Fact-checked by Grok 2 weeks ago

Antenna feed

An antenna feed, also referred to as a feed system, is the primary or component in an antenna that supplies radiofrequency energy to or extracts it from secondary radiating structures, such as reflectors, lenses, or arrays, typically via a or . In continuous aperture antennas, it functions as the main radiating , exemplified by a illuminating a to achieve focused beam patterns. For array antennas, the feed generates the coefficients that determine the amplitude and phase distribution across the elements. Antenna feeds play a critical role in optimizing system performance by ensuring efficient energy transfer, minimizing losses, and controlling radiation characteristics like beamwidth, , and sidelobe levels. They are essential in applications including satellite communications, radar systems, , deep-space probes, and terrestrial links, where precise illumination of the antenna aperture directly impacts overall efficiency and signal quality. The design must account for factors such as to the feed line, polarization purity (linear or circular), and spillover losses in reflector systems, often achieving optimal aperture illumination with patterns tapered to about 10 dB at the reflector edges. Common types of antenna feeds include aperture-based designs like horns and patches for operation and high ; linear antennas such as dipoles or Yagi-Uda elements for simpler structures; and traveling-wave feeds like Vivaldi or rods for performance. Compound feeds, incorporating arrays, dichroic reflectors, or beam waveguides, enable multiband operation and frequency reuse in advanced systems. Feed horns, a prevalent subtype, often feature flared or corrugated waveguides to improve with free space and support dual polarizations via probes or septums. Modern designs increasingly leverage computational methods for and multiband capabilities, enhancing versatility across frequency ranges from VHF to millimeter waves.

Introduction and Fundamentals

Definition and Purpose

According to IEEE Std 145, an antenna feed is defined as follows: For continuous aperture antennas, the feed is the primary radiator; for example, a feeding a reflector. For antennas, that portion of the antenna system which functions to produce the excitation coefficients. This interface serves as the critical link between the transmitter or and the structure, converting guided waves into radiating waves or capturing incoming radiation for delivery. The primary purposes of an antenna feed are to transfer maximum power with minimal losses, ensure proper of the radiated signal, and maintain the integrity of the antenna's . By optimizing , the feed minimizes reflections and dissipation, supporting applications from communications to systems where signal fidelity is essential. It also influences characteristics, such as linear or circular, to match operational requirements and enhance compatibility with other system components. Historically, early antenna feeds in the , developed by , consisted of simple wire connections tied to grounded structures like wooden poles, allowing initial signal transmissions over distances up to 2.4 km. Modern feeds evolved throughout the alongside advancements in radio technology, incorporating more sophisticated designs to handle higher frequencies and power levels in diverse applications. Key performance metrics for antenna feeds include feed efficiency, which in reflector systems is the product of spillover efficiency (the fraction of feed power intercepted by the reflector) and taper efficiency (the uniformity of aperture illumination), contributing to the overall efficiency of the . Another important metric is at the feed interface, which quantifies the signal power reduction due to the feed's components and mismatches, directly impacting overall system performance. The role of in the feed is essential to minimize these losses and achieve high .

Basic Operating Principles

Antenna feeds operate based on the fundamental electromagnetic principles governed by , which describe the interaction of electric and in the propagation of radiofrequency (RF) energy. These equations unify the behavior of time-varying electric and , enabling the analysis of wave propagation in feeds as solutions to the wave equation derived from \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t} and \nabla \times \mathbf{H} = \mathbf{J} + \frac{\partial \mathbf{D}}{\partial t}. In feeds, RF energy primarily propagates in the transverse electromagnetic (TEM) mode, where both electric and are perpendicular to the direction of propagation, approximating plane waves in the far field. Higher-order modes, such as transverse electric () or transverse magnetic (TM), may arise in more complex structures like waveguides, where modes have no longitudinal electric field component and TM modes have no longitudinal magnetic field component. A key parameter in feed design is the Z_0, which represents the ratio of voltage to current for a wave propagating along an infinite and ensures maximum transfer without reflections. For a lossless , Z_0 is derived from the distributed L (henries per ) and capacitance C (farads per ) per of the line. The begins with the from Maxwell's : \frac{\partial V}{\partial z} = -L \frac{\partial I}{\partial t} and \frac{\partial I}{\partial z} = -C \frac{\partial V}{\partial t}, assuming sinusoidal steady-state operation (V(z,t) = \Re\{V(z)e^{j\omega t}\}). Differentiating and substituting yields the wave equation \frac{\partial^2 V}{\partial z^2} = LC \frac{\partial^2 V}{\partial t^2}, with solutions V(z) = V_0^+ e^{-\gamma z} + V_0^- e^{\gamma z} and \gamma = j\omega \sqrt{LC}. The is then Z_0 = \sqrt{\frac{L}{C}}, as the forward and backward waves satisfy V_0^+ / I_0^+ = \sqrt{L/C}. This intrinsic property depends solely on the line's and materials, independent of or frequency in the lossless case. For example, lines exhibit balanced TEM propagation with Z_0 determined by wire spacing and . Feed design must also consider polarization maintenance, where aligns the along a fixed , while requires two orthogonal components of equal magnitude with a 90° difference to produce a rotating . The feed preserves the desired by exciting appropriate distributions on the , preventing depolarization losses. In feeds, and TM modes influence ; modes support transverse to propagation, and TM modes can enable through mode combinations. Losses in feeds degrade efficiency and are categorized as dielectric losses (\alpha_d), due to dissipation in the insulating ; losses (\alpha_c), from finite in metals causing ohmic heating; and losses (\alpha_r), from unintended leakage. The total attenuation is \alpha_{\text{total}} = \alpha_d + \alpha_c + \alpha_r, where each is frequency-dependent and minimized through and geometry optimization.

Types of Feed Lines

Twin-Lead Transmission Lines

Twin-lead transmission lines consist of two parallel conductors, typically made of copper or wire, separated by an insulating material such as spacers or a continuous solid ribbon. This open-wire design allows the electromagnetic fields to propagate primarily in the air between the conductors, with the providing mechanical support and partial dielectric filling. Common implementations include flat ribbon for consumer applications and ladder line variants with periodic spacers for higher-power uses. The of lines typically ranges from 300 to 600 ohms, determined by the , spacing between the wires, and the properties of the insulating material. For instance, 300-ohm uses wires spaced approximately 0.3 inches apart with a solid , while 450- to 600-ohm ladder lines employ wider spacing and mostly air with spaced insulators to achieve higher impedance. Wider spacing increases the impedance by reducing the per unit length, making these lines suitable for matching balanced antennas like dipoles./07%3A_Transmission_Lines_Redux/7.01%3A_Parallel_Wire_Transmission_Line) These lines offer low at high-frequency () and very-high-frequency (VHF) bands, often transmitting over 98% of power over 100 feet at 3.5 MHz even with moderate ratios, due to their predominantly air minimizing losses. They are cost-effective and straightforward to construct or repair, using readily available materials, which makes them popular among hobbyists. However, twin-lead is susceptible to environmental , such as or dirt accumulation that increases losses, and it requires careful —crossing the conductors periodically—when routing around obstacles like metal structures to preserve and avoid common-mode currents. In applications, is commonly employed in setups to feed antennas, enabling efficient multiband operation when paired with a tuner, thanks to its balanced nature. Historically, 300-ohm was the standard for connecting rooftop antennas to receivers from the 1940s through the 1960s, before became prevalent with the rise of . The attenuation constant \alpha for twin-lead lines is approximated by the formula \alpha = \frac{R}{2 Z_0} + \frac{G Z_0}{2}, where R is the series per unit length (primarily from skin effect), G is the shunt conductance per unit length (due to losses), and Z_0 is the ; this expression holds for low-loss conditions, with the first term capturing and the second contributions, typically yielding values under 1 per 100 feet at frequencies. Installation must account for spacing effects, as inconsistent separation alters Z_0 and can introduce mismatches; the , which scales the signal propagation speed relative to free space, ranges from 0.8 for solid dielectrics to 0.95 for air-spaced designs, influencing length calculations./07%3A_Transmission_Lines_Redux/7.01%3A_Parallel_Wire_Transmission_Line)

Coaxial Cables

Coaxial cables are constructed with a central inner , typically made of solid or stranded or , surrounded by a that maintains precise spacing. This is enclosed by a braided or outer , which serves as the return , and often protected by an outer jacket for environmental durability. The of these cables is commonly 50 ohms for RF applications or 75 ohms for video and broadcast uses, determined by the of the inner diameter to the inner diameter of the outer and the constant of the . These cables offer significant advantages in antenna feed systems, including superior shielding that effectively blocks () from external sources, allowing reliable performance in noisy environments. Their flexibility facilitates easy installation and routing around obstacles, while supporting wide bandwidths suitable for multi-frequency operations. However, coaxial cables exhibit higher signal loss than open-wire lines at lower frequencies due to the and resistances, and their —ranging from approximately 0.66 for solid dielectrics to 0.80 for foam dielectrics—results in a propagation speed of 66% to 80% that of light in vacuum. In practical applications, cables serve as the standard feed line for modern high-frequency (), very high-frequency (VHF), and ultra high-frequency (UHF) antennas, as well as feeds and interconnects, due to their balanced performance in power transfer and . Representative types include , a option for short HF jumpers with moderate power handling; RG-213, suited for VHF and UHF runs requiring lower loss and higher durability; and RG-8, which supports up to 1 kW of continuous power in HF bands for high-power amateur and professional setups. Loss in coaxial cables, particularly from skin effect at higher frequencies, is a critical consideration for feed line efficiency. The skin effect confines RF current to a thin layer on the conductor surfaces, increasing effective resistance. The skin depth \delta is given by \delta = \sqrt{\frac{2}{\omega \mu \sigma}}, where \omega = 2\pi f is the angular frequency, \mu is the permeability (typically \mu_0 = 4\pi \times 10^{-7} H/m for non-magnetic conductors), \sigma is the conductivity (e.g., $5.8 \times 10^7 S/m for copper), and f is frequency in Hz. The resistance per unit length for the inner conductor is R'_{\text{inner}} = \frac{R_s}{2\pi a}, where surface resistance R_s = \sqrt{\frac{\omega \mu}{2\sigma}} = \frac{1}{\sigma \delta} and a is the inner conductor radius; for the outer conductor (assuming thickness much greater than \delta), R'_{\text{outer}} = \frac{R_s}{2\pi b} with b the inner radius of the shield. The total series resistance per unit length is R' = R'_{\text{inner}} + R'_{\text{outer}}. The attenuation constant in nepers per meter is then \alpha \approx \frac{R'}{2 Z_0}, where Z_0 = \frac{1}{2\pi} \sqrt{\frac{\mu}{\varepsilon}} \ln\left(\frac{b}{a}\right) is the characteristic impedance, \varepsilon = \varepsilon_r \varepsilon_0, \varepsilon_0 = 8.85 \times 10^{-12} F/m, and \varepsilon_r is the relative permittivity of the dielectric. To convert to decibels per 100 meters, \alpha_{\text{dB/100m}} = 868.6 \alpha (since $1 Np/m = 8.686 dB/m). For skin effect-dominated loss in typical 50-ohm cables, this simplifies to an approximate form \alpha \approx \left(\frac{36.3}{\sqrt{f}}\right) \frac{\sqrt{\varepsilon_r}}{d} dB/100m, where f is in MHz, \varepsilon_r is the dielectric constant, and d represents an effective diameter parameter (e.g., related to a or b/a ratio); this arises from substituting copper parameters and normalizing dimensions, emphasizing the \sqrt{f} dependence and inverse scaling with conductor size. Common connector interfaces for feeds include BNC types, which provide quick coupling for frequencies up to 4 GHz with low voltage standing wave ratio (VSWR) typically below 1.3:1, and N-type connectors, which offer threaded coupling for robust, weatherproof connections up to 11 GHz or higher, maintaining VSWR under 1.15:1 in precision versions to minimize reflections in high-power systems.

Waveguides

Waveguides are hollow metallic structures, typically rectangular or circular in cross-section, that guide electromagnetic waves at and millimeter-wave frequencies for efficient antenna feeding. These structures confine the waves through at the conducting walls, enabling propagation without radiation losses in the dominant modes. Rectangular waveguides are the most common for antenna applications due to their ease of manufacturing and mode control, while circular ones are preferred when or is needed. Operation occurs only above the , beyond which the supports propagating modes. For the dominant _{10} mode in a rectangular , the is given by f_c = \frac{c}{2a \sqrt{\mu \epsilon}} where c is the in , a is the width of the broader wall, and \mu and \epsilon are the permeability and of the medium inside the guide, respectively; for air-filled guides, this reduces to f_c = c / (2a). The corresponding is \lambda_c = 2a, ensuring single-mode operation when the operating satisfies \lambda_c / 2 < \lambda < \lambda_c. The propagation constant along the guide axis (z-direction) for this mode is the phase constant \beta = \sqrt{k^2 - k_c^2} where k = 2\pi / \lambda is the free-space wavenumber and k_c = \pi / a is the cutoff wavenumber, resulting in a guide wavelength \lambda_g = 2\pi / \beta > \lambda. These characteristics ensure controlled wave propagation, with the fields varying sinusoidally across the cross-section in the TE_{10} mode, where the electric field is uniform in the narrow dimension and half-sinusoidal in the wide one. Waveguides offer significant advantages for high-frequency, high-power antenna feeds, including very low attenuation losses in the GHz range—often below 0.1 dB/m for well-designed structures—and exceptional power handling capabilities exceeding kilowatts without dielectric breakdown, as there is no solid insulator. However, they are bulky, with dimensions scaling inversely with frequency (e.g., cross-sections on the order of centimeters for X-band), frequency-limited to bands above cutoff to avoid evanescent modes, and require specialized mode launchers such as probes or loops to excite the desired propagation mode from external sources. In circular waveguides, the dominant TE_{11} mode similarly has a cutoff f_c = c / (1.841 r \sqrt{\mu \epsilon}), where r is the radius, but offers lower attenuation at higher frequencies due to reduced current crowding on the walls. These properties make waveguides ideal for demanding applications such as antennas, where high-power pulses are fed to phased arrays; satellite dishes, for low-loss signal distribution to reflectors; and millimeter-wave systems in or , benefiting from minimal . A representative example is the WR-90 rectangular , standardized for X-band operation from 8.2 to 12.4 GHz with inner dimensions of 0.900 × 0.400 inches and a cutoff for TE_{10} at 6.557 GHz, widely used in military feeds for its balance of size and performance. To connect sections or with antennas, waveguides employ precision connections (e.g., UG-39/U for WR-90) to maintain and integrity, while linear tapers gradually vary the cross-section to transition between sizes, controlling impedance mismatches and minimizing reflections to below -30 dB.

Impedance Matching Techniques

Principles of Impedance Matching

Impedance matching in antenna feeds ensures maximum power transfer from the transmission line to the antenna, minimizing losses due to reflections. This principle is rooted in the , which states that the maximum average power is delivered to a load when its complex impedance Z_L is the complex conjugate of the source impedance Z_S^*. In the context of antenna feeds, the source is the Z_0 of the feed line, and the load is the antenna's ; achieving conjugate matching (Z_L = Z_0^*) results in zero and full power delivery. For real-valued impedances, which are common in many practical feed systems, this simplifies to Z_L = Z_0. The degree of mismatch is quantified by the \Gamma, defined as \Gamma = \frac{Z_L - Z_0}{Z_L + Z_0}, where |\Gamma| < 1 for passive loads. A value of \Gamma = 0 indicates perfect matching with no reflected power, while nonzero \Gamma leads to standing waves along the feed line. The voltage standing wave ratio (VSWR) measures the severity of this mismatch and is given by \text{VSWR} = \frac{1 + |\Gamma|}{1 - |\Gamma|}, with VSWR = 1 for ideal matching and increasing values indicating greater reflection; for example, VSWR = 2 corresponds to |\Gamma| = 0.333 and about 11% reflected power. The power loss due to mismatch is the fraction of incident power reflected, |\Gamma|^2, so the efficiency of power transfer to the antenna is $1 - |\Gamma|^2; mismatches with VSWR > 2 can reduce efficiency below 90%, significantly degrading system performance. These concepts tie directly to the Z_0 of the feed line, which depends on its geometry and materials as outlined in basic operating principles—for instance, cables commonly use 50 Ω to balance power handling and attenuation. VSWR is typically measured using slotted lines, where a slides along a longitudinal slot in the to detect voltage minima and maxima from standing waves, allowing calculation of \Gamma and impedance at frequencies. Antenna input impedances vary with due to changes in distribution and , necessitating matching for single-frequency applications (e.g., VSWR < 2 over <5% ) or broadband strategies for wider operational ranges (e.g., >20% ) to maintain low reflections across the band. The foundational theory of impedance matching in feeds emerged in the alongside advancements in theory by John R. Carson and Otto J. Zobel at Bell Laboratories, who developed analytical methods for wave propagation and filter that underpin modern systems.

Common Matching Methods

Lumped element matching employ discrete inductors (L) and capacitors (C) to achieve conjugate between an and its feed line by canceling the reactive component of the antenna impedance while transforming the real part to match the characteristic impedance of the line, typically 50 Ω. These are particularly useful at lower frequencies (e.g., ) where distributed elements become impractically long and for compact antennas with significant , such as short verticals or loaded dipoles. Common configurations include series or parallel LC circuits; for instance, a series can resonate out capacitive , while a shunt adjusts . The L-section matcher, a basic two-element network, provides a step-by-step process for matching a load impedance Z_L = R_L + jX_L to a source impedance Z_0 (e.g., Ω). First, determine the Q-factor as Q = \sqrt{\frac{Z_0}{R_L} - 1} for R_L < Z_0, assuming the reactive part is absorbed. Then, for a low-pass L-section (series L, shunt C), calculate the series reactance X_L = Q R_L and shunt susceptance B_C = Q / Z_0, yielding L = X_L / \omega and C = B_C / \omega, where \omega = 2\pi f. If X_L is negative (capacitive load), substitute a series capacitor with C = |X_L| / \omega. This design ensures maximum power transfer with minimal loss, though bandwidth is limited by Q (typically 5-10 for narrowband applications). For example, matching a 20 + j25 Ω antenna at 2 GHz to Ω requires a π-network extension for higher Q control. Distributed matching methods utilize transmission line sections to achieve impedance transformation without discrete components, ideal for higher frequencies (VHF/UHF) where lumped elements suffer parasitic effects. The quarter-wave transformer consists of a λ/4 line section with characteristic impedance Z_t = \sqrt{Z_0 Z_L}, transforming the load Z_L to Z_{in} = Z_0 at the design frequency, as derived from the input impedance formula Z_{in} = \frac{Z_t^2}{Z_L}. Design involves calculating the physical length l = \frac{\lambda}{4} = \frac{v}{4f} (v = velocity factor × c) and selecting Z_t via paralleled coaxial cables if needed; for a 200 Ω patch antenna to 50 Ω line, Z_t ≈ 100 Ω yields a narrowband match with ~10% bandwidth for VSWR < 1.5:1. Single- and double-stub tuning employs open- or short-circuited stubs along the feed line to introduce susceptance that cancels the load's reflection coefficient, calculated via the for position and length. For single-stub matching, plot the normalized load admittance on the , move along the constant |Γ| circle to intersect the g=1 conductance circle, then add a stub susceptance to reach the center (matched point); stub length is read as the distance to the short/open circle. Double-stub uses fixed spacing (e.g., λ/8) with adjustable stubs for broader tuning range, suitable for variable loads. These methods provide adjustable matching with VSWR < 1.2:1 over minor frequency shifts but require precise fabrication. Antenna-specific techniques adapt distributed elements to the geometry for direct feed integration. The gamma match, common for Yagi-Uda driven elements, uses a parallel rod (γ-line) connected via a capacitor to the antenna arm, transforming the typically low impedance (20-30 Ω) to 50 Ω while providing some balance. Design starts with rod length ~0.05λ and spacing 0.02λ, tuned by adjusting capacitance (e.g., 120 pF for 50 MHz Yagi) to null reactance, achieving SWR < 1.7:1 over 2:1 bandwidth. The delta match for dipoles employs a triangular conductor fanning from the feed point to a continuous element, matching 50-75 Ω balanced lines without splitting the dipole; angle and width are empirically set for resonance, often paired with a stub for fine adjustment. For folded dipoles, the hairpin match adds a short parallel conductor (λ/4 stub or coil) across the feed gap to raise impedance from ~100 Ω to 300 Ω, tuned by varying length (e.g., 24.5 cm for 30 m band) for 50 Ω match with a 4:1 transformer. Broadband techniques extend matching over wider frequencies by gradual impedance transitions. Tapered transmission lines, such as exponential or linear tapers, vary characteristic impedance smoothly over several wavelengths, reducing reflections for VSWR < 2:1 over 2:1 bandwidth; for example, a Klopfenstein taper matches 50 Ω to 200 Ω with minimal ripple. Ferrite cores on feed lines provide broadband transformation (e.g., 4:1 via stacked toroids) for multiband antennas, absorbing common-mode currents without narrowband resonance. Balun-integrated matches combine these for Yagi feeds, enhancing efficiency across HF bands. Measurement and tuning rely on VSWR meters to quantify match quality, where VSWR = (1 + |Γ|)/(1 - |Γ|) and Γ is the reflection coefficient; ideal is VSWR ≤ 1.5:1. Connect the meter inline, transmit low power, and adjust elements (e.g., stub length) to minimize reading across the band, often starting at band edges. Software tools like (Numerical Electromagnetics Code) simulate these via method-of-moments modeling of wire geometries, predicting impedance and SWR for iterative design; input antenna coordinates, run analysis, and optimize matching components for real-world validation.

Balanced and Unbalanced Feed Systems

Balanced Feeds

Balanced feeds in antenna systems involve transmission lines where the voltages on the two conductors are equal in magnitude but opposite in phase, ensuring no net current flows to ground. This configuration is particularly suited for symmetric antennas such as , where it prevents common-mode currents that could otherwise induce noise pickup or distort the radiation pattern. The absence of common-mode currents also minimizes interference from environmental factors, making balanced feeds common in high-frequency (HF) antenna applications for improved signal integrity. Implementation typically involves direct connection of the antenna to a twin-lead transmission line, such as ladder line, which maintains the balanced condition without introducing asymmetry. For center-fed dipoles, balanced lines such as 300-ohm ladder line are commonly used. However, the antenna's typical input impedance is around 73 ohms, so an antenna tuner is often required to achieve proper matching and preserve efficiency across the operating band. Challenges arise from environmental asymmetries, such as proximity to uneven ground or nearby conductive structures, which can introduce imbalances and allow common-mode currents to develop. The degree of balance is quantified by the common-mode rejection ratio (CMRR), which measures the system's ability to suppress common-mode signals relative to differential signals, with higher values indicating better performance. Examples include ladder line feeds for multi-band antennas, where the balanced line enables operation across multiple frequencies with low loss by connecting to a tuner at the station end. Off-center fed (OCF) dipoles also utilize balanced feeds, often with 4:1 impedance transformation to accommodate the offset feed point while supporting multi-band use. In terms of performance, balanced feeds exhibit lower radiation from the feed lines compared to unbalanced systems, as the equal and opposite currents on the conductors result in field cancellation along the line. This is described by the balanced mode voltage, given by V_{bal} = \frac{V_1 - V_2}{2}, where V_1 and V_2 are the voltages on the two conductors, emphasizing the differential nature that suppresses common-mode effects.

Unbalanced Feeds

Unbalanced feeds refer to antenna feed systems where one conductor is maintained at ground potential, providing a reference for the signal voltage, which simplifies integration with grounded equipment such as transmitters and receivers. This configuration is prevalent in the majority of modern antenna installations due to its compatibility with standard RF equipment and the inherent shielding properties of common feed lines like . In implementation, the signal voltage in an unbalanced feed is defined as V_{\text{unbal}} = V_{\text{signal}} relative to ground, where the outer conductor serves as the ground reference. Coaxial cables are typically connected directly to at the base or to via probe feeds, while end-fed designs may incorporate a to act as a ground plane substitute. A primary challenge in unbalanced feeds is the potential for common-mode currents to flow on the outer surface of the shield, which can generate radio frequency interference (RFI) by coupling noise into nearby electronics or the environment. These currents often necessitate additional measures such as ferrite chokes or physical isolation to suppress them and maintain system integrity. Practical examples include base-fed vertical monopoles, where coaxial cable delivers the signal to the antenna base over a ground plane or radials, and coaxial feeds for parabolic dish antennas, which connect to focal point probes or patches for microwave applications. These setups commonly employ 50-ohm impedance feeds to match standard coaxial cable characteristics and optimize power transfer. Performance trade-offs in unbalanced feeds include a higher likelihood of radiation from the feed line itself, which can distort the intended antenna radiation pattern by effectively extending the radiating structure. Additionally, the common-mode rejection ratio (CMRR) is inherently lower compared to balanced systems, leading to increased susceptibility to external interference and reduced pattern purity.

Supporting Feed Components

Baluns and Transformers

Baluns, or balanced-to-unbalanced transformers, are essential components in antenna feed systems that convert between balanced and unbalanced transmission lines while often providing impedance transformation to optimize power transfer and minimize losses. They prevent common-mode currents that can distort radiation patterns and cause interference, particularly when feeding balanced antennas like dipoles from unbalanced coaxial cables. Transformers within baluns enable voltage or current balancing, with designs tailored for specific frequency ranges and power levels in radio frequency applications. The concept of the modern balun was pioneered in the 1940s by Swiss engineer Gustav Guanella, who developed transmission line-based designs for impedance matching in radio-frequency circuits, initially for television antennas. Guanella's innovations, patented in 1949, laid the foundation for broadband baluns that maintain performance across wide frequency bands without resonant tuning. Baluns are broadly classified into two types: Guanella transmission line baluns, which operate as current baluns by forcing equal and opposite currents in the balanced output to suppress common-mode signals, and Ruthroff transformer baluns, which function as voltage baluns by enforcing equal voltages across the output terminals. Guanella designs use multiple parallel or series-connected transmission lines wound on a core to achieve balance, offering superior common-mode rejection and broader bandwidth compared to Ruthroff configurations, which rely on a single folded transmission line and are simpler but narrower in operational range. Current baluns like the Guanella type are preferred for antenna feeds due to their ability to maintain balance under varying load impedances, whereas voltage baluns like the Ruthroff are more susceptible to core saturation from unbalanced loads. For balance conversion without impedance change, 1:1 baluns are used, such as the coiled coaxial choke balun where the feedline is wound into a coil to create high impedance to common-mode currents on the outer shield. In contrast, transformers in baluns like 4:1 ratios step up impedance—for instance, matching a 50-ohm coaxial input to a 200-ohm balanced load—following the relation Z_{\text{out}} = n^2 Z_{\text{in}}, where n is the turns ratio. These are implemented by winding bifilar or trifilar wires on ferrite cores, with the turns ratio determining the transformation factor. A primary application of baluns is preventing common-mode currents on coaxial feeds to dipole antennas, where unbalanced coax can otherwise radiate or receive interference along its length, degrading antenna efficiency and pattern symmetry. Broadband versions enhance this by using ferrite cores such as Fair-Rite's FT-240-43 toroid, made from NiZn mix 43 material, which provides effective suppression from 25 MHz to 300 MHz for high-frequency applications. These cores support power handling up to 1 kW in balun designs, making them suitable for amateur radio and broadcast systems. Performance metrics for well-designed baluns include common-mode isolation exceeding 20 dB across the operating band, ensuring minimal leakage of unbalanced signals to the antenna. In practice, Guanella current baluns achieve higher isolation under load variations than Ruthroff voltage baluns, which may drop below 20 dB if the load becomes highly unbalanced.

Couplers and Attenuators

Directional couplers are essential components in antenna feed networks, enabling the sampling of a small fraction of the signal power to monitor forward and reflected waves without significantly disrupting the main transmission path. These devices typically couple a defined portion of the power, such as 10 dB coupling, which extracts approximately 10% of the incident power to a secondary port for analysis. The performance of a directional coupler is characterized by its directivity, defined as D = 20 \log \left( \frac{1}{|\Gamma|} \right), where |\Gamma| is the magnitude of the ratio between the reverse coupled signal and the forward coupled signal, quantifying the coupler's ability to distinguish between wave directions. High directivity, often exceeding 20 dB, ensures accurate separation of forward and reflected components, minimizing measurement errors in applications like standing wave ratio (SWR) assessment. Coaxial directional couplers, commonly featuring Type-N connectors, are widely used in antenna feeds operating at frequencies up to several GHz, offering robust construction for outdoor and high-power environments. Waveguide directional couplers, on the other hand, are preferred for microwave frequencies above 1 GHz, where they provide low-loss sampling in rectangular or circular waveguide structures integrated into the feed system. Both types facilitate SWR measurement by directing forward power to one port and reflected power to another, allowing real-time monitoring of impedance mismatches in the antenna feed. Attenuators serve to control power levels within antenna feed networks by intentionally reducing signal amplitude, with attenuation quantified as A = 20 \log \left( \frac{V_{\text{in}}}{V_{\text{out}}} \right), where V_{\text{out}} and V_{\text{in}} are the output and input voltages, respectively (standard positive value). Fixed attenuators commonly employ π or T network configurations using resistors matched to the characteristic impedance Z_0, typically 50 Ω in RF systems; for a symmetrical π-pad attenuator, the series resistors are given by R_1 = R_2 = Z_0 \frac{K + 1}{K - 1} and the shunt resistor by R_3 = Z_0 \frac{2K}{K^2 - 1}, where K = 10^{A/20} > 1. Variable attenuators, adjustable via mechanical, electronic, or mechanisms, allow dynamic control over attenuation levels from 0 to over 30 , enabling fine-tuning in response to varying signal conditions. In antenna feeds, attenuators are placed in-line to protect sensitive receivers from excessive transmitted during testing or , preventing overload and damage. They also support of feed systems by providing known signal reductions for accurate referencing and alignment procedures. However, the inherent of attenuators, which exceeds the intended due to resistive , contributes to overall feed inefficiency, potentially reducing transmitted by 0.5 or more depending on the design.

References

  1. [1]
    [PDF] IEEE Standard Definitions of Terms for Antennas
    Jun 1, 2012 · A linear array antenna feed with radiating element locations and excitations chosen to correct for aberrations present in the focal region ...<|control11|><|separator|>
  2. [2]
    Antenna Feeds - Bird - Major Reference Works - Wiley Online Library
    Apr 15, 2005 · Antenna feeds supply energy to secondary antennas like reflectors, lenses, or beam waveguides, via a transmission line or waveguide.Missing: definition | Show results with:definition
  3. [3]
    [PDF] Radio Antennas, Feed Horns, and Front-End Receivers
    A radio antenna converts electromagnetic radiation to electric current. Transmitting antennas are characterized by gain, while receiving antennas by effective ...
  4. [4]
  5. [5]
    Antenna Feeds - an overview | ScienceDirect Topics
    Antenna feed refers to the mechanism that couples energy into the antenna, enabling the propagation of electromagnetic waves.
  6. [6]
  7. [7]
    Guglielmo Marconi - Radartutorial.eu
    In 1895 he developed a grounded transmitter antenna that could transmit signals to a point 2400 meters away. In his experiments he used a wire tied to a wooden ...
  8. [8]
    Antenna History: Trends in Antenna Design - JEM Engineering
    Key antenna trends include Yagi-Uda (1920s), horn (1930s), dish (1960s), microstrip (1970s), and PIFA (1980s) antennas.Missing: paper | Show results with:paper
  9. [9]
    Antenna Efficiency
    A high efficiency antenna has most of the power present at the antenna's input radiated away. A low efficiency antenna has most of the power absorbed as losses ...
  10. [10]
    Understanding Cable and Antenna Analysis | Anritsu America
    The antenna itself has a return loss of 15 dB but the 5 dB insertion loss improves the perceived system return loss by 10 dB (5 dB *2).
  11. [11]
    The Feynman Lectures on Physics Vol. II Ch. 24: Waveguides
    ... our filter equations—is given by z0=√L0C0,. and is, therefore, a pure resistance. To find the propagation speed v and the characteristic impedance z ...
  12. [12]
    [PDF] Dielectric Permittivity Measurements of ... - Auburn University
    αtot = αc + αd + αr. (2.10). 13. Page 23. Radiation loss can be minimized by avoiding sharp angles or discontinuties in the microstrip transmission line, and ...
  13. [13]
    [PDF] Chapter 24 - Transmission Lines - QSL.net
    Transmission lines carry RF power efficiently. Types include open-wire (two-wire) and coaxial (concentric) lines. Coaxial lines have a shielded outer conductor.
  14. [14]
    What is “twin-lead” transmission line — and what happened to it?
    Jun 19, 2024 · Twin-lead is a balanced, ungrounded transmission line with two parallel conductors, often with a low-loss dielectric for insulation.Missing: disadvantages | Show results with:disadvantages
  15. [15]
    [PDF] Let's Talk Transmission Lines - ARRL
    Twinlead is often used for home TV antennas and has a characteristic impedance of 300 Ω. So-called “window” ladder line is commonly used for amateur work and ...Missing: structure disadvantages
  16. [16]
    [PDF] Transmission Lines
    Oct 15, 2018 · Distributed Circuit Model of a Transmission Line. 413. Z = R + jωL. Y = G + jωC. (10.6.1). This leads to a so-called distributed-parameter ...<|control11|><|separator|>
  17. [17]
  18. [18]
    Coax (Coaxial) Cables, An Introduction - Elliott Sound Products
    The characteristic impedance of a connector is determined by the dimensions of the inner and outer conductors, as well as the type of dielectric used to support ...
  19. [19]
    Understanding Coaxial Cable Terms, Specifications and Applications
    Jun 18, 2015 · Coaxial cable is a two conductor electrical cable consisting of a center conductor and an outer conductor with an insulating spacer between the two.
  20. [20]
    Coaxial Cable Impedance Calculator - everything RF
    Most RF Coaxial Cables have an impedance of 50 ohms or 75 ohms. These are considered to be standardized impedance values with cables readily available.Missing: structure | Show results with:structure
  21. [21]
    Understanding Low Loss Coaxial Cables and Their Applications
    The outer conductor acts as the return path for the inner conductor and carries current in the opposite direction. This, in turn, functions as a shield because ...Missing: structure | Show results with:structure
  22. [22]
    Coax vs. Twinlead, July 1965 Electronics World - RF Cafe
    Mar 21, 2024 · Twin-lead has lower loss but is vulnerable to interference, while coax has higher loss but is less sensitive to routing and external ...Missing: disadvantages EMI
  23. [23]
    coax versus open-wire; velocity factor; electrical length; coaxial ...
    Mar 12, 2020 · In general, coaxial cable transmission lines with a foam dielectric have a higher velocity factor than coaxial cables with a solid dielectric.
  24. [24]
    What coax should be used for HF, VHF, or UHF frequencies?
    RG-58 provides a good balance between price and performance for HF Frequencies and VHF UHF derives fewer losses RG-213 or LMR-400 LMR400.Missing: satellite feeds 8 power
  25. [25]
    MPD Digital 50ft RG8x Coax UHF (PL259) Male/Male 50 ohm ...
    RG-58 is better for short jumpers and work in the HF Bands. For UHF, VHF and CB Radio RG-8x and RG-213 supports better signal retention. Our coax cables are US- ...
  26. [26]
    Is RG-8X the General Purpose Coaxial Cable? - AmateurRadio.com
    Apr 29, 2023 · These power levels are easily handled by RG-8X. In fact, the cable is spec'd for over 1 kW for all of the HF bands. At 50 MHz, the power rating ...
  27. [27]
  28. [28]
    Coax Loss Calculations - Microwave Encyclopedia
    On this page we will review the simple math for calculating the RF losses of coaxial transmission lines over frequency.Missing: 100m | Show results with:100m
  29. [29]
  30. [30]
    6 most widely used RF cable connectors - DOLPH MICROWAVE
    Apr 19, 2024 · ​​Key fact​​: A quality N connector maintains ​​<1.15:1 VSWR up to 18 GHz​​ in precision variants (like 7/16″ N), outperforming SMA and BNC for ...
  31. [31]
    Microwaves101 | Waveguide Primer - Microwave Encyclopedia
    A final disadvantage of waveguide is that you can't pass DC currents along with your RF signal. Fear not, the advantages of power handling often outweigh all of ...
  32. [32]
    Waveguide Mathematics - Microwave Encyclopedia
    The accepted limits of operation for rectangular waveguide are (approximately) between 125% and 189% of the lower cutoff frequency. Thus for WR-90, the cutoff ...
  33. [33]
    [PDF] 6.013 Lecture 19: Waveguides and Applications - DSpace@MIT
    Oct 28, 2002 · The frequencies at which a waveguide can propagate can be represented in a simple diagram such as that presented in Figure 19-4 for TEm0 modes.
  34. [34]
    Microwaves101 | Circular Waveguide - Microwave Encyclopedia
    Standard sizes of circular waveguide. Visit our page on rectangular waveguide dimensions, and download "Uli's List". Circular waveguide internal cross sections ...
  35. [35]
    WR90 | WG16 | R100 - Rectangular Waveguide Size - everything RF
    WR90 Specifications · Recommended Frequency Band: 8.20 to 12.40 GHz · Cutoff Frequency of Lowest Order Mode: 6.557 GHz · Cutoff Frequency of Upper Mode: 13.114 GHz.
  36. [36]
    Waveguide Transitions | Eravant, Formerly SAGE Millimeter
    Waveguide transitions from Eravant. We are a US based manufacturer of waveguide transitions such as circular mode, tapered, and square.
  37. [37]
    [PDF] Antenna Impedance Matching – Simplified - Abracon
    Impedance matching is designing an antenna's input impedance to match the RF circuitry's output impedance, often 50 Ω, for maximum efficiency.
  38. [38]
    [PDF] Impedance Matching - UCSB ECE
    Oct 29, 1999 · ... impedance ratio and maximum SWR, solve for at each band edge in terms of the band-edge reflection coefficient for that SWR. As an example ...
  39. [39]
    Slotted Line Measurements - Microwaves101
    Slotted lines measure VSWR by measuring peak and valley voltages, and frequency by measuring half-wavelength distance. They also measure impedance.
  40. [40]
    [PDF] The Bell System Technical Journal January, 1928
    THE theory of electrical transmission along straight parallel guiding conductors is of fundamental importance to the communication engineer. In Its original, ...
  41. [41]
    ARRL Antenna Book
    This book contains everything you need to understand how radio signals propagate, how antennas work, and how to construct your own antenna system. Use The ARRL ...Missing: common matching<|separator|>
  42. [42]
    [PDF] Basic Amateur Radio - Simple Gain Antennas for the Beginner - ARRL
    An alternative matching method is shown at B of Fig. 5. This involves the use of a toroidal broadband 2:1 matching transformer. Recent editions of The. ARRL ...
  43. [43]
  44. [44]
    EZNEC Antenna Software by W7EL
    View Antenna display, showing the "wires" making up the model of the five-element beam, with currents and 2D slice superimposed to show orientation.NEC-5 · EZNEC Pro+ v. 7.0 Updates · AutoEZ · EZNEC Pro+ Printable Manual
  45. [45]
    [PDF] A practical approach to HF and VHF antennas, plus antenna ... - ARRL
    Properly balanced coaxial or open wire feed lines with balanced antennas should contribute minimal RF at the operating position, even absent a shack ground ...Missing: definition | Show results with:definition
  46. [46]
  47. [47]
  48. [48]
    Balun Basics Primer - High Frequency Electronics
    The amount of attenuation this signal will experience from the balanced to unbalanced port is called common mode rejection ratio (CMRR) and is expressed in dB.
  49. [49]
    Random Length Multiband Dipoles - ARRL
    Feed the dipole in the center with 450-ohm ladder line (available from most ham dealers), and buy an antenna tuner with a balanced output. Feed the ladder line ...
  50. [50]
    Fundamentals of Off-Center-Fed Dipoles - Antentop
    Off-center-fed dipoles (OCFs) are 1/2-wavelength wire antennas fed at neither the center nor the end, but somewhere between--off-center.
  51. [51]
    [PDF] MEASUREMENT OF BALANCED ANTENNAS USING ... - URSI
    If the antenna is perfectly balanced, V1 = −V2, resulting Vd = 2V1 and Vc = 0. A new measurement method for balanced antennas based on mixed-mode network ...Missing: V_bal = | Show results with:V_bal =
  52. [52]
    Microwaves101 | Baluns - Microwave Encyclopedia
    The unbalanced, or "single-ended" signal can be microstrip or stripline; the key characteristic is that one of the two conductors is clearly grounded. In a ...
  53. [53]
    2019 No Nonsense General Class Study Guide: Antennas and Feed ...
    Feb 19, 2019 · TV twinlead has a characteristic impedance of 300 ohms. Ladder line or window line has a characteristic impedance of 450 ohms. Open wire ...
  54. [54]
    Understanding RF Antenna Baluns | Advanced PCB Design Blog
    Sep 29, 2025 · If a balanced antenna is fed directly with a coaxial feeder, the outer shield of the coax can radiate and pick up radio frequency signals. This ...
  55. [55]
    Common mode current - W8JI
    Common mode currents cause conductors to couple to each other. Common mode currents are generally behind desired EM radiation and reception.Missing: monopole challenges
  56. [56]
    Understanding the Guanella Transmission Line Balun
    Feb 4, 2024 · We'll start by examining two basic 1:4 balun configurations that can be used to feed a symmetric dipole antenna from an unbalanced source.Missing: invention 1940s
  57. [57]
    All About the 1:1 Current/Choke Balun
    The 1:1 Guanella-balun is the basic building block for other types ... Guanella or Ruthroff design; strong arguments can be made for each design approach.
  58. [58]
    Current vs. Voltage Baluns - Marki Microwave
    Oct 22, 2014 · Current baluns force equal and opposite currents, while voltage baluns force equal and opposite voltages at the output.
  59. [59]
    A Brief Introduction to Ruthroff Transmission Line Transformers
    Apr 7, 2024 · Compared to the Guanella transformers, the Ruthroff configurations provide a relatively lower bandwidth. ... Understanding the Guanella ...
  60. [60]
    Toroids (5943003801) - Fair-Rite
    Part Number: 5943003801 43 TOROID 43 Material® Inductive Components Closed Magnetic Circuit Printer Friendly Data Sheet Check Stock Chart
  61. [61]
    43 Material® Data Sheet - Fair-Rite
    This NiZn is our most popular ferrite for suppression of conducted EMI from 20 MHz to 250 MHz. This material is also used for inductive applications.
  62. [62]
    The broadband common mode choke, 1:1 balun and 1:1 unun - PA9X
    Coax is an unbalanced feed line. So when you feed a balanced antenna with ... Power ratings per toroid with most common winding methods within an acceptable SWR ...