Sunteți pe pagina 1din 26

LECTURE 12: Aperture Antennas – Part I

(The uniqueness theorem. The equivalence principle. The application of


the equivalence principle to aperture problem. The uniform rectangular
aperture. The tapered rectangular aperture.)

Introduction
Aperture antennas constitute a large class of antennas, which emit
electromagnetic wave through an opening (or aperture). These antennas
have close analogs in acoustics: the megaphone and the parabolic
microphone. The pupil of the human eye is a typical aperture receiver
for optical EM radiation. At radio and microwave frequencies, horns,
waveguide apertures and reflectors are examples of aperture antennas.
Aperture antennas are of common use at UHF and above. It is because
aperture antennas have their gain increase as ∼ f 2 . For an aperture
antenna to be efficient and have high directivity, it has to have an area
comparable or larger than λ 2 . Obviously, these antennas would be
impractical at low frequencies. Another positive feature of the aperture
antennas is their near-real valued input impedance and geometry
compatibility with waveguide feeds.
To facilitate the analysis of these antennas, the equivalence principle
is applied. This allows us to carry out the far-field analysis in the outer
(unbounded) region only, which is external to the radiating aperture and
the antenna. This requires the knowledge of the tangential field
components at the aperture, as it follows from the equivalence principle.

1. Uniqueness theorem
A solution is said to be unique if it is the only one possible among a
given class of solutions.
The EM field in a given region V[ S ] is uniquely defined if
- all sources are given;
- either the tangential Eτ components or the tangential Hτ
components are specified at the boundary S.

1
The uniqueness theorem is proven by making use of the Poynting’s
theorem in integral form:
∫∫ E × (
H *
ds + jω ∫∫∫ µ |)H |2
−ε | E |2
dv + ∫∫∫ σ (
| E |2
dv = )
S V[ S ] V[ S ]
(12.1)
(
− ∫∫∫ E ⋅ J + H ⋅ M dvi* * i
)
V[ S ]

Poynting’s theorem states the conservation of energy law in EM systems.


One starts with the supposition that a given EM problem has two
solutions (due to the same sources and the same boundary conditions):
( ) ( )
E a , H a and E b , H b . The difference field is then formed:
δ E = Ea − Eb
(12.2)
δH = H −H a b

Since the difference field has no sources, it will satisfy the source-free
form of (12.1):
∫∫ (δ E × δ H )ds + jω ∫∫∫ ( µ | δ H | −ε | δ E | ) dv + ∫∫∫ σ | δ E | dv = 0 (12.3)
* 2 2 2

S V[ S ] V[ S ]

Since both fields satisfy the same boundary conditions at S, then δ E = 0


and δ H = 0 over S. This leaves us with
(
jω ∫∫∫ µ | δ H |2 −ε | δ E |2 dv + ∫∫∫ σ | δ E |2 dv = 0 , )
(12.4)
V[ S ] V[ S ]

which is true only if


ω ∫∫∫ ( µ | δ H |2 −ε | δ E |2 ) dv = 0
V[ S ]
(12.5)
∫∫∫ σ | δ E | dv = 0
2

V[ S ]

If we assume some dissipation, however slight, equations (12.5) are


satisfied only if δ E = δ H = 0 everywhere in the volume V[ S ] . This
implies the uniqueness of the solution. If σ = 0 , which is a physical
impossibility, but is often used approximation, multiple solutions
(δ E , δ H ) may exist in the form of self-resonant modes of the structure

2
under consideration. In open problems, resonance is impossible in the
whole region.
Notice that the uniqueness theorem holds if either δ E = 0 or δ H = 0
is true on any part of the boundary.

2. Equivalence principles
The equivalence principle follows from the uniqueness theorem. It
allows us to build simpler to solve problems. As long as the equivalent
problem preserves the boundary conditions of the original problem for
the field at S, it is going to produce the only one possible solution for the
region outside V[ S ] .

( Eo , H o ) ( Eo , H o )

V[ S ] V[ S ]
J se
( Eo , H o ) ⇒ ( Ee , H e )
S sources S no sources M se
n̂ n̂
(a) Original problem (b) General equivalent
problem

( )

J se = nˆ × H o − H e
(12.6)
(
M se = Eo − Ee × nˆ ) ( Eo , H o )

V[ S ]
J s = nˆ × H o Js
(12.7) no fields
M s = Eo × nˆ S no sources
Ms

The zero-field formulation is often n̂


referred to as Love’s equivalence (c) Equivalent problem
principle. with zero fields

3
One can apply Love’s equivalence principle in three different ways:
(a) One can assume that the boundary S is a perfect conductor. This
eliminates the surface electric currents, i.e. J s = 0 , and leaves
just surface magnetic currents M s , which radiate in the presence
of a perfect electric surface.
(b) One can assume that the boundary S is a perfect magnetic
conductor. This eliminates the surface magnetic currents, i.e.
M s = 0 , and leaves just surface electric currents J s , which
radiate in the presence of a perfect magnetic surface.
(c) Make no assumptions about the materials inside S, and define
both J s and M s currents, which are radiating in free space (no
fictitious conductors behind them). It can be shown that these
equivalent currents create zero fields inside V[ S ] .
All three approaches lead to the same field solution according to the
uniqueness theorem. The first two approaches are not very useful in the
general case of curvilinear boundary surface S. However, in the case of
flat infinite planes (walls), the image theory can be used to reduce the
problem to an open one. Image theory can be successfully applied to
curved surfaces provided the curvature’s radius is large compared to the
wavelength. Here is how one can implement Love’s equivalence
principles in conjunction with image theory.

( Eo , H o ) ( Eo , H o ) no fields ( Eo , H o ) no fields ( Eo , H o )
V[ S ] S S S
n̂ n̂ n̂
sources no sources no sources
⇒ Ms ⇒ 2M s
Js = 0 Js = 0

(a) Original problem (b) Equivalent problem (c) Equivalent problem


- electric wall - images

4
( Eo , H o ) ( Eo , H o ) no fields ( Eo , H o ) no fields ( Eo , H o )
V[ S ] S S S
n̂ n̂ n̂
sources no sources no sources
⇒ Js ⇒ 2J s
Ms = 0 Js = 0

(a) Original problem (b) Equivalent problem (c) Equivalent problem


- magnetic wall - images

The above approach is used to evaluate fields in half-space as excited by


apertures. The field behind S is assumed known. This is enough to
define equivalent surface currents. Using image theory, the open-region
far-zone solutions for the vector potentials, A (resulting from J s ) and F
(resulting from M s ), are found from:
e− jβ r
4π r ∫∫
j β rˆ⋅r ′
A( P ) = µ J s ( r ′) e ds′ (12.8)
S

e− jβ r
F ( P) = ε
4π r S ∫∫ M s (r′)e jβ rˆ⋅r′ds′ (12.9)

Here, r̂ denotes the unit vector pointing from the origin of the coordinate
system to the point of observation P. The integration point Q is specified
through the radius-vector r′ . In the far zone, it is assumed that the field
propagates radially away from the antenna. It is convenient to introduce
the so-called propagation vector:
β = β r̂ , (12.10)
which characterizes both the phase constant and the direction of
propagation of the wave. The vector potentials can then be written as:
e− jβ r
A( P) = µ
4π r S ∫∫ J s (r′)e j β ⋅r′ds′ (12.11)

5
e− jβ r
F ( P) = ε
4π r S ∫∫ M s (r ′)e j β ⋅r′ds′ (12.12)

The relations between the far-zone fields and the vector potentials are
rather simple.
E Afar = − jω ( Aθθˆ + Aϕϕˆ ) (12.13)
H far = − jω ( F θˆ + F ϕˆ )
F θ ϕ (12.14)
Since
EFfar = η H Ffar × r , (12.15)
the total far-zone electric field is found as:
E far = E Afar + EFfar = − jω ( Aθ − η Fϕ )θˆ + ( Aϕ + η Fθ ) ϕˆ  (12.16)
Equation (12.16) involves both vector potentials as arising from both
types of surface currents. Computations are reduced in half if image
theory is used in conjunction with an electric or magnetic wall
assumption.

3. Application of the equivalence principle to aperture problems


The equivalence principle is widely used in the analysis of aperture
antennas. To calculate exactly the far fields, the exact field distribution
at the aperture is needed. In the case of exact knowledge of the aperture
field distribution, all three approaches given above will produce the same
results. However, such exact knowledge of the aperture field distribution
is usually impossible, and certain approximations are used. Then, the
three equivalence-principle approaches produce slightly different results,
the consistency being dependent on how accurate our knowledge about
the aperture field is. Usually, it is assumed that the field is to be
determined in half-space, leaving the feed and the antenna behind a
infinite wall S (electric or magnetic). The aperture of the antenna S A is
this portion of S where we have an approximate knowledge of the field
distribution based on the type of the feed line or the incident wave
illuminating the aperture. This is the so-called physical optics
approximation, which certainly is more accurate than the geometrical
optics approach of ray tracing. The larger the aperture (as compared to

6
the wavelength), the more accurate the approximation based on the
incident wave.
Let us assume that the fields at the aperture are known: Ea , H a , and
they are zero everywhere else at S . The equivalent current densities are:
J s = nˆ × H a
(12.17)
M s = Ea × nˆ
Using (12.17) in (12.11) and (12.12) produces:
e− jβ r
A( P ) = µ nˆ × ∫∫ H a e jβ ⋅r′ds′ (12.18)
4π r S

e − jβ r
F ( P) = −ε nˆ × ∫∫ Ea e j β ⋅r′ds′ (12.19)
4π r S
The radiation integrals in (12.18) and (12.19) will be denoted shortly as:
I H = ∫∫ H a e j β ⋅r′ds′ (12.20)
S

I = ∫∫ Ea e j β ⋅r′ds′
E
(12.21)
S

One can find general vector expression for the far-field E vector
making use of equation (12.16) written as:
E far = − jω A − jωη F × rˆ , (12.22)
where the longitudinal Ar component is to be neglected. Substituting
(12.18) and (12.19) yields:
e− jβ r
E = − jβ
far

4π r  ( )
rˆ × ∫∫  nˆ × Ea − η rˆ × nˆ × H a  e j β ⋅r′ ds′ (12.23)

SA

This is the full vector form of the radiated field resulting from the
aperture field, and it is referred to as the vector diffraction integral (or
vector Kirchhoff integral).

7
We shall now consider a practical case of a flat aperture lying in the
x − y plane with nˆ ≡ zˆ . Then:
e− jβ r
A=µ
4π r
( −I yH xˆ + I xH yˆ ) (12.24)
e− jβ r
F = −ε
4π r
( −I yE xˆ + I xE yˆ ) (12.25)
The integrals in the above expressions can be explicitly written for this
case in which r′ = x′xˆ + y′yˆ :
I xE = ∫∫ Eax ( x′, y′)e jβ ( x′ sinθ cosϕ + y′ sinθ sin ϕ ) dx′dy′ (12.26)
SA

I yE = ∫∫ Ea ( x′, y′)e j β ( x′sinθ cosϕ + y′sinθ sin ϕ ) dx′dy′


y
(12.27)
SA

I xH = ∫∫ H a ( x′, y′)e j β ( x′sinθ cosϕ + y′ sinθ sin ϕ ) dx′dy′


x
(12.28)
SA

I yH = ∫∫ H a ( x′, y′)e j β ( x′ sinθ cosϕ + y′ sinθ sin ϕ ) dx′dy′


y
(12.29)
SA

Note that the above integrals are exactly the double inverse Fourier
transforms of the aperture field’s components.
The vector potentials in spherical terms are:
e− j β r  ˆ
A=µ
4π r 
θ cos θ ( I H
x sin ϕ − I H
y cos ϕ ) + ˆ ( I xH cos ϕ + I yH sin ϕ )  (12.30)
ϕ 
e − jβ r  ˆ
F = −ε θ cosθ ( I xE sin ϕ − I yE cos ϕ ) + ϕˆ ( I xE cos ϕ + I yE sin ϕ )  (12.31)
4π r
By substituting the above expressions in (12.16), one obtains the far E
field components as:
e− jβ r E
Eθ = j β [I x cos ϕ + I yE sin ϕ +
4π r (12.32)
η cosθ (I yH cos ϕ − I xH sin ϕ )]
e− jβ r
Eϕ = j β [-η (I xH cosϕ + I yH sin ϕ ) +
4π r (12.33)
cosθ (I yE cos ϕ − I xE sin ϕ )]

8
For apertures mounted on a conducting plane, the preferred
equivalent model is the one with electric wall with magnetic current
density
M s = 2 ⋅ Ea × nˆ ( ) (12.34)
radiating in open space. The solution, of course, is valid only for z ≥ 0 .
In this case, I H = 0 .
For apertures in open space, the dual current formulation is used.
Then, a usual assumption is that the aperture fields are related as in the
TEM-wave case:
1
H a = zˆ × Ea (12.35)
η
This implies that
1 I yE H I xE
I H
= zˆ × I E
or I H
=− ,I y = (12.36)
η η η
x

This assumption is valid for moderate and high-gain apertures; therefore,


the apertures should be at least a couple of wavelengths in extent. The
above assumptions reduce (12.32)-(12.33) to:
e − j β r (1 + cosθ ) E
Eθ = j βη I x cos ϕ + I yE sin ϕ  (12.37)
4π r 2
Eϕ = j βη
e − jβ r
(1 + cosθ ) I E cos ϕ − I E sin ϕ  (12.38)
4π r 2  y x 

9
4. The uniform rectangular aperture on an infinite ground plane
A rectangular aperture is defined in the x − y plane as shown below.

Ly
x

Lx

If the fields are uniform in amplitude and phase across the aperture, it is
referred to as a uniform rectangular aperture. Let us assume that the
aperture field is y-polarized.
L L
Ea = E0 yˆ , | x |≤ x and | y |≤ y (12.39)
2 2
According to the equivalence principle, we assume an electric wall at
z = 0 , where the equivalent magnetic current density is given by
M se = E × nˆ . Applying image theory, one can find the equivalent sources
radiating in open space as:
M s = 2M se = 2 E0 yˆ × zˆ = 2 E0 xˆ (12.40)

10
The only non-zero radiation integral is:

Lx / 2 Ly / 2

∫ ∫
j β x′ sin θ cos ϕ
I = 2 E0
E
y e dx′ ⋅ e j β y′ sinθ sin ϕ dy′ =
− Lx / 2 − Ly / 2

βL   β Ly  (12.41)
sin  x sin θ cos ϕ  sin  sin θ sin ϕ 
= 2 E0 Lx Ly  2 ⋅  2 
β Lx β Ly
sin θ cos ϕ sin θ sin ϕ
2 2

It is appropriate to introduce the pattern variables:


β Lx
u= sin θ cos ϕ
2
(12.42)
β Ly
v= sin θ sin ϕ
2
The complete radiation fields are found by substituting (12.41) in (12.32)
and (12.33):
e − jβ r sin u sin v
Eθ = j β E0 Lx Ly sin ϕ
2π r u v
(12.43)
e − jβ r sin u sin v
Eϕ = j β E0 Lx Ly cosθ cos ϕ
2π r u v
The total-field amplitude pattern is, therefore:
sin u sin v
| E |= F (θ , ϕ ) = sin 2 ϕ + cos 2 θ cos 2 ϕ ⋅ =
u v
(12.44)
sin u sin v
= 1 − sin 2 θ cos 2 ϕ ⋅
u v

11
The principal plane patterns are:
E-plane pattern (ϕ = π / 2)
 β Ly 
sin  sin θ 
Eθ =  2  (12.45)
 β Ly 
 2 sin θ 
 
H-plane pattern (ϕ = 0)
 βL 
sin  x sin θ 
Eϕ = cosθ  2  (12.46)
 β Lx 
 sin θ 
 2 

Principle patterns for aperture of size: Lx = 3λ , Ly = 2λ

30 30
E-plane
H-plane

60 60

1 0.8 0.6 0.4 0.2


90 90

120 120

150 150

180

12
For electrically large apertures, the main beam is narrow and the
1 − sin 2 θ cos 2 ϕ in (12.44) is negligible, i.e. it is roughly equal to 1 for
all observation angles within the main beam. That is why, in the theory
of large arrays, it is assumed that the amplitude pattern of a rectangular
aperture is:
sin u sin v
f (u , v ) (12.47)
u v
β Lx β Ly
where u = sin θ cos ϕ and v = sin θ sin ϕ .
2 2
Here is a view of the | sin u / u | function for Lx = 20λ and ϕ = 0 (H-
plane pattern):

|sin[20*pi*sin(theta)]/[20*pi*sin(theta)]|
1

0.8

0.6

0.4

0.2

0
-1 -0.5 0 0.5 1
sin(theta)

13
Here is a view of the | sin v / v | function for Ly = 10λ and ϕ = 90 (E-
plane pattern):
|sin(10 π sin(theta))/(10 π sin(theta))|
1

0.8

0.6

0.4

0.2

0
-1 -0.5 0 0.5 1
sin(theta)

Beamwidths
(a) first-null beamwidth

One needs the location of the first nulls in the pattern in order to
calculate the FNBW. The nulls of the E-plane pattern are determined
from (12.45) as:
β Ly
sin θ / θ =θ n = nπ , n = 1, 2,… (12.48)
2
 nλ 
⇒ θ n = arcsin   , rad (12.49)
 Ly 
 
The first null occurs at n = 1 .

14
λ 
⇒ FNBWE = 2θ n = 2arcsin   , rad (12.50)
 Ly 
 
In a similar fashion, FNBWH is determined to be:
λ 
FNBWH = 2arcsin   , rad (12.51)
 Lx 

(b) half-power beamwidth


The half-power point in the E-plane occurs when
 β Ly 
sin  sin θ 
 2 = 1 (12.52)
 β Ly  2
 2 sin θ 
 
or
β Ly
sin θ / θ =θ h = 1.391 (12.53)
2
 0.443λ 
⇒ θ h = arcsin 
 Ly 
, rad (12.54)
 
 0.443λ 
HPBWE = 2 arcsin 
 Ly 
(12.55)
 

A first-order approximation is possible for very small arguments in


(12.55), i.e. when Ly 0.443λ (large aperture):
λ
HPBWE 0.886 (12.56)
Ly
The half-power beamwidth in the H-plane is analogous:
 0.443λ 
HPBWH = 2arcsin   (12.57)
 Lx 

15
Side-lobe level
It is obvious from the properties of the | sin x / x | function that the
first side lobe has the largest maximum of all side lobes, and it is:
sin 4.494
| Eθ (θ = θ s ) |= = 0.217 = −13.26 , dB (12.58)
4.494

When evaluating side-lobe levels and beamwidths in the H-plane, one


has to include the cosθ factor, too. The smaller the aperture, the less
important this factor is.

Directivity
In a general approach to the calculation of the directivity, the total
radiated power Π has to be calculated first using the far-field pattern
expression (12.44).
4π U
D0 = = 4π max (12.59)
ΩA Π rad
Here,
1
U (θ ,ϕ ) = | Eθ |2 + | Eϕ |2  r 2 = U max | F (θ , ϕ ) |2 (12.60)
2η  
2π π
ΩA = ∫ ∫ | F (θ ,ϕ ) | sin θ dθ dϕ
2
(12.61)
0 0
However, in the case of an aperture illuminated by a TEM wave, one
can use a simpler approach. Generally, for all aperture antennas, the
assumption of a uniform TEM wave at the aperture ( E = yE ˆ 0 ),
E
H a = − xˆ 0 , (12.62)
η
is quite accurate (although η is not necessarily the intrinsic impedance of
vacuum). The far-field components in this case were already derived in
(12.37) and (12.38). They lead to the following expression for the
radiation intensity:
β2
U (θ ,ϕ ) = (1 + cosθ )2 | J xE |2 + | J yE |2  (12.63)
32π η
2

16
The maximum value of the function in (12.63) is easily derived after
substituting the radiation integrals from (12.26) and (12.27):
2
β 2
U max =
8π 2η ∫∫ Ea ds′ (12.64)
SA
The integration of the radiation intensity (12.63) over a closed sphere is
in general not easy. It can be avoided by observing that the total power
reaching the far zone must have passed through the aperture in the first
place. In the general aperture case, this power is determined as:
1
Π rad = ∫∫ Pav ⋅ ds = ∫∫ | Ea |2 ds (12.65)
S
2η S A

Substituting (12.64) and (12.65) in (12.59) finally yields:


2

∫∫ Ea ds′
4π SA
D0 = (12.66)
λ ∫∫ | Ea | ds′
2 2

SA

In the case of a uniform rectangular aperture,


| E0 |2
Π = Lx Ly (12.67)

2
 Lx Ly  | E0 |2
U max =   (12.68)
 λ  2η
Thus, the directivity is found to be:
U 4π 4π 4π
D0 = 4π max = 2 Lx Ly = 2 Ap = 2 Aeff (12.69)
Π λ λ λ
The physical and the effective areas of a uniform aperture are equal.

17
5. The uniform rectangular aperture in open space
Now, we shall examine the same aperture when it is not mounted on
a ground plane. The field distribution is the same as in (12.39) but now
the H field must be defined, too, in order to apply the general form of
the equivalence principle with both types of surface currents.
ˆ 0 
Ea = yE
 − Lx / 2 ≤ x′ ≤ Lx / 2
E , (12.70)
H a = − xˆ 0  − Ly / 2 ≤ y′ ≤ Ly / 2
η 
Above, again an assumption was made that there is a direct relation
between the electric and the magnetic field components.
To form the equivalent problem, an infinite surface is chosen again to
extend in the z = 0 plane. Over the entire surface, the equivalent J s and
M s must be defined. Both J s and M s are not zero outside the aperture
in the z = 0 plane because the field is not zero there. Moreover, the field
is not known a priori outside the aperture. Thus, the exact equivalent
problem cannot be built in practice (at least, not by making use of the
infinite plane model).
The usual assumption made is that Ea and H a are zero outside the
aperture in the z = 0 plane, and, therefore, so are the equivalent currents
J s and M s :
M s = −nˆ × Ea = − zˆ × yˆ E0 
 − L / 2 ≤ x′ ≤ L / 2
xˆ  x x
E0  ,

(12.71)
J s = nˆ × H a = zˆ × (− xˆ )  − L y / 2 ≤ y ≤ L y / 2
η 
− yˆ 
Since the equivalent currents are related via the TEM-wave assumption,
only the integral J yE is needed for substitution in the far field
expressions derived in (12.37) and (12.38).

18
Lx / 2 Ly / 2

∫ ∫
j β x′ sin θ cos ϕ
I = 2 E0
E
y e dx′ ⋅ e j β y′ sinθ sin ϕ dy′ =
− Lx / 2 − Ly / 2

 β Lx  sin  β Ly sin θ sin ϕ  (12.72)


sin  sin θ cos ϕ   2 
= 2 E0 Lx Ly  2  ⋅  
β Lx β Ly
sin θ cos ϕ sin θ sin ϕ
2 2
Now, the far-field components are obtained by substituting in (12.37)
and (12.38):
Eθ = C sin ϕ
(1 + cosθ ) sin u sin v
2 u v
(12.73)
Eϕ = C cos ϕ
(1 + cosθ ) sin u sin v
2 u v
where:
e− j β r
C = j β Lx Ly E0 ;
2π r
β Lx
u= sin θ cos ϕ ;
2
β Ly
v= sin θ sin ϕ .
2
The far-field expressions in (12.73) would be identical to those of the
aperture mounted on a ground plane if cosθ were replaced by 1. Thus,
for small values of θ , the patterns of both apertures are practically
identical.
An exact analytical evaluation of the directivity is difficult.
However, according to the approximations made, the directivity formula
derived in (12.66) should provide accurate enough value. According to
(12.66), the directivity is the same as in the case of the aperture mounted
on a ground plane.

19
6. The tapered rectangular aperture on a ground plane
The uniform rectangular aperture has the maximum possible effective
area (for an aperture-type antenna) equal to its physical area. This also
implies that it has the highest possible directivity for all constant-phase
excitations of a rectangular aperture. However, directivity is not the only
important factor in the design of an antenna. A factor that frequently
comes into a conflict with the directivity is the side-lobe level (SLL).
The uniform distribution excitation produces the highest SLL of all
constant-phase excitations of a rectangular aperture. It will be shown
that the reduction of SLL can be achieved by tapering the equivalent
sources distribution from a maximum at the aperture’s center to zero
values at its edges.
One very practical aperture of tapered source distribution is the open
rectangular waveguide. The dominant (TE10) mode has the following
distribution:
 π  − Lx / 2 ≤ x′ ≤ Lx / 2
ˆ 0 cos  x′  , 
Ea = yE (12.74)
 Lx   y − L / 2 ≤ y ′ ≤ L y / 2
y

Ey x

The general procedure for the far-field analysis is the same as before (in
Section 4). The only difference is in the field distribution. Again, only
the integral J yE is to be evaluated.

20
 π  jβ x′ sinθ cosϕ
Lx / 2 L /2
y

I = 2 E0 ∫ cos  x′  e
E
y dx′ ⋅ ∫ e j β y′sinθ sin ϕ dy′ (12.75)
− Lx / 2  Lx  − Ly / 2

The integral of the y variable was already encountered in (12.41):


βL 
Ly / 2 sin  y sin θ sin ϕ 
I ( y ) = ∫ e j β y′sinθ sin ϕ dy′ = Ly  2  (12.76)
β L
− Ly / 2 y
sin θ sin ϕ
2
The integral of the x variable is easily solved:
π 
Lx / 2

I ( x) = ∫ cos  x′  e j β x′sinθ cosϕ dx′ =


− Lx / 2  Lx 
π 
Lx / 2

= ∫ cos  x′  cos ( β x′ sin θ cos ϕ ) + j sin ( β x′ sin θ cos ϕ )  dx′ =


− Lx / 2  Lx 
1 x   π    π   
L /2

= ∫ cos  − β sin θ cos ϕ  x  + cos  + β sin θ cos ϕ  x′ dx′ +


2 − Lx / 2   Lx  

 Lx   
j x   π    π   
L /2

+ ∫ sin  β sin θ cos ϕ −  x′ + cos  β sin θ cos ϕ +  x′ dx′


2 − Lx / 2   Lx    Lx   
 β Lx 
cos  sin θ cos ϕ 
π Lx  2 
⇒ I ( x) = 2
2  π  β Lx
  − sin θ cos ϕ
2 2
 β Lx βL
 
cos  sin  y sin θ sin ϕ 
sin θ cos ϕ 
⇒ J yE = π E0 Lx Ly  2 
 2  (12.77)
   β Ly 
2
 π  β Lx   2 sin θ sin ϕ 
 
 2  − sin θ cos ϕ   
2
  v
 u 
To derive the far-field components, (12.77) is substituted in (12.32) and
(12.33).

21
π cos u sin v
Eθ = − C sin ϕ
2  2  π 2  v
u −   
  2  
(12.78)
π cos u sin v
Eϕ = − C cosθ cos ϕ
2  2  π 2  v
u −   
  2  
where:
e− j β r
C = j β Lx Ly E0 ;
2π r
β Lx
u= sin θ cos ϕ ;
2
β Ly
v= sin θ sin ϕ .
2

Principle plane patterns


In the E-plane, the aperture is not tapered. As expected, the E-plane
principal pattern is the same as that of a uniform aperture.

E-plane (ϕ = 90 ):
 β Ly 
sin  sin θ 
Eθ =  2  (12.79)
 β Ly 
 2 sin θ 
 
H-plane (ϕ = 0 ):
 βL 
cos  x sin θ 
Eϕ = cosθ  2  (12.80)
 β Lx  π 
2 2

 sin θ  −  
 2   2

22
H-plane pattern – uniform vs. tapered illumination ( Lx = 3λ ):

0
uniform
30 30 tapered

60 60

1 0.8 0.6 0.4 0.2 90


90

120 120

150 150

180

The lower SLL of the tapered-source far field is obvious. It is better


seen in the rectangular plot given below. The price to pay for the lower
SLL is the decrease in directivity (the beamwidth of the major lobe
increases).

23
tapered
1 uniform
0.9

0.8
H-plane amplitude pattern

0.7

0.6

0.5

0.4

0.3

0.2

0.1

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1


sin(theta)

The above example of Lx = 3λ is illustrative on the effect of source


distribution on the far-field pattern. However, a more practical example
is the rectangular-waveguide open-end aperture, where the waveguide
operates in a dominant mode, i.e. λ0 / 2 < Lx < λ0 . Here, λ0 is the
wavelength in open space ( λ0 = c / f0 ). Consider the case Lx = 0.75λ .
The principal-plane patterns for an aperture on a ground plane look like
this:

24
0 H-plane
30 30 E-plane

60 60

90 1 0.8 0.6 0.4 0.2


90

120 240

150 210
180

In the above example, a practical X-band waveguide was considered


whose cross-section has the following sizes: Lx = 2.286 cm, Ly = 1.016
cm. Obviously, λ0 = 3.048 cm, and f0 = 9.84 GHz.
The case of a dominant-mode open-end waveguide radiating in free
space can be analyzed following the approaches outlined in this Section
and in Section 5.
The calculation of beamwidths and directivity is analogous to
previous cases. Only the final results will be given here for the case of
the x-tapered aperture on a ground plane.
8  4π 
Directivity: D0 = 2  2 Lx Ly  (12.81)
π λ 
8
Effective area: Aeff = 2 Lx Ly = 0.81Ap (12.82)
π
Note the decrease in the effective area.

25
Half-power beamwidths:
50.6
HPBWE = , deg. (= HPBWE of the uniform aperture) (12.83)
Ly / λ
68.8
HPBWH = , deg. (> HPBWH of the uniform aperture) (12.84)
Lx / λ
The above results are approximate. Better results would be obtained
if the following factors were taken into account:
• the phase constant of the waveguide β g is not equal to the free-
space phase constant β 0 = ω µ 0ε 0 ; it is dispersive;
• the abrupt termination at the waveguide open end introduces
reflection, which affects the field at the aperture;
• there are strong fringe currents at the waveguide walls, which
contribute to the overall radiation.

26

S-ar putea să vă placă și