Notch Antennas

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

NASA/TM—2004-213057

Notch Antennas

Richard Q. Lee
Glenn Research Center, Cleveland, Ohio

July 2004
The NASA STI Program Office . . . in Profile

Since its founding, NASA has been dedicated to • CONFERENCE PUBLICATION. Collected
the advancement of aeronautics and space papers from scientific and technical
science. The NASA Scientific and Technical conferences, symposia, seminars, or other
Information (STI) Program Office plays a key part meetings sponsored or cosponsored by
in helping NASA maintain this important role. NASA.

The NASA STI Program Office is operated by • SPECIAL PUBLICATION. Scientific,


Langley Research Center, the Lead Center for technical, or historical information from
NASA’s scientific and technical information. The NASA programs, projects, and missions,
NASA STI Program Office provides access to the often concerned with subjects having
NASA STI Database, the largest collection of substantial public interest.
aeronautical and space science STI in the world.
The Program Office is also NASA’s institutional • TECHNICAL TRANSLATION. English-
mechanism for disseminating the results of its language translations of foreign scientific
research and development activities. These results and technical material pertinent to NASA’s
are published by NASA in the NASA STI Report mission.
Series, which includes the following report types:
Specialized services that complement the STI
• TECHNICAL PUBLICATION. Reports of Program Office’s diverse offerings include
completed research or a major significant creating custom thesauri, building customized
phase of research that present the results of databases, organizing and publishing research
NASA programs and include extensive data results . . . even providing videos.
or theoretical analysis. Includes compilations
of significant scientific and technical data and For more information about the NASA STI
information deemed to be of continuing Program Office, see the following:
reference value. NASA’s counterpart of peer-
reviewed formal professional papers but • Access the NASA STI Program Home Page
has less stringent limitations on manuscript at http://www.sti.nasa.gov
length and extent of graphic presentations.
• E-mail your question via the Internet to
• TECHNICAL MEMORANDUM. Scientific [email protected]
and technical findings that are preliminary or
of specialized interest, e.g., quick release • Fax your question to the NASA Access
reports, working papers, and bibliographies Help Desk at 301–621–0134
that contain minimal annotation. Does not
contain extensive analysis. • Telephone the NASA Access Help Desk at
301–621–0390
• CONTRACTOR REPORT. Scientific and
technical findings by NASA-sponsored • Write to:
contractors and grantees. NASA Access Help Desk
NASA Center for AeroSpace Information
7121 Standard Drive
Hanover, MD 21076
NASA/TM—2004-213057

Notch Antennas

Richard Q. Lee
Glenn Research Center, Cleveland, Ohio

National Aeronautics and


Space Administration

Glenn Research Center

July 2004
Available from
NASA Center for Aerospace Information National Technical Information Service
7121 Standard Drive 5285 Port Royal Road
Hanover, MD 21076 Springfield, VA 22100

Available electronically at http://gltrs.grc.nasa.gov


Notch Antennas
Richard Q. Lee
National Aeronautics and Space Administration
Glenn Research Center
Cleveland, Ohio 44135

Introduction
A notch antenna, also known as tapered slot antenna (TSA), is an end-fire traveling wave antenna.
Like microstrip antennas, the TSA features low profile, light weight, easy fabrication by photo-
etching, conformal installation and compatibility with microwave integrated circuits (MIC).
Performance wise, TSA has demonstrated multi-octave bandwidth, moderate gain (7–10 dB)
and symmetric E- and H- plane beam patterns. However, TSA lacks the versatility of microstrip
antennas in terms of multifunction operations such as dual-frequency, dual-polarization etc., and
loses its planar architecture when used as circularly polarized antennas or in 2-D arrays. The
antenna was first proposed by Lewis et al [1] in 1974. Since then, much technological advances
have been made in both designs and applications. In designs, various numerical techniques have
been developed and applied to aid the design of the TSA antennas [2–4]. Full-wave numerical
techniques based on Method of Moments are now commercially available providing the ability to
shorten the design cycle as have been demonstrated in a recent CAD benchmark demonstration
[5]. In applications, TSAs have been mainly used for spatial-power combining and as focal-plane
feed arrays for imaging reflector systems [6–7]. In more recent applications, TSA arrays capable of
multi-function operations are emphasized. Examples of these include the dual-band dual-polarized
TSA array of nested lattice architecture [8], and multi-frequency full-duplex phased array system
with notch elements [9].

This chapter will focus on the most recent advances in TSA and their applications. In order to
provide some basic understanding of the design issues, the chapter will begin with some
discussions on the design of single tapered slot antennas. Then, TSA arrays and their applications
will be highlighted.

Basic Geometries
A typical TSA consists of a tapered slot section which is produced by the gradual widening of a
slot line. The antenna, excited by a slotline feed, is fabricated on a metallization layer with or
without a supporting dielectric substrate. The tapered slot section which constitutes the radiating
region can take on any of three geometric profiles: (1) nonlinear taper (Vivaldi, tangential or
parabolic), (2) linear taper (LTSA), and (3) constant width (CWSA) [10]. The majority of previous
studies have been focused on TSA with linear and exponential profiles. Prasad and Mahapatra [11]
first introduced the linearly tapered slot antenna (LTSA) in 1979. Their antenna of about one free
space wavelength long was etched on a 25-mil alumina substrate. In the same year, Gibson [12]
reported a TSA with exponentially tapered profile, also known as AVivaldi@ antenna, which
demonstrated a bandwidth of 8–40 GHz. The schematics of the LTSA and the Vivaldi antennas
are shown in fig. 1. In later years, most of the subsequent developments of TSA arrays are derived
mainly from these two antenna types.

NASA/TM—2004-213057 1
In order to facilitate integration of antenna and devices at sub-millimeter wave frequencies, a new
variant of the TSA, generally known as V-antennas, was introduced in 1979 for coupling radiation
into diode mixers [13]. Years later, other applications of V-antennas were demonstrated at
microwave frequencies [14–15]. These antennas with finite width ground planes can enhance the
antenna performance by suppressing the surface waves. Another advantage of the V-antennas is
its compatibility with uniplanar feed structures which allow to excite the antenna with either a
microstrip line or a waveguide [14]. The schematics of a V-LTSA and a V-Vivaldi excited by a
coplanar stripline are shown in fig. 2.

Another variant of the conventional TSA is the antipodal tapered slot antenna. In practice, the
conventional planar TSA is fed by a balanced slotline. One serious drawback of the conventional
TSA is in the fabrication and impedance matching of the slotline. Slotline fabricated on a low
dielectric constant substrate has relatively high impedance which makes matching to a microstrip
feed very difficulty. The antipodal TSA is designed to overcome this impedance matching
difficulty; however, it has exhibited very poor cross-polarization characteristics. The antenna is
formed by gradually flaring the strip conductors of the balanced microstrip on opposite sides of
the dielectric substrate with respect to the antenna axis, thus allowing the antenna to be directly fed
by a microstrip feed [16]. To circumvent the high cross-polarization problem, a balanced antipodal
Vivaldi consisting of three metallization layers was introduced. This new antenna, shown in fig. 3,
has demonstrated –15 dB lower cross-polarization across a 18:1 band as compared to the
conventional Vivaldi antenna [17].

NASA/TM—2004-213057 2
TSA Design
For years, the design of TSA has been primarily based on empirical approach, which as an initial
step, could start with the following simple guidelines:

The aperture width of slot: W ≥ λ0

1.05 ≤ c/v ≤ 1.2

The effective thickness: teff ≤ .03 ≤ λ0

The taper angle, 2 α, is typically 5 to 12º

The length of the TSA, L, is typically 2 to 12 λ0

where c is the velocity of light, v is the velocity of the field along the slot, and λ0 is the free-space
wavelength. With the availability of electromagnetic simulators, engineers can now check the
accuracy of their designs before carrying out the actual fabrication and characterization. In general,
the design of TSA involves two major tasks: (1) the design of a broadband transition and feed
structure with very wide frequency range and low return loss, and (2) determining the dimensions
and shape of the antenna in accordance with the required beamwidth, side lobe, and back lobe etc.
over the operating frequency range. The geometric parameters such as length, width, dielectric
thickness, ground-plane size, and taper-profile which have direct impact on the impedance,
directivity, bandwidth, and radiation pattern of the antenna will be discussed in the sections below.

Feeding Methods
Almost always planar tapered slot antennas such as LTSA and Vivaldi are excited by a slotline;
thus, feeding a TSA requires building a transition between the slotline and other transmission
media. In the case of microstrip or coaxial lines, the transition is a balun. In general, the transition
should have small parasitic inductance and capacitance for wide bandwidth, and hence,
electromagnetic coupling is preferred over wire bonding or solder connection. In the following,
we will consider three basic feed structures: (1) coplanar waveguide (CPW), (2) microstrip line,
and (3) coaxial lines.

NASA/TM—2004-213057 3
Coplanar Waveguide-Feed: CPW possesses many useful design characteristics such as low
radiation loss, less dispersion and uniplanar configuration. In [10], we have presented several
feeding techniques where the transmission media are CPW which excite the TSA via
electromagentic coupling. Figure 4(a) shows a strip-to-slot coupled LTSA excited by a conductor-
backed finite ground-plane CPW feed. This antenna has demonstrated a VSWR of 2:1 over
nearly 3:1 bandwidth as indicated by the measured return loss shown in fig. 4(b).

The TSA can also be excited by direct slot-to-slot feeding where the CPW and the antenna are on
the same side of the substrate. In order to excite the odd mode where the electric fields in both
slots of the CPW have the same amplitude but opposite direction, a bondwire is usually placed
across the slots to keep the two ground planes at equal potential. The CPW feed enables direct
integration of active solid state devices with the antenna. Figure 5 shows a LTSA integrated with a
FET amplifier which demonstrated a gain of more than 14 dB over a bandwidth of 1.75 GHz [18].

NASA/TM—2004-213057 4
CPW can be designed as a tapered CPW with constant characteristic impedance, but with different
ratio of slot to center strip widths [19]. This unique feature is useful for exciting twin tapered slot
antennas of large separations.

Figure 6 shows the application of tapered CPW transitions to twin Vivaldi antennas to achieve
pattern diversity for transponders and monopulse systems through even and odd mode excitations
[20]. With this approach, exciting the even mode produces a maximum radiation, while exciting
the odd mode produces a minimum radiation in the boreside direction.

NASA/TM—2004-213057 5
Microstrip Line Feed: Microstrip line is in widespread use in power combining circuit. Since
microstrip line is an unbalanced line and slotline is a balanced line, feeding a TSA with a
microstrip line requires a balanced-to-unbalanced transition (balun) to avoid compromising the
broadband antenna performance. The most common microstrip/slot transition, the Marchand balun
[21], has demonstrated a VSWR of 2:1 over an octave bandwidth with an integrated wideband
V–Vivaldi antenna [22]. The transition and the corresponding transmission line model are shown
in fig. 7. The balun consists of four quarter-wave sections with the end open-circuited section
extended past the center of the slotline by about one quarter of a guided wavelength. Further
broadening of the bandwidth can be achieved when the microstrip is terminated by a radial stub
and the slot line is terminated by an elliptical shaped cavity [10].

The double-Y balun is another microstrip-to-slotline transition which has no inherent bandwidth
limitation other than parasitic inductances and capacitances. As indicated in fig. 8, the double-Y
balun is based on the 6-port double-Y junction which consists 3 balanced and 3 unbalanced lines
placed alternately around the center of the structure. For the balun to provide perfect transmission
between an opposite balanced and unbalanced ports, opposite pairs of lines should have reflection
coefficients with opposite phases; thus requiring one pair of lines to be short circuit while the other
open circuit. TSA with double-Y balun feed structure has achieved an impedance bandwidth of
more than 2:1 [23].

NASA/TM—2004-213057 6
The balun requirement often complicates the feed design and limits the antenna bandwidth if it is
not designed properly. The V- and antipodal tapered slot antennas offer a way to circumvent this
problem since they are not directly fed by a slotline. Figure 9 shows a V–LTSA with a uniplanar
microstrip-to-coplanar stripline feed [24]. In the feed network, a microstrip line of characteristic
impedance of 50 Ω and width Wl is coupled to two orthogonal microstrip lines of characteristic
impedance of 70 Ω and width W2 through a quarter-wave impedance transformer of width W1 .
The mean path length of the folded loop “a through b” is equal to half a guide wavelength λg of
the 70 Ω microstrip line at the design frequency. However, for wideband operation, the mean
path length needs to be increased to 0.638 λg to compensate for the right-angle bend parasitic
elements. This design ensures that the signal phase from points “a” to “b” are 180 degree out
of phase; thus the coupled microstrip lines are excited in the odd mode with the electric field
predominantly across the gap S 2 . The dimension of S 2 is chosen such that the characteristic
impedance of the coupled microstrip line is 50 Ω . The measured return loss of the V–LTSA is
better than 10dB over the frequency range of 5 to 15 GHz.

NASA/TM—2004-213057 7
Figure 10 shows an antipodal LTSA with a microstrip-to-balanced microstrip feed. The balanced
microstrip is formed with the ground plane tapered to a width equal the strip width Wl and the
radius R 2 of the arc is arbitrarily chosen to provide a smooth taper transition for matching the
characteristic impedance of the conventional microstrip to the balanced microstrip.

NASA/TM—2004-213057 8
The electric field lines at different cross section along the feed and the antenna are illustrated
in fig. 11. The electric field lines which are spread out in the conventional microstrip structure
concentrate between the metal strip of the balanced microstrip, and finally rotate while traveling
along the LTSA. The field rotation inadvertently introduces higher cross-polarizations in these
types of antennas. The antipodal LTSA has demonstrated over 4:1 (8–32 GHz) bandwidth for
2:1 VSWR at the design frequency of 18 GHz [25].

Coaxial feed: Coaxial line is useful as a feed structure because of its compatibility with the
slotline, coplanar microstrip, or balanced microstrip to form wideband transitions; thus, it can be
used to excite all three variants of TSA described above. The disadvantage is that it is not planar
and has high losses at high frequencies. Further, coaxial line is an unbalanced feed line. All of
the currents flow inside the line, i.e. the inner connector and the inside of the shield. Feeding a
balanced antenna with unbalanced coaxial feed may cause currents to flow on the outside of the
shield, which could results in significant power loss and serious distortion in radiation pattern.
Since the magnitude of the current on the outside shield is a function of the impedance to ground
of the conductor, one common approach to cutoff the flow of current is with an open circuit which
is created by placing a “skirt” one quarter of a wavelength long around the outside shield and
shorted to the shield at one quarter of a wavelength away from the load.

The most direct way of exciting a TSA with a coaxial feed is to extend its center conductor over
the slotline section of the TSA and anchor the coaxial feed with solder connection to the ground
plane. Recently, an improved balun design which has direct application to V-antennas has been
proposed [26]. The balun impedance is tapered by cutting open the outer wall of the coax so that
the reflections at the input are minimized. The length of the balun is determined by the lowest
operating frequency and the maximum reflection coefficient which occurs in the passband. Figure
12 shows the cross-section of the tapered coaxial line balun and its integration with a V–Vivaldi
antenna.

NASA/TM—2004-213057 9
Radiation Characteristics
A good understanding of the effects of the design parameters (length, width, dielectric thickness,
ground-plane size and taper profile etc.) on antenna characteristics would facilitate the task of
antenna design. TSA is known to have good performance over a wide bandwidth. In practice, the
operating bandwidth is often limited by the slotline-to-feedline transition and by the finite width of
the antenna. To achieve wideband operation requires the TSA to operate in a traveling-wave mode
with perfect impedance match at both the feed transition and the slot termination. In this section,
some important characteristics of TSA, which are based on experimental data as well as from
previous publications, will be presented.

Effect of length, width and taper profiles: Tapered slot antennas radiate in the endfire direction
with fairly symmetric radiation patterns, but have high cross-polarization levels in the diagonal
plane. The lowest cross-polarization level in the diagonal plane is about 10 to 15 dB higher
than that of the principal planes which are typically better than –15 dB. In general, the cross-
polarization characteristics of planar TSA are superior to those corresponding to their antipodal
counterparts.

Being a traveling wave antenna, the phase velocity and guide wavelength, λg, varies with any
change in the geometrical and material parameters of the antenna, which in turn impacts the
radiation characteristics of the antenna. The gain of a TSA increases with the length, L, of the
antenna typically from a few dB to over 10 dB as L increases from 2–5λg [10]. Maximum
measured gain of 16–17 dB with radiation efficiency of 80 percent has been reported for long
TSA with L greater than 6 λ0 [27]. Similarly, the beamwidths, being inversely proportional to
gain, decrease rapidly as the length is increased; however, the H-plane beamwidth varies more
slowly in comparison to the E-Plane beamwidth particularly for L less than 5λg.

NASA/TM—2004-213057 10
The taper profile has been found to have strong effects on both the beamwidth and sidelobe level
of the antenna. In general, the beamwidths are narrower for the CWSA, followed by the LTSA,
and then Vivaldi for antennas with the same length, same aperture size, and on the same substrate.
Similarly, the sidelobes are highest for the CWSA, followed by the LTSA, and the Vivaldi.

As a traveling wave antenna, the H-plane beamwidth follows 1 L dependence, while the
E-plane beamwidth depends more on the aperture width or tapered angle [28]. Varying the
tapered angle will change the phase velocity and hence λg , which will in turn change the E-plane
beamwidth. Thus, constant beamwidth in both E- and H-plane can be achieved with proper
choices of L and tapered angles. Gibson obtained approximately constant beamwidth in both
E- and H-plane for a Vivaldi antenna fabricated on alumina substrate with tapered profile
defined by [12].

y = ± 0.125exp(0.052x)

Recently, approximately constant E- and H-plane beamwidths over a 6 to 18 GHz bandwidth


was demonstrated for a balanced antipodal antenna with elliptical radiating taper [17], and nearly
identical E- and H-plane 3-dB beamwidth was observed for a LTSA with tapered angles in the
range of 15 to 20 degrees [29].

Effect of dielectric: The performance of a TSA is sensitive to the thickness and dielectric constant
of the dielectric substrate. The presence of a dielectric substrate has the primary effect of
narrowing the main beam of the antenna. The effects of dielectric substrates on the E-and H-plane
beamwidths of a Vivaldi antenna on polyethylene have been investigated at 18.5 GHz [30]. By
varying substrate thickness from 10 to 15 and 27 mils, the H-plane beamwidths become narrower,
while the E-plane beamwidths remain essentially constant, indicating that the dielectric has a
strong effect on the H-fields. Increasing the dielectric thickness generally results in increased
gain, but with higher sidelobes. For good performance, a TSA should have an effective substrate
thickness in the range of 0.0025λ0 ≤ teff ≤ 0.028λ0 where teff = t ( ε r –- 1) is the effective
thickness of the substrate. For substrate thickness above the upper bound of 0.028λ0, unwanted
substrate modes develop which degrade the antenna performance resulting in low efficiency
and narrow bandwidth, particularly for dielectrics with high dielectric constants. Further, for
millimeter-wave operations, the upper bound on the effective thickness constrains to using
mechanically fragile substrates with thickness of only a few hundreds of microns. To overcome
these problems, it has been proposed to fabricate notch antennas either on thick perforated
substrates with an array of regularly and closely placed holes [31] or on photonic bandgap
dielectrics as shown in fig. 13 [32]. The former approach reduces the effective substrate thickness,
while the latter approach suppresses the excitation of the surface modes.

NASA/TM—2004-213057 11
Ground Plane Effect: Placing a ground plane immediately below the TSA will cause the main
beam to steer away from the endfire direction by about 50º in the H-plane [10]. The amount of
scan will be reduced as the distance between the ground plane and the antenna increases, and
becomes negligible when the distance of separation is sufficiently large. For the E-plane patterns,
no significant change was observed except that the TSA had to be tilted in the elevation to receive
full power. Figure 14 shows measured H-plane patterns for a LTSA over a ground plane with a
foam spacer in between.

NASA/TM—2004-213057 12
Impedance Characteristics
Imput impedance: The impedance characteristic of TSA is important for designing match
transitions; yet very little information is available from previous publications. Based on conformal
mapping method for a fin antenna without a dielectric, Yngvesson et al. reported that the
impedance of a LTSA with thin dielectric is essentially constant at frequencies from 4 to 12.4 GHz
and is close to 80 Ω [33]. In a recent study, the input impedance of LTSA was measured using a
microwave wafer probe and a set of on wafer Thru-Reflect-Line (TRL) slotline standards [34].
The LTSA was excited through a short length of a slotline by a ground-signal microwave probe
(Picoprobe, Inc.). In the measurement, the reflection coefficient of the LTSA is de-embedded
from the measured reflection coefficient at the input terminal of the slotline using the NIST
de-embedded software [35]. Figure 15 shows the real and imaginary parts of the input impedance,
Z in , versus frequency for a LTSA with α= 5º and L = 2.54cm.

The impedance plot exhibits a series of resonance over the frequency band from 2 to 26.5 GHz.
These resonances are caused by imperfect impedance match transitions at the feed end and the
termination end of the LTSA. The first resonance mode at frequencies below 4 GHz has very large
input impedance, typically over 1000 Ω , which could be caused by large electric fields associated
with a very small W/λ0. For frequencies above 4 GHz, the impedances are bounded between
145 Ω and 40 Ω with a mean value of about 92 Ω . Similar results were also obtained for LTSA
of different lengths and tapered angles [10].

Impedance matching: Impedance matching of TSA has commonly been performed at the feed
end of the antenna. Recently, an approach illustrated in fig. 16 for impedance matching at the
termination end of the antenna has been proposed [36].

NASA/TM—2004-213057 13
The new approach uses a two-step dielectric transformer to match the antenna to free space. The
de-embedded real and imaginary parts of the input impedance with and without the dielectric
transformer are shown in fig. 17.

Mutual impedance: Mutual coupling can produce profound impacts on the performance of an
antenna array causing impedance mismatch and scanning blindness particularly for antennas on
thick dielectric substrates. In general, mutual coupling can occur between two TSAs arranged
either in coplanar or stacked configurations. Recently, the mutual coupling between two LTSAs
in close proximity and excited by two calibrated ground-signal microwave probes was obtained
from an S-parameter measurement of | S21| [37]. Figure 18 shows the measured mutual coupling
between two LTSAs in coplanar and stacked arrangement as a function of separations. Results
indicate that the mutual coupling decreases with the distance of separation.

NASA/TM—2004-213057 14
The coupling coefficients as a function of frequencies were also measured first with the LTSAs
fabricated on a continuous substrate and then repeated after a narrow slot was cut in the dielectric.
In the latter case, coupling between antennas were confined to space waves only. Results indicate
stronger mutual coupling for the TSAs on continuous substrate as a result of additional coupling
through the supporting dielectric; however, both cases display decreasing |S21| with frequencies
[10]. This is due to the fact that at higher frequencies the separation is electrically larger and
separations between antennas are greater with respect to wavelengths.

TSA Array and Applications


Over the years, many different applications of TSA arrays have been reported in the literature.
TSA arrays have been successfully applied as focal plane feed arrays for a multi-beam imaging
reflector system. For the same beamwidth, a TSA array occupies about 1/4 of the area of a feed
array with waveguide elements; thus, imaging system with TSA focal plane array can be designed
to yield high angular resolution [27].

TSA arrays have also been advantageously used for spatial power combining to achieve power
amplification and harmonic generation. Recently, a 120-watt, X-band spatially combined solid-
state amplifier has been demonstrated [6]. The spatial combiner consists of six identical “trays”
with each containing four input/output broadband TSA integrated with four MMIC amplifiers via
a slot-to-microstrip transition. Space power combining can be applied to large antenna aperture
such as space-fed lens arrays. The concept of space power amplification has been demonstrated
experimentally with a three-element LTSA array module. Figure 19 shows a photo of the array
module and the measured H-plane pattern.

NASA/TM—2004-213057 15
In this approach, an array of active antenna modules, constructed from non-planar TSA’s and
monolithic microwave integrated circuit (MMIC) multistage GaAs power amplifiers, receives the
signal at lower power and after amplification re-radiates the signal into free space. Polarization
diversity is employed to permit accurate measurement of the amplified signal radiated from the
transmitting array. With the amplifiers turned on, the three-element active array module produced a
gain of 30 dB at 20 GHz [38]. Replacing the MMIC multistage power amplifier with a broadband,
GaAs MMIC distributed amplifier, frequency multiplication and space power combining was
demonstrated with the same active LTSA array module which receives signal at 9.3 GHz, and after
amplification, radiates the second harmonic signal into free space. Results indicated a fundamental-
to-second-harmonic conversion efficiency of 8.1 percent [39]. For experimental details and
measured results, the reader is referred to reference [10].

With the rapid development of the wireless industry in recent years, a K-band circular LTSA array,
shown in fig. 20, has been proposed for mobile communications [40]. The 16-element array is fed
by a 1:16 microstrip line power splitter composed of T-junctions and right-angle bends. The output

NASA/TM—2004-213057 16
ports of the splitter are electromagnetically coupled to the slotline of the LTSA through a
microstrip-to-slotline transition with the slotline and microstrip line characteristic impedances
chosen to be 120 Ω and 100 Ω respectively. The array was placed over a reflecting ground plane
to displace the beam above the horizon, and foam spacers of different thickness are used to control
the amount of scan. Figure 21 shows the measured E- and H-plane patterns for 0.286λ0 separation.
As shown, the circular LTSA array produces an omnidirectional pattern in the azimuthal plane
and a beam displacement of about 28° in the elevation plane.

NASA/TM—2004-213057 17
Using a similar approach, an omni-directional quasi-optical circular array of 12 Vivaldi-coupled
oscillator elements has been designed for LMDS (Local Multipoint Distribution Services)
applications [41]. The array, which is powered from a single direct-current power supply, has
demonstrated an 87 percent power combining efficiency, and 144 mW radiated power at 28 GHz.

TSA array has found applications in intelligent transportation systems. A wide-scan spherical-lens
antenna at millimeter-wave frequency (77 GHz) was recently introduced for automobile radars [42].
The multibeam antenna system is composed of planar TSA which are positioned around a
homogeneous spherical Teflon lens. Beam scanning is achieved by switching between elements.
Figure 22 shows a 33-beam spherical Teflon lens system which has demonstrated an angular
resolution of 5.5º and –3.5 dB crossover of adjacent beams.

In shared aperture technology development where the functionality of several antennas are
combined into one aperture using wideband multiple beam technology, dual-polarized tapered slot
antennas with “double Y” microstrip-to-slotline transition have been used to achieve wideband
operations [43]. To realize horizontal and vertical linear polarization, two TSA elements of like
polarization are first combined into a two-element subarray. Then, two such subarrays are joined
at 90° through a mechanical slot cut in the top of one subarray and the bottom of the other subarray

NASA/TM—2004-213057 18
to form a quad-element subarray module. Figure 23 shows an 8×8 array composed of quad-element
subarray modules as building blocks. To obtain right- and left-hand circular polarization, the
horizontal and vertical polarizations are combined with a 90° phase-offset in a quadrature hybrid
circuit.

Combining the functionality of several antennas into one shared aperture presents many technical
challenges. Recently, a dual-band dual-polarized nested antipodal Vivaldi array has been developed
for a new radio-telescope, the Square Kilometer Array (SKA) [8]. In order to meet the 5:1
bandwidth (0.8 to 4.0 GHz) requirement, the array consists of a lower-band and an upper-band
subarray with each made up of different-size antipodal Vivaldi quad-elements having overlapping
bandwidths. The array of nested lattice architecture has multilevel ground planes which allow
height-adjustment to achieve optimization of the array performance. A sketch of a 20-element dual-
band dual-polarized multilevel nested antipodal Vivaldi array is shown in fig. 24.

NASA/TM—2004-213057 19
In a recent development of a low-cost, multi-frequency and full-duplex phased array transceiver that
transmits at 10 or 19 GHz and receives at 12 or 21 GHz, Vivaldi antennas were used to achieve
wide-band performance. In addition, the system, as illustrated in fig. 25, features a low-loss and
low-cost multi-line phase shifter controlled by dual piezoelectric transducers (PET) and four-
channel microstrip multiplexers [9]. The array system has demonstrated scan angles of about 40º at
10 to 21 GHz using the PET phase controls. The use of PET greatly simplifies the integration and
high insertion loss problems generally associated with solid-state type phased shifters used in
conventional phased arrays.

NASA/TM—2004-213057 20
Conclusion
The objective of this chapter is to provide an overview of notch antennas including underlying
principles and limitations, general design guidelines, performance characteristics, and new
applications. Due to limited coverage, the chapter focuses mainly on LTSA and Vivaldi, while
many other noteworthy works have not been able to include in the discussion. The general design
guideline outlined in the chapter is intended as an initial step in the experimental or analytical
processes leading to the final design. The discussions on feeding techniques and antenna
characteristics are useful in understanding the design and the radiation process of the notch
antennas. Finally, the chapter ends with some recent advances and innovative applications of
TSA arrays based on shared aperture technology.

References
[1] L.R. Lewis, M. Fassett and J. Hunt, “A Broadband Stripline Array Element,”
1974 IEEE AP–S International Symposium, Atlanta, GA June 1974, pp. 335–337.
[2] Ramakrishna Janaswamy, “An accurate Moment Method Model for the Tapered Slot
Antenna,” IEEE Trans. Antenna and Propagation, Vol. 37, No. 12, December 1989.
pp. 1523–1528.
[3] Eric Thiele and Allen Taflove, “FD–TD Analysis of Vivaldi Flared Horn Antennas
and Arrays,” IEEE Trans Antenna and Propagation, Vol. 42, No. 5, May 1994, pp.
633–641.
[4] Liang-Chen Kuo, Meng-Chung Tsai and Huey-Ru Chang, “3–D FDTD Design
Simulation and Experimental Measurement of a Ka-Band Planar Antipodal Linearly-
Tapered Slot Antenna (ALTSA),” IEEE Microwave and Wireless Components
Letters, Vol. 11, No. 9, Sept. 2001, pp. 382–384.
[5] “CAD Benchmark on a Balanced Antipodal Vivaldi Antenna,” Microwave
Engineering Europe (http://www.mwee.com/magazine/2000/CAD-nov.htm),
October 2000.
[6] Nai-Shuo Cheng, Pengcheng Jia, David B. Rench and Robert A. York, “A 120-W
X–Band Spatially Combined Solid-State Amplifier,” IEEE Trans. Microwave Theory
and Techniques, Vol. 47, No. 12, Dec. 1999, pp. 2557–2559.
[7] T.L. Korzeniowski, D.M. Pozar, D.H. Schaubert, and K.S. Yngvesson, “Imaging
System at 94 GHz Using Taperted Slot Antenna Elements,” Proceedings of the 8th
International Conference on Infrared and Millimeter Waves, 1983.
[8] Hung Loui, Jan Peeters Weem and Zoya Popovic, “A Dual-Band Dual3-Polarized
Nested Vivaldi Slot Array With Multilevel Ground Plane,” IEEE Trans. Antennas
and Propagation, Vol. 51, No. 9, September 2003, pp. 2168–2175.
[9] Tae-Yeoul Yun, Chunlei Wang, Paola Zepeda, Christopher T. Rodenbeck, Mathew R.
Coutant, Ming-yi Li, and Kai Chang, “A 10–21 GHz, Low-Cost, Multifrequency, and
Full-Duplex Phased-Array Antenna System,” IEEE Trans AP–S, Vol. 50, No. 5, May
2002, pp. 641–649.
[10] Kai Fong Lee and Wei Chen, Advances in Microstrip and Printed Antennas, Wiley
Interscience, New York 1997. Chapter 9, p. 443.
[11] S.N. Prasad and S. Mahapatra, “A Novel MIC Slot Line Aerial,” Proc. 9th
European Microwave Conf., Brighton, U.K. 1979, pp. 120–124.
[12] P.J. Gibson, “The Vivaldi Aerial,” Proc., 9th European Microwave Conf., Brighton,
U.K. 1979, pp. 101–105.

NASA/TM—2004-213057 21
[13] T.L. Hwang, D.B. Rutledge, and S.E. Schwarz, “Planar Sandwich Antennas for
Submillimeter Applications,” Appl. Phys. Lett. Vol. 34 (1), 1 January 1979, pp. 9–11
[14] R.N. Simons, N.I. Dib, R.Q. Lee, and L.P.B. Katehi, “Integrated Uniplanar
Transition for Linearly Tapered Slot Antenna,” IEEE Trans. Antennas and
Propagation, Vol. 43, No. 9, 1995, pp. 998–1002.
[15] J.J. Lee and S. Livingston, “Wide Band Bunny-Ear Radiating Element,” 1993 IEEE
AP–S International Symposium Digest, June 1993, pp. 1604-1607.
[16] E. Gazit, “Improved Design of the Vivaldi Antenna,” IEE Proc., Part H, Vol. 135,
No. 2, 1988, pp. 89–92.
[17] J.D.S. Langley, P.S. Hall, and P. Newham, “Balanced Antipodal Vivaldi Antenna for
Wide Bandwidth Phased Arrays,” IEE Proc. Antennas Propag., Vol. 143, No. 2, April
1996, pp. 97–102.
[18] X.D. Wu and K. Chang, “Compact Wideband Integrated Activge Slot Antenna
Amplifier,” Elect. Lett. Vol. 29, No.5, March 1993, pp. 496–497.
[19] Guiping Zheng, Ahmed A. Kishk, Allen W. Glisson, and Alexander B. Yakovlev,
“Slot Antenna Fed by a CPW Line with Tapered Transition,” Microwave and
Optical Tech. Lett., Vol. 38, No. 6, September 2003, pp. 465–467.
[20] I. Linardou and C. Migliaccio, “Twin Vivaldi Antenna for Transponders and
Monopulse Systems,” Microwave and Optical Technology Letters, Vol. 27, No. 3,
Nov. 2000, pp. 207–209.
[21] N. Marchand, “Transmission Line Conversion,” Electronics, Vol. 17, 1944,
pp. 142–145.
[22] A.B. Smolders and M.J. Arts, “Wide-Band Antenna Element with Integrated
Balun,” 1998 IEEE AP–S Int. Symposiums Digest, Atlanta, USA, June 1998.
[23] V. Trifunovic and B. Jokanovic, “Review of Printed Marchand and Double-Y
Baluns: Characteristics and Application,” IEEE Trans. On Microwave Theory and
Techniques, Vol. 42, No. 8, Aug. 1994, pp. 1454–1462.
[24] R.N. Simons, N.I. Dib, R.Q. Lee, and L.P.B. Katehi, “Integrated Uniplanar
Transition for Linearly Tapered Slot Antenna,” IEEE Trans. Antennas Propagation,
Vol. 43, No. 9, 1995, pp. 998–1002.
[25] R.N. Simons, R.Q. Lee, and T.D. Perl, “Non-Planar Linearly Tapered Slot Antenna
with Balanced Microstrip Feed,” 1992 IEEE AP–S International Symposium, Vol. 4,
1992, pp. 2109–2112.
[26] P. Knott and A. Bell, “Coaxially-Fed Tapered Slot Antenna,” Electronics Letters,
Vol. 37, No. 18, Aug. 2001, pp. 1103–1104.
[27] K. Sigfrid Yngvesson, T.L. Korzeniowski, Young-Sik Kim, Erik L. Kollberg, and
Joakim F. Johansson, “The Tapered Slot Antenna—A New integrated Element for
Millimeter-Wave Application,” IEEE Trans. Microwave and Techniques, Vol. 37,
No. 2, Feb. 1989, pp. 365–374.
[28] T. Thungren, E.L. Kollberg and K.S. Yngvesson, “Vivaldi Antennas for Single
Beam Integrated Receiver,” Proceedings of the 12th European Microwave
Conference, 1982, pp. 475–480.
[29] P.S. Kooi, T.S. Yeo, and M.S. Leong, “Parametric Studies of the Linearly Tapered
Slot Antenna (LTSA),” Microwave and Optical Tech. Lett., Vol. 4, No. 5, April
1991, pp. 200–206.
[30] D.C. Hogg and W.E. Legg, “A Finline Radiator,” The Bell System Technical
Journal, Vol. 52, No. 7, September 1973, pp. 1249–1253.
[31] J.S. Colbum and Y. Rahmat-Samii, “Printed Antenna Pattern Improvement Through
Substrate Perforation of High Dielectric Constant Material: An FDTD Evalution,”
Microwave and Optical Technology Letters, Vol. 18, No. 1, May 1998, pp. 27–32.

NASA/TM—2004-213057 22
[32] Thomas J. Ellis and Gabriel M. Rebeiz, “MM-Wave Tapered Slot Antennas on
Micromachined Photonic Bandgap Dielectrics,” 1996 IEEE MTT–S International
Symposium Digest, pp. 1157–1160.
[33] K.S. Yngvesson, D.H. Schaubert, T.L. Korzeniowski, E.L. Kolberg, T. Thungren,
and J.F. Johnson, “Endfire Tapered Slot Antennas on Dielectric Substrate,” IEEE
Trans. Antennas Propagat., Vol. 33, No. 12, 1985, pp. 1392-1400.
[34] R.N. Simons and R.Q. Lee, “Linearly Tapered Slot Antenna Impedance
Characteristics,” 1995 IEEE AP–S International Symposium Digest, Vol. 1, 1995,
pp. 170–173.
[35] NIST De-embedding software, Program DEEMBED, Rev. 4.04, 1994.
[36] R.N. Simons and R.Q. Lee, “Characterization of Miniature Millimeter-Wave Vivaldi
Antenna for Local Multipoint Distribution Service,” 49th ARFTG/MTT Conference
Digest, 1997, pp. 95–99.
[37] R.N. Simons and R.Q. Lee, “On-Wafer Characterization of Millimeter-Wave Antennas
for Wireless Applications,” IEEE Trans. Microwave Theory and Techniques, Vol. 47,
No. 1, Jan. 1999, pp. 92–96.
[38] R.N. Simons and R.Q. Lee, “Space Power Amplification with Active Linearly
Tapered Slot Antenna Array,” Proceedings of the 1993 IEEE MTT–S International
Microwave Symposium, Vol. II, pp. 623–626.
[39] R.N. Simons and R.Q. Lee, “Spatial Frequency multiplier with Active Linearly
Tapered Slot Antenna Array,” Proceedings of the 1994 IEEE MTT–S International
Microwave Symposium, Vol. III, pp. 1557–1560.
[40] R.N. Simons, E. Kelly, R.Q. Lee, and S.R. Taub, “Radial Microstrip Slotline Feed
Network for Circular Mobile Communications Array,” 1994 IEEE AP–S International
Symposium Digest, Vol. 2, pp. 1024–1027.
[41] M.J. Vaughan and R.C. Compton, “28 GHz Omni-Directional Quasi-Optical
Transmitter Array,” IEEE Trans. Microwave Theory and Techniques, Vol. 43, No. 10,
1995, pp. 2507–2509.
[42] Bernhard Schoenlinner, Xidong Wu, Jim P. Ebling, George V. Eleftheriades, and
Gabriel M. Rebeiz, “Wide-Scan Spherical-Lens Antennas for Automotive Radars,”
IEEE Trans. on Microwave Theory and Techniques, Vol. 50, No. 9, Sept. 2002,
pp. 2166–2175.
[43] T.A. Axness, R.V. Coffman, B.A. Kopp, and K.W. O’Haver, “Shared Aperture
Technology Development,” Johns Hopkins APL Technical Digest, Vol. 17, No. 3,
1996, pp. 285–294.

NASA/TM—2004-213057 23
Form Approved
REPORT DOCUMENTATION PAGE OMB No. 0704-0188
Public reporting burden for this collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources,
gathering and maintaining the data needed, and completing and reviewing the collection of information. Send comments regarding this burden estimate or any other aspect of this
collection of information, including suggestions for reducing this burden, to Washington Headquarters Services, Directorate for Information Operations and Reports, 1215 Jefferson
Davis Highway, Suite 1204, Arlington, VA 22202-4302, and to the Office of Management and Budget, Paperwork Reduction Project (0704-0188), Washington, DC 20503.
1. AGENCY USE ONLY (Leave blank) 2. REPORT DATE 3. REPORT TYPE AND DATES COVERED
July 2004 Technical Memorandum
4. TITLE AND SUBTITLE 5. FUNDING NUMBERS

Notch Antennas

WBS–22–302–20–1D
6. AUTHOR(S)

Richard Q. Lee

7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES) 8. PERFORMING ORGANIZATION


REPORT NUMBER
National Aeronautics and Space Administration
John H. Glenn Research Center at Lewis Field E–14503
Cleveland, Ohio 44135 – 3191

9. SPONSORING/MONITORING AGENCY NAME(S) AND ADDRESS(ES) 10. SPONSORING/MONITORING


AGENCY REPORT NUMBER
National Aeronautics and Space Administration
Washington, DC 20546– 0001 NASA TM—2004-213057

11. SUPPLEMENTARY NOTES

Responsible person, Richard Q. Lee, organization code 5640, 216–433–3489.

12a. DISTRIBUTION/AVAILABILITY STATEMENT 12b. DISTRIBUTION CODE

Unclassified - Unlimited
Subject Category: 32 Distribution: Nonstandard
Available electronically at http://gltrs.grc.nasa.gov
This publication is available from the NASA Center for AeroSpace Information, 301–621–0390.
13. ABSTRACT (Maximum 200 words)

Notch antennas, also known as the tapered slot antenna (TSA), have been the topics of research for decades. TSA has demonstrated
multi-octave bandwidth, moderate gain (7 to 10 dB), and symmetric E- and H- plane beam patterns and can be used for many
different applications. This chapter summarizes the research activities on notch antennas over the past decade with emphasis on their
most recent advances and applications. This chapter begins with some discussions on the designs of single TSA; then follows with
detailed discussions of issues associated with TSA designs and performance characteristics. To conclude the chapter, some recent
developments in TSA arrays and their applications are highlighted.

14. SUBJECT TERMS 15. NUMBER OF PAGES


29
Slot antenna; Antenna array; Antenna design 16. PRICE CODE

17. SECURITY CLASSIFICATION 18. SECURITY CLASSIFICATION 19. SECURITY CLASSIFICATION 20. LIMITATION OF ABSTRACT
OF REPORT OF THIS PAGE OF ABSTRACT
Unclassified Unclassified Unclassified
NSN 7540-01-280-5500 Standard Form 298 (Rev. 2-89)
Prescribed by ANSI Std. Z39-18
298-102

You might also like