider negligible, in
addition to 1/r with respect to k, the change of direction of the unit vectors of the
transverse dyadic on the aperture. This corresponds to assume almost parallel, in

g.(2.2), the unit vectors r and R.

M. Orece: Antennas 2013, Aperture antennas

22

Moreover we neglect in eq.(2.24) the amplitude variation of 1/r, which is taken


equal to the constant value (for each observation point) 1/R.
Conversely, we must deal more carefully with the variation of r on the aperture,
as far as the phase term ejkr is concerned. The distance of the observation point
P from the generic source point P  is, in general:

r = (x x )2 + (y y  )2 + z 2

1/2

(2.25)

In the paraxial region, i.e. for small and z |x x |, |y y  |, eq.(2.25) may


be developed as:


(x x )2 (y y  )2
+
r =z 1+
z2
z2

1/2

(x x )2 (y y  )2
+
+ ...
=
=z+
2z
2z

(2.26)

The approximation Fresnel zone consists in neglecting the terms of order higher
than the second in the development (2.26).
An alternative for of expansion of eq.(2.25) is obtained by using spherical coordinates fore the observation point P (see g.2.2):
x = R sin cos = Ru
y = R sin sin = Rv
z = R cos

(2.27)

By introducing eqs.(2.27) in (2.25) we obtain, after having developed in Taylor


series up to the rst three terms, and neglecting the powers of 1/R of order higher
than 1 (which is equivalent to assume ux /R, vy  /R
1):
x2 + y 2 (ux + vy  )2


= R + rb
+
vy
)
+
r
R

(ux
=
2R

(2.28)

Eq.(2.24) becomes therefore:


jZ0 ejkR 
E(P )
Jt (x , y  )ejkrb dx dy 
=
2 R A

(2.29)

In the Fraunhofer region further approximations are made on the phase term
. We neglect in (2.28) the terms of order higher than 1 in the coordinates y 
e
and x on the aperture, obtaining:
jkr

j ejkR 


Jt (x , y  )ejk sin (x cos +y sin ) dx dy 
E(P )
=
2 R A

(2.30)

Consequently, in the Fraunhofer zone the eld distribution is similar (i.e. scaled
by a factor) on all spheres centered in the origin, where the aperture is placed: this
may therefore be considered as a point source.
If we write:
kx = k sin cos
(2.31)
ky = k sin sin

M. Orece: Antennas 2013, Aperture antennas

23

eq.(2.30) becomes:
E(P ) =


with
F (kx , ky ) =

j jkR
e
F (kx , ky )
2R


Jt (x , y  )ej(kx x +ky y ) dx dy 

(2.32)
(2.33)

Lets consider the plane z = 0. The sources (related to the eld) on the aperture
may be seen as a function Jt (x, y) dened on all the plane z = 0:
Jt (x, y) = Jt (x , y  )
Jt (x, y) = 0

inside A
outside A

(2.34)

F (kx , ky ) is clearly the double Fourier transform of Jt (x, y):


F (kx , ky ) = F[Jt (x, y)] =

 

Jt (x, y)ej(kx x+ky y) dx dy

(2.35)

and, vice versa:


Jt (x, y) = F

1 
[F (kx , ky )] = 2
F (kx , ky )ej(kx x+ky y) dkx dky
4

(2.36)

As it will be seen later,in the part of the Fresnel zone nearer to the sources, the
eld is essentially given by the geometrical optics contribution, with additional eld
oscillations due to diraction eects. In general, it is assumed as maximum limit of
this region the distance D2 /2 (i.e. 1/4 of the Fraunhofer distance) called Rayleigh
distance; the region within this limit is called Rayleigh zone. Here, the phase surfaces
radiated by a plane aperture illuminated in phase remain approximately planar and
parallel each other. A typical example is the paraboloid, whose phase surfaces on the
aperture are parallel planes, so that in the near region we can consider a propagation
by parallel rays.
The Raleigh zone can be considered as the near eld region for the aperture,
and therefore with dimension depending on the aperture itself, while the near eld
region for the sources depends only on the wavelength.

2.4

Gain ed eciency of an aperture

Eq.(1.21) requires the integration of the angular power density within the solid angle,
which in many case, in particular when dealing with apertures, is usually expressed
by complicated functions, often not analytically integrable.
For an aperture it exists a dierent formulation, usually simpler, to express the
power at the denominator of eq.(1.14) or (1.16) as the integral of the power density
on the aperture. Assuming that the eld on the aperture is TEM (which it is true
for apertures larger than one wavelength) the power density is S = |E|2 /Z0 , and the

M. Orece: Antennas 2013, Aperture antennas

24

directivity on the z axis (which is usually the direction of the maximum) is given

2


by:
 E(P  )dA


4
(2.37)
G(0) D(0) = 2  A

|E(P  )|2 dA
A

where P is a generic point on the aperture, and A is the aperture area.


If E(P  ) is constant on the aperture, the maximum gain is clearly
G(0) =

4
A
2

(2.38)

and the eective area is coincident with the geometric area. We may therefore dene
the aperture eciency as the ratio between the directivity of the aperture and that
of the same aperture uniformly illuminated: the eciency is given by

2


 E(P  )dA

1 A
= 

2.5

(2.39)

|E(P )| dA
2

Rectangular aperture

Lets consider a rectangular aperture on the (x , y  ) plane, with dimensions a and b
(see g. 2.3) and area A. Eq.(2.33) becomes:
F (, ) =

 a/2  b/2
a/2

b/2

Jt (x , y  )ejk sin (x

cos +y  sin )

dx dy 

(2.40)

We will now consider a few particular cases of practical interest.

2.5.1

Uniform amplitude and phase

For a uniformly illuminated aperture it is U (x , y  ) = 1 (normalized value); the


integral (2.40) can be easily computed and we nd:

F (, ) = A

sin

a
sin cos

a
sin cos

sin

b
sin sin

b
sin sin

(2.41)

The diagrams in the principal planes (x, z) and (y, z) are particularly interesting.
In the former ( = 0) eq.(2.41) becomes:
F () = A

sin

a
sin

a
sin

(2.42)

In the latter ( = /2) the eld is still given by eq.(2.42) with b instead of a: therefore in both principal planes the radiation patterns have, except for the horizontal

M. Orece: Antennas 2013, Aperture antennas

25

x^

z^

0 .5

0
0

-1 5

15

Figure 2.3: Normalized diagram of the eld intensity radiated from a rectangular
aperture uniformly illuminated (linear scale).
scale, the same behavior. The normalized radiation pattern is shown in g.(2.3) as
a function of the normalized variable


w=

k0 a
2 b

sin

(2.43)

The scale in the ordinates is linear in eld; in this case the minima are nulls, and
are located in the points w0 = n, with n = 1, 2, .... In the u, v plane, the nulls
are on a rectangular grid; moreover the highest secondary lobes are in the principal
planes, while in the other planes = const., e.g. the diagonal ones, their level is much
lower. A 3-D representation of the radiation pattern for a rectangular aperture, in
dB scale, is shown in g.(2.4), for the case of uniform illumination: we may note the
symmetry with respect to the principal planes, typical of the rectangular apertures
with separable variables.

2.5.2

Separable eld distributions

A common type of aperture illumination is that where the eld distribution is separable in the product of two functions:
Jt (x , y  ) = u1 (x )u2 (y  )

(2.44)

In this case the integral (2.40) is easy to be solved. Eq.(2.40) becomes:


F (, ) =

 a/2
a/2

u1 (x )ejkx

sin cos

dx

 b/2
b/2

u2 (y  )ejky

sin sin

dy 

(2.45)

The patterns in the principal planes (x, z) and (y, z) are therefore respectively:

M. Orece: Antennas 2013, Aperture antennas

26

10
0
10
20
30
40
2
2

1
1

0
0

1
2

Figure 2.4: Three-dimensional radiation pattern of a square aperture with uniform


illumination.


F (, 0) =

b/2
b/2

u2 (y )dy



a/2
a/2

u1 (x )ejkx

sin

dx
(2.46)



F (, /2) =

a/2

a/2

u1 (x )dx



b/2

b/2

u2 (y  )ejky

sin

dy 

The eects of the non uniform illumination and the errors in the phase patterns
on the principal planes can therefore be studied as one-dimensional problems, provided that the aperture eld be separable both in amplitude and phase in the form
given by eq. (2.44).
With these approximations, the expressions of the radiated eld can be reduced
to Fourier transforms of a single variable functions, many of which are found in
analytical form on the Fourier transforms tables. A few examples of such functions,
with the analysis of the main radiation characteristics, are shown in Table 2.1.
The behavior of the radiated eld from an aperture has, in general, some common
characteristics, which are connected to the tapering 3 of the aperture illumination
functions.
In fact, from Table 2.1we may observe that, for constant size of the aperture,
with the increase of the tapering the eciency decreases, the main beam widens
3
Tapering is the ratio (usually in dB but sometimes in number) between the eld intensity at
the center of the aperture and that at its edges. Usually this ratio is greater than 1, and therefore
in dB it is positive.

M. Orece: Antennas 2013, Aperture antennas

Type of
distribution
Uniform
Parabolic
1 (1 )x2
=1
= 0.8
= 0.5
=0
n x
cos 2 |x| < 1
n=0
n=1
n=2
n=3
n=4
Triangular
1 |x| |x| < 1

HPBW
(rad)
0.88/L

27

Position
First side
of rst null lobe level
(rad)
(dB)
/L
-13.2

Aperture
eciency
1

0.88/L
0.92/L
0.97/L
1.15/L

/L
1.06/L
1.14/L
1.43/L

-13.2
-15.8
-17.1
-20.6

1
0.994
0.97
0.833

0.88/L
1.2/L
1.45/L
1.66/L
1.93/L

/L
1.5/L
2/L
2.5/L
3/L

-13.2
-23
-32
-40
-48

1
0.81
0.667
0.575
0.515

1.28/L

2/L

-26.8

0.75

Table 2.1: Characteristics of the radiation patterns for various types of aperture
distributions.

(and consequently also the angle of the rst null), and the sidelobes decrease. We
may also note that also the derivatives of the eld at the edges of the aperture have
inuence on the radiation pattern, in the sense that a low value of such derivative
increases the eect of the tapering.
The choice of one aperture distribution or another thus will depend on the design
specications (for example, whether a high gain or low sidelobes are preferred). In
practice, however, the actual aperture distributions can be represented only roughly
by those reported in the Table, and it will be therefore often necessary perform numerical integrations or Fourier transforms, or to represent the aperture distribution
with a sum of functions which may be transformed analytically.
By introducing a suitable normalization, the previous equations, for example
eq.(2.45), can be rewritten in a more universal form.
Considering a separable distribution where the aperture eld is expressed by a
given function along y; for example, if it is uniform it is u2 (y  ) = 1. The radiation
pattern in the plane x = 0 ( = /2), except for a multiplicative constant, is, as
previously seen, of the type:
F () =

sin

b
sin

b
sin

(2.47)

M. Orece: Antennas 2013, Aperture antennas

28

while in the plane y = 0 ( = 0):


F () =

 a/2
a/2

u1 (x )ejkx

dx

(2.48)

a
sin

(2.49)

sin

If we now introduce new variables:


x=

2x
a

u=

then u1 (x ) f (x) and F () F (u); so that eq.(2.48) becomes:


F (u) =

a 1
f (x)ejux dx
2 1

(2.50)

It is clear that, if we have the same type of distribution on two apertures with
dierent size, these will produce the same radiation pattern as a function of the
variable u dened in the second of eqs.(2.49).

2.5.3

Eects of the phase error

We may have a phase error on a radiating aperture for various reasons: for example,
in the case of a parabolic reector antenna, when the primary feed is displaced with
respect to the focus; or because of the presence of a phase error in the primary feed;
or, for a horn antenna, it may be due to the sphericity of the eld wavefront incident
on the aperture.
Moreover, we will assume that the eld on the aperture is given by a separable
variables function, so that we will consider only one variable at a time. Assuming the
aperture distribution along x of the type f (x)ej(x) , where f (x) is the amplitude
(real) and (x) is the phase function, the expression for the eld radiated from the
aperture becomes proportional to:
F (u) =

a 1
f (x)ej[ux(x)] dx
2 1

(2.51)

We will limit the discussion to the following particular forms of (x):


linear error: (x) = x
quadratic error: (x) = x2
(a) Linear error - By inserting the expression (x) = x in eq.(2.51) we obtain:
F (u) =

a 1
f (x)ej(u)x dx
2 1

(2.52)

It is clear that eq.(2.52) has the same form as eq.(2.50) with u replaced by (u).
Consequently the distribution of the radiated eld is equal to that obtained with

M. Orece: Antennas 2013, Aperture antennas

29

constant phase distribution, but shifted along u by a quantity . The maximum


will be for u = , i.e. in the direction:
0 = sin1

(2.53)

The radiation pattern is therefore the same of an aperture with constant phase
distribution, but tilted by an angle 0 in opposite direction to the increasing sense
of the phase. Note that, since u is proportional to sin , in the variable there is
not only a rotation but also a deformation of the radiation pattern in abscissas.
(b) Quadratic error - In this case the radiated eld is proportional to:
a 1
2
f (x)ej(uxx ) dx
F (u) =
2 1

(2.54)

where, with this normalization, is the maximum phase error (obviously at the
edges of the aperture).
The computation of this integral is in general very dicult and usually it is done
numerically or with approximate methods.
The eect of a quadratic phase error is illustrated in g.(2.5) for two types of
). In both cases
illumination: uniform (f (x) = 1), and tapered (f (x) = cos2 x
2
= /2. The maximum is always for = 0 and, since the phase error is symmetric
with respect to the center of the aperture, also the radiation pattern will be symmetric with respect to the direction = 0. However we nd that, when becomes
suciently large, in some cases the main beam bifurcates, and we have the largest
maxima on both sides of the direction = 0.
1

0.9

0 .9

0.8

0 .8

0.7

0 .7

0.6

0 .6

0.5

0 .5

0.4

0 .4

0.3

0 .3

0.2

0 .2

0.1

0 .1

0
-10

-5

10

0
-10

-5

10

Figure 2.5: Eect of the quadratic phase error, for (a) uniform illumination; (b)
illumination of the type cos(x/2), for dierent values of the maximum phase error
(from the bottom: 0, /4, /2, ).
The general eect of the quadratic phase error is of reducing the maximum gain,
widening the main beam, increasing the sidelobes level and lling the nulls. For a

M. Orece: Antennas 2013, Aperture antennas

30

tapered illumination, since the sidelobes of the starting pattern are low, the lling
of the nulls makes often impossible to distinguish the rst sidelobe from the main
beam (see g.2.5, b). The gain loss due to this type of phase error is shown, in
percentage, in g.(2.6) for a uniformly illuminated aperture.
100

90

%G

80

70

60

50

40
0

0.05

0.1

0.15

0.2

0.25
beta/2pi

0.3

0.35

0.4

0.45

0.5

Figure 2.6: Gain loss due to the quadratic phase error, as a function of the maximum
error (in wavelengths): uniform illumination.

2.6

Circular aperture

The considerations and the results for a rectangular aperture, concerning to the
relationship existing between the eld on the aperture and the radiation pattern,
can also be applied to a circular aperture. However, in this case it is convenient to
use cylindrical coordinates (see g.2.7), which are related to y  and x by:
x = cos 

y  = sin 

(2.55)

Indicating with U (,  ) the eld distribution on the aperture, the expression for
the radiated eld (2.33) becomes:
F (, ) =

 2  a
0

U (,  )ejk sin cos( ) dd

(2.56)

where a = D/2 is the aperture radius. By introducing the normalized variables


r=

u=

2a
D
sin =
sin

(2.57)

M. Orece: Antennas 2013, Aperture antennas

31

2a

z
Figure 2.7: Circular aperture and coordinate systems.
U (,  ) and F (, ) are transformed into the functions f (r,  ) and F (u, ), respectively; moreover we will assume normalized to 1 the maximum of f (r,  ). The
radiated eld then becomes:
2

F (u, ) = a

 2  1
0

f (r,  )ejur cos( ) rdrd

(2.58)

Note that eq.(2.58) is simply the expression of the Fourier transform in cylindrical
coordinates, and it is called Fourier-Bessel transform of f (r,  ).

2.6.1

Uniform amplitude and phase

Assuming f (r,  ) = 1 in eq.(2.58) and carrying out the integration  , we obtain4 :


F (u) = 2a2

 1
0

J0 (ur)rdr

(2.59)

where J0 (ur) is the Bessel function of order 0. Then, by integrating with respect to
r we obtain:
J1 (u)
(2.60)
F (u) = 2a2
u
Note that F (0) = a2 , i.e. the aperture area.
The behavior of F (u) is easily predictable if we consider that J1 (u) has the
oscillating behavior shown in g.2.8.
Integrate from 0 to 2 in  the relationships (see, e.g., M.Abramowitz, I.Stegun: Handbook of
mathematical functions, New York: Dover Publ., 1965, pp.361)
4

cos(z sin  ) = J0 (z) + 2

cos(2k )J2k (z)

k=1

sin(z sin  ) = 2

sin[(2k + 1) ]J2k+1 (z)

k=1

recalling that only the integrals with k = 0 are non-zero.

M. Orece: Antennas 2013, Aperture antennas

32

1
0.8
0.6

0.4
0.2
0
-0.2
-0.4
-0.6
0

10

12

14

16

Figure 2.8: Behavior of the Bessel functions of order 0,1,2,3,4.


The normalized eld pattern F (u) is shown in g.(2.9) as a function of u. The
HPBW (in radians) is given by:
= 2 sin1 (0.51/D)
= 1.02/D

(2.61)

and the rst sidelobe is 17.5 dB below the main beam.


Such values are comparable with the corresponding values (
= 0.88/a for the
HPBW and -13.5 dB for the rst sidelobe) derived for the rectangular aperture. The
dierent levels of sidelobes should not however suggest that the circular aperture
has less power in the sidelobes than the rectangular aperture: in fact, in this latter
case, the rst SLL is constant in all the azimuth planes, while in the rectangular
aperture there are high sidelobes only in the principal planes (see g. 2.4).

2.6.2

Non uniform illumination

The eects of the tapering are similar to those seen for the rectangular aperture:
gain reduction, main beam widening and increase of the sidelobes. These eects can
be illustrated considering the series of aperture eld distributions
f (r) = (1 r2 )p

(2.62)

with p positive integer. The aperture distribution is therefore symmetric with respect to  . The normalized radiated eld is given by:
Fp (u) = 2a

 1
0

(1 r2 )p J0 (ur)rdr

(2.63)

2p+1 p!Jp+1 (u)


up+1

(2.64)

from where
Fp (u) = a2

M. Orece: Antennas 2013, Aperture antennas

33

10

G dB

20

30

40

50

60
0

10

15

Figure 2.9: Radiation pattern of an uniformly illuminated circular aperture.


The main characteristics of the radiated eld for aperture distributions of this
type with p = 0, 1, 2, 3, 4, are summarized in table (2.2), where is the aperture
eciency, is the HPBW and 0 is the angular position of the rst null, and shown
in g.(2.9) and (2.10).
In general, an arbitrary aperture distribution (provided that it is regular and
independent on  ) can be approximated with the sum
f (r)
= a0 + a1 (1 r2 ) + a2 (1 r2 )2 + a3 (1 r2 )3 + ... =

an (1 r2 )n

(2.65)

where each term give rise to an integral which may be solved analytically (eq. 2.64).
Moreover, by using the change if variable x = 1 r2 the coecients ai of the RHS of
eq.(2.65) are given directly by those of a polynomial expansion of f (x): the radiation
from the aperture will therefore be expressed by:
F (u) = a2

N

ap 2p p!Jp+1 (u)
0

up+1

(2.66)

Another interesting type of distribution is the following:


f (r,  ) = 1 r2 cos2 

(2.67)

M. Orece: Antennas 2013, Aperture antennas

p
0
1
2
3
4

1.00
0.75
0.56
0.44
0.36

1.02/D
1.27/D
1.47/D
1.65/D
1.81/D

0
1.22/D
1.63/D
2.03/D
2.42/D
2.79/D

34

SLL
17.6
24.6
30.6
36
40.5

Table 2.2: Characteristics of the radiation from circular apertures with


illumination of the type (1 r2 )p .
0

10

G dB

20

30

40

50

60
0

10

15

Figure 2.10: Radiation patterns from circular apertures with illumination of the
type (1 r2 )p .

M. Orece: Antennas 2013, Aperture antennas

35

which is uniform in the plane  = /2 and has some tapering (cosine square) in
 = 0. By replacing eq.(2.67) in eq.(2.58) we get:


F (u, ) = 2a


1  2  1 2
J1 (u)
2  jur cos( )


r cos e
r dr d
u
2 0 0

(2.68)

The integral in eq.(2.68) is solved by using the expansion:




ejur cos( ) =

Jn (ur)ejn( )

(2.69)

n=

and using some integral expression of the Bessel functions that allow to reduce it to
combinations of eqs.(2.64).
The radiation patterns in the planes = 0 and = /2 are shown g. 2.11.
Note that the HPBW is larger in the = /2 plane than in = 0 , clearly because
the aperture distribution is tapered in that plane, while it is uniform in the other.
However, the sidelobe level is not the same that it would be with the same tapering
in both planes: in other words, in a circular aperture the tapering in a principal
plane inuences the sidelobe level also in the other plane, unlike the rectangular
aperture.
0

10

G(u)

15

20

25

30

35

40
0

10

15

Figure 2.11: Radiation pattern of a circular aperture with distribution of the type
f (r,  ) = 1 r2 cos2  , in the planes = 0 (-) and = /2 (- - -).
Exercise: Show that the aperture eciency for a distribution with tapering t of the type
t + (1 t)(1 r2 ) (called parabola-on-a-pedestal), is given by:

M. Orece: Antennas 2013, Aperture antennas

36

3 (1 + t)2
4 1 + t + t2

and that for a distribution of the type t + (1 t)(1 r2 )2 (called square-parabola-on-a-pedestal),


is given by:
5 (1 + 2t)2
=
3 3 + 4t + 8t2

2.7

Phase center

The phase center of an aperture and, in general, of an antenna, is the center of


curvature of the radiated phase surface and therefore, in the assumption of far eld
conditions, is the center of the spherical wave radiated by the antenna, i.e. of
the constant phase (spherical) surfaces. From this denition, this point should be
independent on the angle and the distance of observation.
But in practice this is not true. In fact, computing the eld radiated by an
aperture with the various formulas previously shown (e.g. with the Fourier transform
of the aperture eld), we nd that on a spherical observation surface, any point be
taken as center, the phase is not constant, except that for some particular case, for
example when the eld distribution on the aperture plane is symmetric and with
constant phase. And this is valid only for the main beam: in fact, the Fourier
transform of real symmetric function is also real and symmetric, but usually with a
change of sign for any successive lobe: consequently, on a sphere at large distance
a previously dened, the phase will switch between 0 and 180 from one lobe to
another, passing through an amplitude null.
In general, the phase of the radiated eld on a sphere al large distance will always
vary with the observation angle, and it doesnt exist a point which may be taken
as the center of a sphere where the phase of the radiated eld is exactly constant
for all angles. An example of the behavior of the phase is shown in g.(2.12).
In conclusion, the constant phase surface, even at large distance, is not exactly a
sphere, and therefore the phase center varies its position with the observation angle.
However, it is possible to nd a xed point for which the phase error is small in a
reduced angle (for example, that corresponding to the main beam of the antenna)
which can be considered, with these limitations, the phase center.
The determination of the position of the phase center is particularly important
for the sources used as reector antenna feeds: as an example, in a paraboloid the
wave radiated from the feed must be spherical with center in the paraboloid focus,
and any error with respect to this requirement produces a degradation of the antenna
pattern. In this case, we must aim to minimize the phase error on the paraboloid
aperture by nding a spherical surface where the phase variation of the eld radiated
from the feed is minimum5 , and making coincide its center with the focus of the
paraboloid.
5

The minimization criterion can be variable, as it will be seen in the following.

M. Orece: Antennas 2013, Aperture antennas

0.5

0.5

-1

0
-10

10

-2.5
-10

37

10

Figure 2.12: Example of a radiation pattern (in amplitude, left, and phase, right)
from an aperture with phase error.
We must also note that the phase pattern of the eld radiated from an aperture is
usually dierent in the principal planes, so that a certain point may be the optimum
phase center for one plane but not for the other: this is the case of apertures with
unequal eld distribution in the principal planes (for example, the rectangular and
circular horns, with fundamental mode only).
a/2
r
0

-a / 2
d

Figure 2.13: Geometry for the determination of the phase center.


The approximate position of the phase center can be found observing that, by
moving the origin of the reference system from O to O with a displacement d
(positive or negative) along the z axis (see g.2.13), the vector r describing the
space of the sources becomes
= x
x + y y d
z

(2.70)

so that the radiated eld E on the sphere with center in O is related to the eld E0
on the sphere with center in O from the relationship
E = E0 ejkd cos

(2.71)

Thus the displacement of the origin produces only a phase variation.


If we consider, in order to nd the phase center, the relative phase shift ()
with respect to the direction = 0 (the z axis), i.e. xing (0) = 0, we will obtain
() = 0 () kd(1 cos )

(2.72)

M. Orece: Antennas 2013, Aperture antennas

38

where the suxes have the same meaning as in the previous equation.
With a suitable choice of d we can therefore obtain that for a given angle of
observation 0 it is (0 ) = (0), i.e. that the constant phase surface surface is
tangent to the sphere and moreover it intersects it in 0 . If then 0 0, we will
obtain the osculating circle to the curve at constant phase in its intersection with
the z axis: on this circle, the phase variation will be innitesimal of higher order,
and its center will be coincident with the phase center on the z axis, and in this
region it will be ()
= (0).
By expanding, in eq. (2.72), the cosine in Taylor series and expressing both
phase and observation angles in degrees, we obtain that the distance of the phase
center from the reference point (respect to which the phase was computed) is
given by:
(0 (0 ) 0 (0))
(2.73)
d/ = 18.24
02
The sign indicates that, if the computed (or measured) phase variation vs. , on a
sphere with a given center, is positive, the phase center is on the opposite direction
with respect to the radiation direction.
The previous considerations have been carried out for the z axis. If we compute
(or measure) the phase with respect to the point determined in this way, that will
not be exactly constant: in fact, it has been attened in the neighborhood of the
axis but not for wide angles. For using an antenna as a reector feed, covering wider
angles, it may be more convenient to choose as phase center (i.e. where the focus of
the paraboloid must be) a point minimizing the phase error according some dierent
(and in some way arbitrary) criteria; some examples of these are shown in g.(2.14):
as an example, the same phase on the axis and on the maximum illumination angle
(A), or on the axis and for a given angle or a given tapering level, or minimization
of the maximum absolute value of the error (B), weighting the dierent illumination
in the center and at the edges, etc.

A
B

Figure 2.14: Behavior of the phase of the radiated eld from an aperture, for dierent
positions of the phase reference point.

M. Orece: Antennas 2013, Aperture antennas

2.8

39

Apertures with arbitrary eld distribution

The eld radiated by an aperture can be expressed analytically only in particular


cases, when the shape of the aperture and the eld distribution function are very
simple. More generally, it is necessary to resort to numerical integrations which, in
many cases, can be simple two-dimensional FFTs.
Alternatively, if the aperture shape is relatively simple, the eld can be expressed as a sum of analytically integrable functions: an example (dened in an
unidimensional domain) is the expansion of eq.(2.65). Techniques of this type were
particularly useful when fast computers were not available, but they are anyway
valid when optimizations requiring repeated integrations are to be carried out, to
reduce the total computational load.
One of the rst examples of expansion in two-dimensional domain is that of an
elliptical aperture6 , which can be suciently easily analyzed by using an illumination
function expressed in Fourier series, and then using the Fourier-Bessel transforms of
sinusoidal functions.
The radiation patterns computed through this assumption are in good agreement
with those determined experimentally, especially for what concerns the gain, the
HPBW, the rst sidelobe and the position of the other sidelobes. Also in this case
the sidelobes level in one of the principal planes depends, although less signicantly,
on the illumination in the other principal plane. Fig.(2.15) shows the sidelobe level
18
20

Liv. lobo secondario (dB)

22
24
26
28
30
32
34
36
38
2

8
10
12
Tapering piano H (dB)

14

16

18

Figure 2.15: Sidelobe level in the H (-) and E (- - -) planes, as a functions of the
edge illumination in the H plane (from Adams and Kelleher).
computed for apertures obtained with reector antennas, by using measured feed
6

See R.J. Adams e K.S. Kelleher: Pattern calculation for antennas of elliptical aperture,
Proc. IRE, vol.38, Sept. 1950. The paper is in fact a short summary, where it is also reported the
diagram of g.(2.15), which was later reported on Jasiks Antenna engineering handbook and often
used for the preliminary design of reector antennas.

M. Orece: Antennas 2013, Aperture antennas

40

radiation patterns, keeping constant (-13 dB) the taper in the E plane and varying
it in the H plane: we may observe that in the plane where the edge taper varies,
the sidelobe level decreases for increasing taper, while in the other there is a slight
increase, still however remaining almost constant within a couple of decibels.

M. Orece: Antennas 2013, Aperture antennas

2.9
2.9.1

41

Horn antennas
Sectoral and pyramidal horns

The sectoral horns are obtained from a rectangular waveguide having dimensions
allowing the propagation of the fundamental mode TE10 alone, by widening to a
dihedron two of the four walls, as shown in g.(2.16): the cross section widens
gradually in one direction, while it remains constant in the perpendicular one. It is
therefore possible to have sectoral horns with directive pattern only in the H or E
plane.
lh

le
a

E
a

Figure 2.16: Sectoral horns in the H (left) or E (right) planes.


The higher order modes arising in correspondence of the discontinuity are below
cut and therefore are exponentially attenuated, and in the horn cross sections that
would allow their propagation their amplitude is almost negligible: therefore there
is always only the fundamental mode TE10 . On the aperture the amplitude of
the transverse electric and magnetic eld will vary following the fundamental mode
characteristics, with half-sinusoid in one of the principal directions and constant in
the other.
Between the two dihedral walls the wave is approximately cylindrical, with axis
in the intersection of the two planes. Consequently the eld phase behavior is
approximately quadratic, with a maximum deviation (assuming a are angle of
the horn not too large) equal (in radians) to
a2
(2.74)
4
where a is the aperture width and is the length of the oblique side of the horn,
still measured from the vertex of the dihedron.
It can be easily derived that the exact expression of the phase error as a function
of the dimension and the aperture angle is

a
tan
(2.75)
=

2
The radiation patterns are shown in g.(2.17) for the two principal planes. The
corresponding phase patterns are also shown in g.(2.18): both are plotted vs. the
=

M. Orece: Antennas 2013, Aperture antennas

42

dimensionless quantities (b/) sin and (a/) sin , having as parameter the maximum phase error in wavelength, s or t in both planes, given respectively by:
s=

b2
8 E

t=

a2
8 H

(2.76)

It can be seen that, when the maximum deviation is high, the main beam in the
E plane (and also, although less noticeably, in the H plane) splits in two. In general,
the sidelobes are higher in the E plane than in the H plane.
The patterns of g.(2.17) are suciently precise for horns of a few wavelengths;
for smaller apertures, up to about 2, the values obtained shall be multiplied by an
obliquity factor which depends on the reection coecient and on the wavelength on
the horn aperture, With sucient approximation this factor is given7 by cos(/2).
The gain of a sectoral horn can be computed from the curves of g.(2.19), provided that the smallest dimension of the aperture is at least one wavelength. Measurements carried out on several horns show that the dierences between theoretical
and experimental values are within a few tenths of dB.
The pyramidal horn - of which a few samples, operating in various frequency
bands, are shown in g.(2.20) - is obtained by widening in pyramidal shape all the
4 planes which bound the rectangular waveguide. In general, the two couples of
planes do not intersect each other at the same distance from the aperture, so that
in fact it is not a perfect pyramid.
The pyramidal horn, thanks to its good reproducibility and easy manufacturing,
and to the excellent agreement between theoretical and experimental results, is
frequently used as a Standard Gain Horn (SGH) in antenna gain measurements as
a reference standard. Its gain can be computed precisely through the relationship
(in decibel):
ab
(2.77)
G = 10(log 0.81 4 2 ) Le Lh

The rst term of eq.(2.77) is the gain of a rectangular aperture illuminated with
a TE10 mode and constant phase, and it can be computed analytically through
eq.(2.37). The other two terms represent the gain loss due to the phase error,
respectively in the E and H plane, and can be derived from the curves of g.(2.21).
For not too high phase errors (e.g. 0.8 in the E plane and 0.6 in the H plane),
this gain reduction can be approximated by (dashed curves g.(2.21))
Le = 0.405 2
Lh = 0.177 2

(2.78)

The gain can also be derived from the product of the two values that can be read
in the ordinates in g.(2.19), with a further multiplication by /32.
7

See R.C.Johnson, H.Jasik: Antenna engineering handbook - 2nd ed., New York: Mc. Graw-Hill,
1984, p.15-5.

M. Orece: Antennas 2013, Aperture antennas

43

0
s=1

3/4
10
1/8

E/Emax

15

1/4

3/8

1/2

20

25
0
30

35

40
0

0.5

1.5
u

2.5

2.5

0
t=1

3/4
10
1/4

E/Emax

15

3/8

1/2

1/8
20

25
0
30

35

40
0

0.5

1.5
u

Figure 2.17: Radiation patterns of sectoral horns with phase error, in both principal
planes. In abscissas u = (b/) sin and u = (a/) sin , respectively.

M. Orece: Antennas 2013, Aperture antennas

44

150

1/8
1/4
3/8
1/2

100

3/4
P has e (E pl.)

50
t=1
0

-50

-100

-150
0

0.5

1.5
b/ s in

2.5

2.5

150

1/8
1/4
3/8
1/2
3/4
t=1

100

P has e (H pl.)

50

-50

-100

-150
0

0.5

1.5
a/ s in

Figure 2.18: Phase patterns of sectoral horns with phase error, in both principal
planes. In abscissas u = (b/) sin and u = (a/) sin , respectively.

M. Orece: Antennas 2013, Aperture antennas

45

150

l=100
100

75
50

30
50
20
15
10
6

10

15
b/lambda

20

25

20

25

150
l=100

75
100

50

30
20
50
15
10
8

10

15
a/lambda

Figure 2.19: Gain of sectoral horns in the E and H planes (respectively for a/ = 1
and b/ = 1).

M. Orece: Antennas 2013, Aperture antennas

46

Figure 2.20: Examples of pyramidal horns.


From the behavior on g.(2.17) we may observe that the radiation patterns of a
square horn are not symmetric about the axis, but they dier in the E and H planes.
However, with a suitable choice of the b/a ratio, it is possible to obtain a relatively
symmetric beam, e.g. by imposing that the HPBWs be equal (or the angles at
10 dB, etc.). If the criterion chosen is that of the HPBWs, from g.(2.17) we derive
that they are respectively (in radians) 0.88/b for the E plane (uniform distribution,
constant phase) and 1.2/a for the H plane (cosine distribution, constant phase).
For the symmetry of the beam at this level, it must be a/b = 1.2/0.88 = 1.36. As a
practical rule, usually the ratio 4/3 is used.
In general, both to reduce the phase error on the aperture and to avoid the persistence of the higher order modes generated by the discontinuity, it is recommended
that the are angle be small, and consequently the horn should be long; problems
may therefore arise in order to the excessive axial length.
To avoid an excessive length, it is worth noting that, for constant , with the
increase of the aperture angle also the aperture area becomes larger (and so the
gain) but there is also a larger phase error (which reduces the gain): so the eects
on the gain are contrasting. The maximum gain is obtained when the maximum
phase error on the aperture is respectively of 3/8 for the H plane and of /4 for

M. Orece: Antennas 2013, Aperture antennas

47

12

10

Le

0
0

0.1

0.2

0.3

0.4

0.5
s

0.6

0.7

0.8

0.9

0.5
t

0.6

0.7

0.8

0.9

Lh

0
0

0.1

0.2

0.3

0.4

Figure 2.21: Gain loss due to the phase error in the E and H planes.
the E plane. This condition is called optimum horn. This corresponds to:

a = 3 h
b = 2 e

(2.79)

where a and b are the dimensions of the aperture of the optimum horn in the H and
E planes, respectively. The gain of an optimum horn, in decibels, is:
G = 10(0.808 + log

ab
)
2

(2.80)

with a loss of 2 dB with respect to the case without phase error.


The optimum horn, since it maximizes the gain for a given horn length, corresponds to the locus of the maxima on the curves of g.(2.19). However, with

M. Orece: Antennas 2013, Aperture antennas

48

constant geometry, the gain is not stationary with respect to the frequency, since in
this case the ratio / changes, and with it the parameter of the curve in g.(2.19).
Example:
Design a rectangular horn in the Ku band (12-18 GHz) with a gain of 20 dB at 15
GHz, with symmetric main beam.
Since the gain is high, we should apply the criterion of the optimum horn, and we get
from (2.80) that the area must be 15.562 ; since a/b = 4/3, it results: a = 4.555(9.11 cm);
b = 3.416(6.83 cm). The lengths in the two planes, from (2.79), result to be respectively
6.92 (13.84 cm) in the H plane and 5.83 (11.66 cm)in the H plane.
Since the standard waveguide in that band is the WR62 (15.8 7.9 mm), with a
simple geometrical construction we may derive that the axial length of the longitudinal
horn section should be respectively of 10.80 cm in the H plane and 9.86 cm in the E plane
(half are angles of 19.22 and 17.03 ). A waveguide/horn connection at dierent lengths
is complicated, so that it is preferred to assume the same value of distance aperture waveguide junction for both planes: as an example, the average value of 10.33 cm. Keeping
the same size of the horn aperture, the new values of become respectively 13.30 cm and
12.17 cm, with phase errors of 0.240 and 0.390: the gain remains the same within 0.1 dB.
At lower and upper frequencies, the gain is respectively 18.6 dB and 20.7 dB.

2.9.2

Smooth circular horns

Smooth circular horns are, together with the rectangular horns, among the oldest
types of horn antennas. The earliest works published on this subject are of the
thirties of the 20th century.8 They are derived from the circular waveguide, and
are obtained by connecting to it a truncated cone. For this reason the antenna is
also called conical horn. The prole, instead of being rectilinear, can be shaped
(to reduce the phase error on the aperture) or tapered (to reduce the reection
coecient in the junction waveguide - cone). To this family belongs also the conical
horn with zero are angle, i.e. the truncated cylindrical waveguide.
The eld topography on the aperture of a conical horn can be considered as
derived from that of a circular waveguide, whose fundamental mode is the TE11 .
The modes in a conical waveguide are dierent from those in a cylindrical waveguide
(associated Legendre and spherical Bessel functions instead of Bessel and exponential
functions), but for small angles it can be shown their substantial coincidence.
Actually, as for the rectangular horns, this eld topography is not exactly as
that of the waveguide, because the truncation of the cone also generates higher
order modes. In particular, there may be the generation of currents, and therefore
radiation, on the outer surface of the horn and of the guide, especially for small
aperture diameters.
8

Se e.g. G.C.Southworth and A.P.King: Metal horns as directive receivers of ultra-short


waves, Proc.IRE, Vol.27, pp.95-102, Feb.1939. This paper, together with many others concerning
the horn antennas, was collected by A.W.Love in Electromagnetic horn antennas, IEEE Press,
1976.

M. Orece: Antennas 2013, Aperture antennas

49

BOCCA DELLA TROMBA

dm
L1
L

Figure 2.22: Geometry of a conical horn.


Assuming a eld topography on the aperture given by the circular TE11 , with no
phase error, and integrating to obtain the radiated eld9 we obtain an illumination
eciency of about 0.83 (0.8 dB). In presence of phase error, the gain decreases,
and it is given by the formula:
G = 20 log

C
L

(2.81)

where C is the circumference of the circular aperture; L is the term taking into
account the gain loss for non uniform illumination and for phase error, and it is
shown in g.(2.23). The curve may be approximated with the quadratic expression
L
= 17 s2 , where s is the phase error measured in wavelength.
5.5
5
4.5
4

L, dB

3.5
3
2.5
2
1.5
1
0.5
0

0.1

0.2

0.3
s

0.4

0.5

0.6

Figure 2.23: Correction factor for the gain of a conical horn (solid curve) and its
approximate expression (dashed line).
Also in this case, keeping constant , by increasing the are angle the aperture
area increases as well as the phase error, which have opposite eects on the gain.
9

Details on the computation ar reported in S.Silver,op.cit., pp.336-341.

M. Orece: Antennas 2013, Aperture antennas

50

The optimum value (optimum horn condition) is for s = 3/8 = 0.375, corresponding,
with the small are angles approximation, to

d = 3
(2.82)
For wide are angles, eq.(2.82) is no longer valid, and it must be expressed exactly
using trigonometric functions. The gain of an optimum conical horn, in dB, is given
by:
C
(2.83)
Gopt = 20 log 2.82

Consequently, and similarly to the rectangular case, the optimum horn has a
gain 2 dB less than a horn of the same size and constant phase on the aperture.
In g.(2.24) the diameter and the corresponding axial length are reported for an
optimum horn for a required gain, while in g.(2.25) is shown the gain of a conical
horn vs. diameter, with the axial length (in wavelengths) as a parameter.
The behavior of the radiation pattern and of the sidelobes can be derived with
the usual integration on the aperture plane: a few results are shown in g.(2.26)
with varying phase error on the aperture edge. It can be seen how the deformation
of the pattern for strong phase errors is very similar to the case of a rectangular
horn. In fact, the fundamental modes in rectangular and circular waveguides have
some similarities (tapered in the H plane, almost constant in the E plane).

2.9.3

Beam symmetry and cross polarization

By carrying out the integral on a circular aperture, with azimuthal variation with
m = 1, we obtain a eld expression of the type:
V0 jkr

e
(FE () sin + FH () cos )
(2.84)
r
where the functions FE,H contain Bessel and trigonometric functions. However that
expression is approximately valid, at least for the main beam, also for non circular
apertures, as the rectangular horn, and it has therefore a suciently general validity:
in fact, eq.(2.84) can be seen as the st term of a Fourier series in the azimuthal
variable .
By projecting eq.(2.84) on the co- and cross- polarized unit vectors p and p given
in eq.(1.9), we may observe that:
E=

in the principal planes the cross-polarization vanishes;


the maximum of the cross polarization, for any , is in the 45 planes, where
it is given by:
Ep 0.5(FE + FH )
Eq 0.5(FE FH )

(co polar component)


((cross polar component)

(2.85)

Consequently, the more symmetric the radiation pattern is in the principal


planes, the lower is the cross polarization.

M. Orece: Antennas 2013, Aperture antennas

51

10

d/lambda, L/lambda

10

10

10

10

12

14

16

18
20
G_opt

22

24

26

28

Figure 2.24: Diameter and axial length of an optimum conical horn.

L=100
30
75
50
30
25

20

G (dB)

15
10
8

20
6
4
15

10 0
10

10
d/lambda

Figure 2.25: Gain of a conical horn vs. diameter.

M. Orece: Antennas 2013, Aperture antennas

52

Ap. efficiency (dB)

10
s=1
3/4

15

1/2
1/8

20

3/8
1/16

25

1/4

30

35
0

4
u

Ap. efficiency (dB)

10

15

s=1

1/4

20

3/4

3/8

1/2
25

1/8

30

1/16

35
0

4
u

Figure 2.26: Normalized radiation patterns of conical horns, for varying phase error
on the aperture, in the E (above) and H (below) planes.

M. Orece: Antennas 2013, Aperture antennas

53

Due to the non excellent symmetry of the horns seen up to now, their cross
polarization characteristics are poor out of the axis. The rectangular horn with
a/b = 4/3 may have some symmetry of the principal planes, but the phase patterns
are dierent, so that the cross polarization is not small. Similarly for the circular
horn, which has dierent HPBWs in the principal planes, except for the case d/
=
1.1, where the cross polarization has a minimum (lower than -30 dB). At such
optimum condition it corresponds however a radiation pattern with xed and wide
beam (the total angle at -10 dB is about 120 ), which can be used only in particular
cases.
a

a
y
b

Figure 2.27: Scheme of the octagonal horn.

10

15
a/b=1

Max X Pol (dB)

20
0.6
25
0.4
30
0
35

40
0.6

0.8

1.2

1.4
d/lambda

1.6

1.8

2.2

Figure 2.28: Behavior of the cross polarization of an octagonal horn for varying a/b
and d/.

M. Orece: Antennas 2013, Aperture antennas

54

In the literature, we may nd other types of horn antennas, designed to control


both the beamwidth and the cross polarization, which may be signicantly reduced,
at least for particular congurations. A few examples are:
The octagonal horn10 (see g.2.27): the behavior of the cross polarization is
shown in g.(2.28) for various dimensions and shapes;
the diagonal horn11 , whose eld topography, shown in g.(2.29a), is obtained
by exciting two in-phase linear polarizations in a square horn: the radiation
pattern is reported in g.(2.29b).
All the previously mentioned horns owe their low cross polarization to the fact
that, in particular conditions, the relationship between direct and cross polarization
of the eld components on the aperture is similar to the optimum for minimum
radiated cross polarization, which, as previously mentioned (Sect. 2.2.1), requires a
non-vanishing cross polarized component on the aperture, if the dimensions of this
latter are small (up to about 2 diameter).

2.9.4

Bimodal horn

For larger and more directive horns, to reduce the cross polarization we need to
minimize the cross polarization on the aperture. A horn which was invented to
satisfy this requirement is the bimodal horn, also called Potter feed12 . This horn
blends in a suitable manner the TE11 and TM11 modes, in order to obtain almost
parallel eld lines (see g.2.30) except obviously in the region of the boundary (where
the electric eld must be perpendicular to it). The higher order mode TM11 is
generated by a radial discontinuity, as shown in g.(2.31).
The expressions of the radiated eld, for the TE and TM modes, assuming
a negligible phase error (i.e. approximating to a cylindrical waveguide), are the
following:
e
EH =
2

jkR

e
EH = ka
2

EE =

R
jkR

1+

jkR
kaK211E e R

11H
k

11H cos
k

sin )
J1 (K11H a) J1 (ka
sin
sin

+ cos J1 (K11H a)

11E
k

+ cos

J1 (ka sin )

k sin
K11H

2 cos

J1 (K11E a) J1 (ka sin )



2
sin
K11E
1 k sin

(2.86)
sin

EE = 0
where:
10

See A.W. Rudge, IEE Conf. Publ. 169. Antennas and Propagat., Nov. 1978, pp.360-363.
See A.W.Love, op.cit., p.189.
12
P.D.Potter: A new horn antenna with suppressed sidelobes and equal beamwidths, Microwave J., vol.6, June 1963, pp.71-76.
11

M. Orece: Antennas 2013, Aperture antennas

55

1.5

.2
.5
.7

.2
.5
.7

-10
dB
-20

1.5

-30
-40
-2

-1

Figure 2.29: Field topography (a) and radiation pattern (b) of the diagona horn.

TE 11

TM 11

Figure 2.30: Field lines of the TE11 and TM11 .


MATCHING IRIS
HORN FLARE
MODE SUPPRESSOR
HORN
APERTURE

PHASING
SECTION
TAPER
CIRCULAR
GUIDE

2a0

2a

2 a1
l

TE o11
TE o11

TE o11
TE o11 TM o11
TE o11 TM o11

Figure 2.31: Bimodal horn and excitation scheme of the TM11 .

M. Orece: Antennas 2013, Aperture antennas

56

K11E a = 3.832 is the rst root of J1 ;


K11H a = 1.841 is the rst root of J1 ;
=

k02 K 2 in both cases.

The radiation pattern of the TM11 has a null on the axis, and the rst sidelobes,
not so small, are out of phase to those of the TE11 .
In g.(2.32) are shown the amplitude and phase patterns (referred to the center
of the aperture) of a horn13 with only the mode TE11 , from where it can be seen a
signicant dierence of the radiation in the principal planes.
Introducing various values of blending ratio of the two modes, we obtain radiation
patterns with a better symmetry in the principal planes (see g.(2.33)). Note that,
for dierent values of the parameter, the radiation pattern in the H plane is relatively
stable, while in the E plane the beam widens and the sidelobe decrease. Clearly,
there will be an optimum condition when the E plane beamwidth will reach that in
the H plane. With reference to the powers, it can be found that the optimum ratio
is when the power associated to the TM11 is about 15% of that of TE11 .
In g.(2.34) the amplitude and phase patterns of this latter case are plotted.
Note that also the phase pattern is about the same in both planes, so that the phase
centers in the E and H plane will be coincident.

2.9.5

Corrugated horns

The corrugated horn is one of the most sophisticated primary radiators for reector
antennas, and it has excellent characteristics of beam symmetry (and cross polarization), phase center stability and beam eciency (dened as the percentage of the
power radiated in the main beam). In its most common version, it consists of a
conical horn whose walls, instead of being smooth, are corrugated with azimuthal
grooves, usually periodic (see g. 2.35). The groove, usually with rectangular section
(although there are many examples with dierent proles), has therefore a width g
and a depth h; the wall between grooves (called tooth) has a thickness t.
Lets examine rst a simplied model of the corrugated wall in order to derive
some easily manageable boundary conditions. For sake of simplicity, we will consider
a corrugated metallic wall where the envelope of the corrugations is a plane (see
g.(2.36)). Moreover, we will assume that g
, and t
g. The unit vectors
r and are respectively perpendicular and parallel to the grooves on the plane,
and we will nd the boundary conditions for the tangent components of E on the
separation plane between the external medium (typically, air) and the corrugations.
The E component will be certainly zero in correspondence of the (metallic) teeth;
but it will be zero also in correspondence of the groove, because, being g
, its
13

It is the same conguration studied by Potter in its original paper, but taking into account
the phase error.

M. Orece: Antennas 2013, Aperture antennas

57

10

Rel.gain (dB)

20

30

40

50

60

70
0

10

20

30

40
theta

50

60

70

80

30

40
theta

50

60

70

80

150

100

Phase (deg.)

50

50

100

150
0

10

20

Figure 2.32: Amplitude and phase patterns for the TE11 mode only, with phase
error.

M. Orece: Antennas 2013, Aperture antennas

58

Figure 2.33: Radiation patterns in the principal planes of a bimodal horn, with
dierent blending rations of the TE11 and TM11 modes, with phase error.

M. Orece: Antennas 2013, Aperture antennas

59

10

20

Rel.gain (dB)

30

40

50

60

70
0

10

20

30
theta

40

50

60

30
theta

40

50

60

150

100
E

Phase (deg.)

50

50

100

150
0

10

20

Figure 2.34: Amplitude and phase patterns of a bimodal horn in the optimum
conguration.

M. Orece: Antennas 2013, Aperture antennas

60

R
h
h

Figure 2.35: Scheme of a corrugated horn.

^r

z=0

Figure 2.36: Corrugated wall.


walls will behave as a parallel plate waveguide below cut for the components parallel
to them. Consequently on all the separation surface it will be
E = 0

(2.87)

For the component Er perpendicular to the wall, the groove behaves like a parallel
plate waveguide of innite width, so that the eld propagates with k = k0 (where
k0 is the propagation constant of the medium). Since the groove has depth h, it will
be equivalent to a short circuited transmission line of length h, with characteristic
impedance Z0 . Consequently, on the separation surface, the ratio between electric
and magnetic eld, also considering that, on the teeth, Er is anyway zero) will be
in average
g
Er
Z0 tan kh
=j
(2.88)
H
g+t
Equations (2.87) and (2.88) can therefore be considered as macroscopic boundary
conditions for a corrugated planar surface. They can also be applied to the concave
surface of the horn, provided that the radius of curvature is suciently large. On
the contrary, for a typical corrugated waveguide, where the diameter is less than ,
and also for the rst part (the throat) of the corrugated horn, we must take into
account that the groove can be seen as a short circuited segment of radial waveguide,
so that the value of the characteristic impedance and the propagation constant of
the equivalent line in eq.(2.88) are modied.

M. Orece: Antennas 2013, Aperture antennas

61

More precisely, in this radial waveguide, which has a zero azimuthal electric
eld component (E ), the fundamental TEM mode cannot be considered, because
the typical excitation of a circular waveguide has an azimuthal index m = 1, or
higher, but never equal to 0. The lowest order mode will therefore be the TM01
whose characteristics can be found in the literature 14 . This however doesnt change
substantially the considerations previously developed.
Lets apply now these boundary conditions to a corrugated conical waveguide to
determine the possible propagation modes.
Notice that, in a conical waveguide, the forward TE and TM modes, with polarization mainly parallel to y, are given, for the modes with m = 1 (typical of the
structures with rotational symmetry), by the following expressions:

TE modes :

E (r, , ) = A
E (r, , ) = A

P1
(cos ) (2)
H
hH (kr) sin
sin

dP1
(cos ) (2)
H
hH (kr) cos
d

Er (r, , ) = 0

H (r, , ) = YT E

H (r, , ) = YT E

H (H + 1) (2)

hH (kr) cos
Hr (r, , ) = jAY0 P1
(cos )

kr




H (H +1)

1/YT = ZT = Z0 1

TM modes :

E (r, , ) = B

(2.89)

k2 r2

(cos ) (2)
dP1
E
hE (kr) sin
d

E (r, , ) = B

(cos ) (2)
P1
E
hE (kr) cos
sin

E (E + 1) (2)

Er (r, , ) = jBZ0 /ZT P1


hE (kr) sin
(cos )

kr

H (r, , ) = YT E ; H (r, , ) = YT E ; Hr (r, , ) = 0




1/Y = Z = Z 1 E (E +1)
T

k2 r2

(2.90)
14

N.Marcuvitz, Waveguide handbook, New York: Mc.Graw-Hill, 1951, pp. 90.

M. Orece: Antennas 2013, Aperture antennas

62

where P1 (cos ) is an associate Legendre function of the rst kind 15 , of order -1


and degree , and h(2)
(kr) is a spherical Hankel function (or also spherical Bessel
function of the third kind 16 ), dened as


(2)
H
(x)
2x +1/2

(2.91)

H(2) (x) = J (x) jY (x)

(2.92)

h(2)
(x)

with
The asymptotic expression of h(2)
(x) represents a spherical wave, and it is given
by

(+1)

ej[x 2 ]

(2.93)
=
x
It can be easily seen that modes purely TE or TM are not possible, because the
application of the boundary conditions (2.87) and (2.88) leads to all elds equal to
zero on the boundary and consequently in all the volume.
On the contrary, there is the possibility of hybrid modes, i.e. modes consisting
of couples of modes TE and TM, combined with a ratio = B/A, with the same
longitudinal propagation constant. Their expression will therefore be:
h(2)
(x)

dP 1 () (2)
P 1 ()
+
h (kr) sin
E (r, , ) = A
sin
d
P 1 () dP1 () (2)
+
h (kr) cos
E (r, , ) = A
sin
d
Er (r, , ) = j ZZT0 A (+1)
P1 ()h(2)
(kr) sin
kr

(2.94)

H (r, , ) = YT E
H (r, , ) = YT E
( + 1) (2)
h (kr) cos
kr
By applying eqs.(2.87) and (2.88) for a cone half-angle = 0 we will have,
Hr (r, , ) = jAY0 P1 (cos )

15

See Appendix I and also M.Abramowitz, I.Stegun: Handbook of mathematical functions, New
York: Dover Publ., 1965, pp.332 and .
16
See Appendix I and also M.Abramowitz, I.Stegun: ibid., pp.437 and .

M. Orece: Antennas 2013, Aperture antennas

63

respectively for the conditions on E and E :




dP1 () 
P 1 (0 )
+
=0

d =0
sin 0
jZ0 tan kh =

1
)
j Z0 ( + 1)P (
0
kr P 1 ( )
1

0

+ dPd () 
sin 0
=0

(2.95)

which, combined together, become:


tan kh =

( + 1) sin 0
kr 1/

(2.96)

From eq.(2.96) we may observe as the condition with = 1 is particularly


interesting, because in this case the dependance of the transverse components of the
eld on is the same, and therefore the eld topography is perfectly symmetric. This
condition, called balanced hybrid mode, corresponds to a resonance of the grooves,
and it is when h = /4 (or odd multiples of such value). In this case, the separation
surface between corrugations and outer medium behaves as a PEC surface for the
azimuthal component, and as a PMC surface for the longitudinal component.
The condition h = /4 is clearly fullled at a single frequency; for higher or lower
frequencies the groove will have a reactance respectively capacitive or inductive, and
it will be = 1. As a consequence, the eld topography will be no longer symmetric:
however, in the case of a capacitive wall, the asymmetry is not very high, and the
edge taper is such to still maintain low sidelobes. Conversely, an inductive wall
increases the eld at the edges, and it is therefore a condition to be avoided.
Consequently, the transition between smooth and corrugated waveguide must
not be done (as it could seem reasonable) with a gradual increase of the grooves
depth from 0 to /4, because this will create in the waveguide a region with an
inductive wall. On the contrary, the transition can start with a groove of depth
/2 (equivalent to a short circuit) and then reducing the groove depth down to the
desired value.
The normalized radiation patterns, having as a parameter the maximum phase
error in wavelengths, are shown in g.(2.37) for the hybrid balanced mode condition,
where the radiation patterns in the E and H plane are the same. The gure may be
compared with gs.(2.17-2.26) for rectangular and smooth circular horns.
The behavior of the phase pattern, taking dierent reference points (i.e. phase
centers), is shown in g.(2.38) for the case of a maximum phase error of 0.25. L
is the axial length, and D is the distance of the reference point from the aperture
(with positive sign when to the vertex of the horn).
Due to their very interesting radiating characteristics, corrugated horns have
been in the past decades deeply investigated my many researchers: an interesting

M. Orece: Antennas 2013, Aperture antennas

64

10

Rel.gain (dB)

15
0.25

0.5

0.75

s=1

20

25
0.125
30

35

40
0

10

15

theta

Figure 2.37: Normalized radiation patterns for a corrugated horn in the hybrid
balanced mode, with various maximum phase errors.
10

d/L=0

0.138

0.224

gradi

1.

10

15

20

25
0

0.5

1.5

2.5
u

3.5

4.5

Figure 2.38: Phase patterns for a corrugated horn in the hybrid balanced mode,
with maximum phase error of 0.25).

M. Orece: Antennas 2013, Aperture antennas

65

historical review was published years ago on the subject17 . Design criteria and
more detailed analyses can be found in 18 .
The practical realization of a corrugated horn may be a challenging task, because
the realization of many narrow deep grooves, especially if perpendicular to the axis,
is not so easy. For this reason this type of horn is very expensive and it is used
mainly when high quality antennas are required, and there are not stringent cost
limitations. However, more recent solutions with the grooves parallel to the axis
allow easier manufacturing and reduced costs. An example is shown in g.(2.39) for
a feed of a VSAT in Ka band.

Figure 2.39: Corrugated horn with grooves parallel to the axis.

Sections 2.10 - 2.24 under revision


17
B.Mc.A.Thomas: A review of the early developments of Circular Aperture Hybrid-Mode
Corrugated Horns, IEEE Transactions Antennas Propagat., AP-34, pp.930-935, July 1986
18
P.J.B. Clarricoats, A.D. Olver: Corrugated horns for microwave antennas, London: Peter
Peregrinus Ltd., 1984

S-ar putea să vă placă și