Fact-checked by Grok 2 weeks ago

Antenna types

Antennas are specialized transducers that convert guided electromagnetic waves into radiated waves for transmission or vice versa for reception, enabling wireless communication, broadcasting, radar, and sensing applications across radio frequencies. They are broadly classified by physical structure into categories such as wire antennas (e.g., dipoles and monopoles), aperture antennas (e.g., horns and reflectors), microstrip or patch antennas, and array antennas, each designed to achieve specific radiation patterns, bandwidths, and efficiencies suited to operational requirements. This diversity arises from the need to optimize performance metrics like directivity, gain, and impedance matching in varying environments, from mobile devices to satellite systems. Key classifications also consider radiation patterns, distinguishing between isotropic (ideal uniform radiation, theoretical reference), omnidirectional (uniform in one plane, e.g., dipole antennas with figure-eight patterns), and directional types (focused beams for higher gain, e.g., Yagi-Uda or helical antennas). Feeding mechanisms further categorize antennas as balanced (symmetric feeds like dipoles requiring baluns) or unbalanced (asymmetric, e.g., monopoles over ground planes), influencing integration with transmission lines. Common examples include the half-wave dipole, a foundational wire antenna with linear polarization and narrow bandwidth; the rectangular patch, a low-profile microstrip type for compact devices like smartphones; and horn antennas, aperture-based for broadband, high-power radar use. Array configurations, such as phased arrays, enable beam steering and enhanced directivity by combining multiple elements, critical for modern 5G and satellite communications. These types continue to evolve with advancements in materials and frequencies, incorporating features like polarization diversity (linear, circular) to mitigate signal fading and support applications from VHF broadcasting to millimeter-wave sensing. Selection depends on factors including frequency range, size constraints, and environmental robustness, ensuring optimal electromagnetic coupling in practical deployments.

Classification of Antennas

By Physical Configuration

Antennas are classified by their physical configuration into categories such as wire, aperture, and planar types, providing a structural taxonomy that emphasizes form over operational principles or radiation characteristics. This approach highlights how the geometric arrangement of conductive elements influences radiation and reception properties, serving as a foundational framework for antenna design. Wire antennas consist of one or more thin, straight, or curved conducting wires that support current distributions to radiate or receive electromagnetic waves. Examples include linear dipoles, where two collinear wire segments form a resonant structure, and loop antennas formed by closed wire paths. Aperture antennas, in contrast, feature openings or slots in a larger conducting surface, such as waveguides or ground planes, through which fields propagate to create radiation. Representative examples are horn antennas, which flare out from a waveguide to form a radiating aperture, and slot antennas cut into planar conductors. Planar antennas are fabricated on flat dielectric substrates with conductive patterns etched or deposited on the surface, enabling conformal and integrated designs. Common instances include microstrip patch antennas, where a rectangular or circular metallic patch over a ground plane resonates to produce radiation. The evolution of antenna configurations began with wire designs in the late 19th century, pioneered by experiments demonstrating electromagnetic wave propagation using simple straight and looped wires. These early forms laid the groundwork for wireless communication, with refinements like the dipole emerging by the 1880s. Aperture antennas gained prominence in the 20th century for microwave applications, building on waveguide technology developed in the 1930s. Planar antennas emerged conceptually in the 1950s but became practical in the 1970s, facilitated by advancements in printed circuit board (PCB) manufacturing that allowed low-profile, mass-producible structures. Wire antennas offer advantages in simplicity of construction and low cost, making them suitable for broadband or low-frequency applications, though they often require larger sizes relative to wavelength and exhibit moderate gain. Aperture antennas excel in achieving high gain through larger effective radiating areas, ideal for directive microwave links, but they can be bulkier and more complex to fabricate at lower frequencies. Planar antennas provide superior integration into compact devices like mobile handsets and arrays, with low profile and ease of arraying, yet they typically suffer from narrower bandwidths and lower efficiency due to substrate losses. Key performance metrics for these configurations include the effective length for wire antennas and aperture efficiency for aperture and planar types. The effective length l_e of a wire antenna quantifies its receiving capability, defined such that the open-circuit voltage V_{oc} = l_e E_i, where E_i is the incident electric field magnitude; for a short dipole of physical length l, l_e = l/2. For aperture antennas, the effective aperture A_e relates to the antenna's ability to capture power, given by A_e = \frac{\lambda^2 G}{4\pi}, where \lambda is the wavelength and G is the antenna gain, with aperture efficiency \eta_a = A_e / A_p comparing A_e to the physical aperture area A_p. These parameters underscore the trade-offs in structural design for optimizing radiation efficiency.

By Operating Principle

Antennas can be classified by their operating principle, which describes the fundamental electromagnetic mechanisms governing their radiation and resonance behavior. This categorization emphasizes how energy is stored, propagated, and radiated, distinguishing between structures that rely on standing waves, progressive waves, or aperture openings. Such principles trace back to early experiments and theoretical developments that established the foundations of antenna theory. Resonant antennas operate at discrete frequencies where their physical dimensions are fractions of the operating wavelength, leading to standing wave patterns along the structure that efficiently couple energy to free space. The concept originated with Heinrich Hertz's 1888 experiments, in which he demonstrated the generation and detection of electromagnetic waves using resonant dipole-like structures excited by spark gaps, confirming Maxwell's predictions of wave propagation. A classic example is the half-wave dipole, where the antenna length equals half the wavelength, resulting in an input impedance of approximately 73 Ω at resonance, which facilitates matching to transmission lines without significant reactive components. This resonance condition maximizes radiation efficiency but limits bandwidth, as detuning from the resonant frequency introduces high reactance. Traveling wave antennas, in contrast, support progressive waves that travel along the antenna structure without reflection, enabling broadband operation and directional radiation. Theoretical advancements in the late 1930s by W. W. Hansen and J. R. Woodyard introduced principles for enhancing directivity in such antennas by optimizing phase progression, particularly in end-fire configurations where the wave velocity is controlled to align with the desired beam direction. The helical antenna exemplifies this principle, where a coiled wire guides a traveling wave with phase velocity adjusted via geometry—typically near the speed of light for axial-mode operation—to produce circular polarization, making it suitable for satellite communications. Aperture antennas radiate electromagnetic energy through an opening in a conducting screen or waveguide, treating the aperture as a source of equivalent currents that excite the surrounding space. The field equivalence principle, formalized by S. A. Schelkunoff in the 1930s, states that the radiation from an aperture is identical to that produced by surface electric and magnetic currents distributed along the aperture edges, allowing complex structures to be analyzed as simplified current sheets. This approach underpins the design of horn and reflector antennas, where the aperture illumination determines gain and pattern, often achieving high directivity through focused wavefronts. Leaky-wave antennas operate by progressively leaking energy from a guided wave structure, such as a perturbed waveguide, into space along its length, resulting in a scanned beam whose angle depends on frequency. Early developments trace to W. W. Hansen's 1940 work on slotted waveguides, which initiated controlled leakage for radar applications. The propagation is characterized by a complex wavenumber given by k = \beta + j\alpha, where \beta is the phase constant determining the beam direction via \sin\theta = \beta / k_0 (with k_0 the free-space wavenumber), and \alpha is the attenuation constant controlling the leakage rate and aperture efficiency. This dispersion relation enables frequency-scanning capabilities, though it requires careful design to balance leakage and minimize end reflections.

By Radiation Pattern and Polarization

Antennas can be classified based on their radiation patterns, which describe the spatial distribution of radiated power, and their polarization, which refers to the orientation of the electric field vector in the radiated wave. Omnidirectional antennas produce a radiation pattern that radiates power equally in all directions within a plane, typically the azimuthal plane, making them suitable for applications requiring uniform coverage such as mobile communications or broadcasting. For instance, a vertical monopole antenna exhibits an omnidirectional pattern in the horizontal plane, with nearly uniform radiation around its axis. In contrast, directional antennas concentrate radiated energy into specific directions to achieve higher gain and range, often at the expense of coverage in other directions; examples include endfire arrays, where the main beam points along the array axis. This focusing enhances signal strength in targeted areas, such as point-to-point links. Polarization classification further refines antenna selection, encompassing linear, circular, and elliptical types, each defined by the locus traced by the electric field vector over time. Linear polarization occurs when the electric field oscillates along a fixed axis, either horizontal or vertical, as seen in dipole antennas that produce vertical linear polarization when oriented accordingly. Circular polarization features the electric field rotating in a circle, with left-hand (LHCP) or right-hand (RHCP) senses depending on the rotation direction relative to propagation, while elliptical polarization is a general case combining linear and circular components with varying axial ratios. Proper matching of transmit and receive antenna polarizations is critical to minimize losses; mismatch can result in polarization loss factors up to 20 dB or more in near-orthogonal cases, severely degrading signal quality in wireless systems. The isotropic radiator serves as a theoretical benchmark for radiation patterns, representing a hypothetical point source that radiates uniformly in all directions over a sphere, yielding a directivity and gain of 0 dBi. Practical antennas approximate this only closely in limited scenarios, such as small loops or short dipoles in their equatorial planes, but real-world implementations face efficiency and pattern imperfections that prevent exact uniformity. Directivity D quantifies pattern focusing and relates to the beam solid angle \Omega_A—the solid angle over which the normalized power pattern integrates—as D = 4\pi / \Omega_A, where \Omega_A = \iint (U(\theta, \phi)/U_{\max}) \sin\theta \, d\theta \, d\phi and U is the radiation intensity. Early advancements in polarization, such as Guglielmo Marconi's use of vertical linear polarization in transatlantic experiments around 1901, enabled reliable long-distance propagation by aligning with ground-wave and ionospheric reflection characteristics. In modern contexts, 5G systems leverage dual-polarization MIMO configurations to double spectral efficiency by simultaneously using orthogonal polarizations, enhancing capacity in dense urban environments.

Wire and Simple Antennas

Dipole Antennas

Dipole antennas are fundamental resonant wire structures consisting of two conductive elements extending from a central feed point, typically arranged collinearly and symmetrically to support balanced currents. They operate by exciting electromagnetic waves along their length, radiating energy efficiently when the total length approximates an odd multiple of half-wavelengths. The simplest and most common form is the half-wave dipole, which serves as a building block for more complex antenna systems due to its straightforward design and predictable performance characteristics. The basic half-wave dipole has a total length l = \lambda/2, where \lambda is the operating wavelength, making it resonant at the design frequency. Its radiation pattern exhibits a figure-8 shape in the E-plane (the plane containing the electric field and the direction of maximum radiation), with nulls along the axis of the dipole, while being nearly omnidirectional in the H-plane (the plane containing the magnetic field and the direction of maximum radiation). The input impedance is approximately $73 + j42.5 \, \Omega, with the real part representing radiation resistance and the imaginary part arising from capacitive reactance at exact half-wavelength. This configuration provides a gain of about 2.15 dBi and is widely used for its simplicity and moderate efficiency. For lengths much shorter than the wavelength (l \ll \lambda), the short dipole approximation applies, treating the structure as a Hertzian dipole with nearly uniform or triangular current distribution. The radiation resistance is given by R_{\rm rad} = 20 \pi^2 (l/\lambda)^2, which becomes very small, leading to low efficiency as ohmic losses dominate over radiated power. Such antennas require high input voltages to achieve meaningful radiation and are less practical for most applications without additional matching or loading techniques. To improve impedance matching, particularly for connection to 300-Ω transmission lines, the folded dipole variant incorporates two parallel half-wave elements connected at the ends, with the feed applied between them. This design increases the input impedance to approximately 300 Ω and enhances bandwidth to about 20%, compared to the narrower bandwidth of a standard dipole, due to the effective thickening of the structure and reduced Q factor. The folded configuration maintains a similar radiation pattern to the simple dipole but offers better tolerance to environmental variations. Feeding a dipole with an unbalanced line, such as coaxial cable, often excites common-mode currents on the outer conductor, distorting the radiation pattern and causing feedline losses. A balun (balanced-to-unbalanced transformer) is essential to suppress these currents, ensuring balanced operation and symmetric radiation by isolating the feedline from the antenna. Common balun types include choke designs or transformer-based structures tailored to the dipole's impedance. Dipole antennas trace their practical origins to demonstrations by Oliver Lodge in 1888, who used wire dipoles to transmit and receive electromagnetic waves, confirming Hertz's predictions and advancing early radio technology. During World War II, dipoles were integral to radar systems, forming arrays for directional sensing in applications like airborne detection and fire control, leveraging their reliable patterns for signal transmission and reception. Design variability, particularly the ratio of element diameter to wavelength (d/\lambda), significantly affects performance; thinner elements (d/\lambda \ll 1) yield higher Q factors and narrower bandwidths, while thicker elements reduce Q inversely proportional to d/\lambda, broadening bandwidth and improving impedance stability. This trade-off allows optimization for specific applications, such as broadband communication versus narrowband efficiency.

Monopole Antennas

A monopole antenna consists of a straight conductive element, typically a rod or wire, mounted perpendicular to a conducting ground plane, serving as an efficient vertical radiator that produces vertically polarized electromagnetic waves. This configuration leverages the ground plane to enhance radiation efficiency, making it suitable for applications requiring omnidirectional coverage in the horizontal plane above the ground. The antenna's simplicity and effectiveness stem from its ability to approximate the performance of more complex structures while occupying minimal space. The canonical form is the quarter-wave monopole, where the physical height h equals one-quarter of the operating wavelength \lambda/4, resulting in a nearly hemispherical radiation pattern confined to the space above the ground plane with omnidirectional characteristics in azimuth. Its input impedance is approximately 36.5 \Omega, roughly half that of a full-wave dipole due to the ground plane's reflective properties. According to image theory, the ground plane mirrors the monopole's current distribution, forming a virtual image that behaves as a complete dipole of height \lambda/2, ensuring no radiation penetrates below the plane and directing all power into the upper hemisphere for improved directivity over open-space equivalents. To accommodate space constraints at lower frequencies where a full \lambda/4 height becomes impractical, loading techniques employ inductive elements such as coils at the base or top of the monopole to shorten the physical length while maintaining resonance. These coils introduce additional inductance that electrically extends the antenna's effective length, allowing operation at desired frequencies; for a thin-wire approximation, the self-inductance can be estimated as L \approx \frac{\mu_0 h}{2\pi} \ln\left(\frac{2h}{a}\right), where a is the wire radius, facilitating tuning for compact designs. Whip antennas represent a flexible variant of monopoles, utilizing bendable rods or wires for rugged, portable deployment in dynamic environments. Monopole antennas gained early prominence in automotive radio systems during the 1910s, enabling mobile reception amid the rise of wireless broadcasting, and remain integral to contemporary vehicle communications for AM/FM, GPS, and cellular services as well as base stations in wireless networks. Despite their advantages, quarter-wave monopoles exhibit inherent bandwidth limitations, typically spanning 5-10% of the center frequency for a voltage standing wave ratio below 2:1, often necessitating traps or matching networks to broaden operational range. The monopole's performance above the ground plane mirrors that of a dipole in free space, providing a balanced reference for comparative analysis.

Loop Antennas

Loop antennas consist of a continuous closed loop of conductive wire, typically circular or square in shape, designed to radiate or receive electromagnetic waves through magnetic field coupling. Unlike open-wire antennas such as dipoles, which primarily interact with the electric field, loop antennas excel in scenarios with strong electric interference, such as near metallic objects, due to their insensitivity to capacitive coupling. They find applications across radio frequency bands, from low-frequency (LF) to very high-frequency (VHF), valued for their simplicity, compactness, and directional properties. In general theory, electrically small loops—where the loop circumference is much less than the operating wavelength—function as equivalent magnetic dipoles, with the effective dipole moment oriented perpendicular to the loop's plane. The far-field azimuthal electric field component for such a small loop is expressed as E_\phi = \frac{\mu_0 I_0 k^2 A \sin\theta \, e^{-j k r}}{4\pi r}, where \mu_0 is the permeability of free space, I_0 is the peak current, k = 2\pi/\lambda is the wave number, A is the loop area, \theta is the polar angle from the loop axis, and r is the radial distance in the far field. For electrically large loops, where the circumference approaches the wavelength, the structure behaves analogously to an electric dipole, exhibiting radiation patterns with maxima along the loop axis. The polarization characteristics of loop antennas depend on their size and orientation: small loops, when mounted horizontally, generate a horizontally polarized electric field in the far field, while large loops, typically oriented in a vertical plane, produce a vertically polarized field. This circular or elliptical polarization potential arises from the loop's geometry, enabling effective use in systems requiring specific polarization matching, though linear polarization dominates in standard configurations. Small loop antennas exhibit a high quality factor (Q), often exceeding 100, which confines their bandwidth to narrow ranges—typically less than 1% of the center frequency—making them suitable for resonant, narrowband operations but requiring tuning circuits for broader use. In high-frequency propagation through the ionosphere, Faraday rotation—caused by the magneto-ionic medium—alters the polarization plane, degrading loop antenna performance by introducing fading; this effect is proportional to the square of the frequency and the total electron content along the path. Early theoretical foundations for loop antennas emerged in the 1890s through experiments by physicist George Francis FitzGerald, who in 1883 derived the radiated power from a small loop, demonstrating its viability as a radiator and influencing subsequent wireless developments. In practical applications, loop antennas have been integral to AM radio broadcasting since the early 20th century, primarily as compact receiving elements in broadcast receivers, leveraging their high directivity and noise rejection to enhance signal selectivity in the medium-frequency band (0.5–1.6 MHz).

Aperture and Reflector Antennas

Horn Antennas

Horn antennas are a type of aperture antenna consisting of a flared waveguide section that transitions from a waveguide feed to a larger opening, enabling efficient radiation of electromagnetic waves in the microwave and millimeter-wave frequency bands. This design directs energy through the aperture to produce a directive beam with relatively high gain, making them suitable for high-frequency applications where precise control over the radiation pattern is essential. The flare shape expands the wavefront gradually, minimizing reflections and optimizing impedance matching between the waveguide and free space. Common types of horn antennas include sectoral horns, which flare in only one plane—either the E-plane (electric field direction) for E-plane sectoral horns or the H-plane (magnetic field direction) for H-plane sectoral horns—and pyramidal horns, which flare in both planes to form a rectangular aperture. Conical horns feature a circular cross-section that expands symmetrically, offering rotational symmetry useful for certain polarization needs. The choice of type depends on the desired beam shape and polarization characteristics, with pyramidal and conical designs being prevalent for their balanced performance. The design of horn antennas emphasizes minimizing phase errors across the aperture to achieve uniform wavefronts and maximum gain. Phase error arises from path length differences between the waveguide throat and various points on the aperture plane, which can distort the radiation pattern if the flare angle is too large. Optimal flare angles are selected to limit this error, ensuring near-uniform phase distribution; for instance, the flare is dimensioned such that the maximum phase deviation is kept below approximately 90 degrees for efficient operation. This approach ties into broader aperture theory principles, where the antenna's performance is analyzed as an equivalent radiating aperture. The resulting directivity leads to a gain given by G = \frac{4\pi A_e}{\lambda^2}, where A_e is the effective aperture area and \lambda is the wavelength, reflecting typical aperture efficiencies of 0.5 to 0.8 in practical designs. Horn antennas can achieve gains up to 25 dBi, depending on aperture size and frequency, with standard gain horns commonly providing 10 to 25 dBi for calibration and testing purposes. Their bandwidth is notably wide for aperture antennas, often reaching 10:1 to 20:1 frequency ratios in broadband configurations, such as ridged or exponential horns, allowing operation over multi-octave spans without significant performance degradation. The modern horn antenna was invented in the late 1930s at Bell Laboratories, with key contributions from George C. Southworth, who developed early waveguide-based horns for microwave transmission experiments. These antennas are standard components in radar systems for their directive beams and in satellite communication feeds due to their low sidelobe levels and efficient energy coupling. Pyramidal horns exhibit high polarization purity, naturally supporting linear polarization aligned with the waveguide mode, while modifications such as mode launchers or dielectric inserts enable circular polarization through excitation of orthogonal modes like TE11 and TM11.

Parabolic and Reflector Antennas

Parabolic reflectors, a type of reflector antenna, utilize a curved surface shaped as a paraboloid of revolution to focus electromagnetic waves onto a specific point or direct them into a narrow beam, achieving high directivity essential for long-distance communications and radio astronomy. The surface equation for a parabolic reflector is typically given by y^2 = 4fx, where f is the focal length, determining the position of the feed antenna. The focal length can be calculated as f = \frac{D^2}{16h}, with D as the reflector diameter and h as the depth of the dish. The gain G of such an antenna follows the aperture efficiency formula G = \frac{4\pi A \eta}{\lambda^2}, where A = \pi (D/2)^2 is the physical aperture area, \eta is the aperture efficiency (typically 0.55–0.70 for well-designed systems), and \lambda is the wavelength. The design originated from experiments by Heinrich Hertz in 1887–1888, who employed cylindrical parabolic reflectors to generate and detect electromagnetic waves at around 450 MHz, demonstrating reflection and polarization properties. Parabolic antennas saw widespread adoption in the 1960s with the rise of satellite communications, exemplified by systems like Telstar, which relied on large parabolic dishes for transatlantic links. Feed configurations play a critical role in performance; prime focus places the feed directly at the focal point for simplicity, but it exposes spillover radiation toward the ground, increasing noise. The Cassegrain configuration, using a hyperbolic subreflector to redirect waves from a feed behind the main reflector, minimizes spillover by directing excess energy toward the cooler sky, enhancing system temperature and efficiency. To mitigate aperture blockage by the feed structure, which reduces effective area and introduces sidelobes, offset feeds position the feed asymmetrically outside the main beam path. Shaped reflectors modify the paraboloid surface to produce contoured beams tailored for specific coverage areas, such as regional satellite broadcasting, using optimization techniques like B-spline curvature synthesis. Polarization handling in these antennas supports dual linear or circular modes through ortho-mode transducers (OMTs), which separate or combine orthogonal polarizations in the waveguide feed with high isolation (>40 dB typically). These systems are often illuminated by horn feeds for broadband operation and low sidelobes.

Array Antennas

Linear and Planar Arrays

Linear arrays consist of multiple antenna elements arranged along a straight line, enabling enhanced directivity and beam steering through controlled phase and amplitude excitation of each element. Early developments in linear arrays date back to the early 1900s, with Edmond Bruce patenting the Bruce array in 1931, a meandered wire configuration used for direction finding in radio astronomy and communications. By the 1950s, linear arrays evolved into phased arrays for military radar applications, allowing rapid electronic beam repositioning to track fast-moving targets, as demonstrated in developments at the Applied Physics Laboratory. In linear arrays, radiation patterns are shaped by the array factor, which for an N-element uniform array is given by AF(\theta) = \sum_{n=0}^{N-1} e^{j (n k d \cos\theta + \beta_n)}, where k = 2\pi / \lambda is the wavenumber, d is the inter-element spacing, \theta is the angle from the array axis, and \beta_n is the progressive phase shift for the nth element. Broadside arrays achieve maximum radiation perpendicular to the array axis (\theta = 90^\circ) by setting all elements in phase (\beta_n = 0), resulting in a symmetric pattern normal to the line of elements. In contrast, endfire arrays direct maximum radiation along the array axis (\theta = 0^\circ) using a progressive phase shift \delta = -k d, which aligns the wavefronts constructively in the forward direction while suppressing backward radiation. Planar arrays extend linear configurations into two dimensions, using rectangular or circular grids of elements to provide control over beams in three-dimensional space, offering higher gain and precise scanning for applications like radar and satellite communications. In phased planar arrays, electronic beam steering is achieved by applying phase gradients across rows and columns, with the scan angle given by \theta_{\text{scan}} = \arcsin(\delta / k d), where \delta is the phase shift between adjacent elements. However, mutual coupling between closely spaced elements in both linear and planar arrays alters input impedances and current distributions, typically reducing radiation efficiency by 1-3 dB and distorting the intended beam pattern.

Parasitic and Driven Arrays

Parasitic arrays consist of a single driven element coupled to one or more passive elements that are not directly connected to the feed line, relying on mutual coupling to shape the radiation pattern. These passive elements function as reflectors or directors based on their lengths relative to the wavelength, enhancing directivity and gain in a specific direction without requiring active phasing. In contrast, driven arrays feature multiple elements, each fed by separate transmission lines, allowing for pattern diversity through selective excitation or switching of feeds to produce varied radiation patterns for applications like diversity reception. The Yagi-Uda antenna exemplifies a parasitic array, comprising a driven element—typically a half-wave dipole—paired with a reflector slightly longer than 0.5λ positioned approximately 0.2λ behind it, and one or more directors shorter than 0.5λ (around 0.44λ) spaced similarly ahead. The reflector, by being longer, reflects energy forward, while directors progressively shorter guide the wave in the endfire direction, resulting in typical gains of 10-15 dBi and front-to-back ratios exceeding 20 dB for multi-element designs. Invented in the 1920s by Shintaro Uda and Hidetsugu Yagi at Tohoku Imperial University in Japan, the Yagi-Uda became a staple for television reception due to its compact size, high directivity, and simplicity in construction. Optimization of Yagi-Uda arrays involves tuning element lengths and spacings to achieve desired impedance, often around 50 Ω, using techniques like folded dipoles or matching networks to minimize return loss. For instance, reflector lengths near 0.5λ and director lengths around 0.4λ, with uniform 0.2λ spacing, balance gain and bandwidth, though variations can enhance performance for specific frequencies. A broadband variant, the Log-Yagi, combines Yagi principles with logarithmic scaling of elements to extend operational frequency range, though detailed designs fall under log-periodic antennas. Driven arrays with multiple feeds enable pattern diversity by exciting subsets of elements to generate orthogonal or complementary beams, improving signal reliability in multipath environments without complex phasing networks. These configurations, often employing switching or hybrid feeds, provide flexibility for applications requiring adaptive coverage, such as mobile communications, while maintaining higher impedance matching ease compared to parasitic types.

Traveling Wave Antennas

Helical Antennas

Helical antennas consist of a conducting wire wound in the shape of a helix, typically mounted perpendicularly over a ground plane, and operate based on traveling wave principles where the current propagates along the helix structure. Invented by John D. Kraus in 1947 at Ohio State University, this design emerged from studies on endfire arrays and has become a foundational element in radio engineering for its ability to produce directive beams with circular polarization. The antenna supports two primary operating modes: axial and normal, distinguished by the helix dimensions relative to the operating wavelength λ. In the axial mode, the helix circumference C is approximately equal to λ (typically 0.75λ to 1.25λ for optimal performance), and the pitch angle α—defined as the angle between the tangent to the helix and its axis—is set between 12° and 14° to achieve circular polarization. This configuration results in a half-power beamwidth (HPBW) of about 52° and directivity gains ranging from 10 to 15 dBi, depending on the number of turns N (often 6 to 10 for practical designs), making it suitable for applications requiring focused, circularly polarized radiation along the helix axis. The axial mode inherently provides a bandwidth of approximately 1.2:1 for a voltage standing wave ratio (VSWR) below 2:1, with an input impedance around 140 Ω, which often requires matching networks for standard 50 Ω systems. To suppress backlobes and enhance forward gain, the helix is mounted on a ground plane reflector with a diameter of about 0.75λ, which reflects energy away from the rear direction. In contrast, the normal mode occurs when the helix dimensions are small compared to λ (C << λ), producing an omnidirectional pattern with linear polarization perpendicular to the axis, similar to a short dipole or loop but with reduced efficiency. This mode is less common for directive applications but useful in compact, broadband scenarios where directionality is not required. A variant, the quadrifilar helical antenna (QHA), uses four intertwined helices to provide hemispherical coverage with right-hand circular polarization (RHCP), widely adopted for GPS receivers due to its compact size, low axial ratio, and resistance to multipath interference.

Log-Periodic Antennas

Log-periodic antennas are frequency-independent structures designed for wideband operation through self-similar geometric scaling of their elements, ensuring consistent electrical performance across a broad frequency range. The log-periodic dipole array (LPDA), a prominent variant, consists of multiple dipole elements arranged along a common axis with lengths and spacings that decrease logarithmically toward the high-frequency end. Key design parameters include the scale factor τ, typically ranging from 0.8 to 0.95, which defines the geometric ratio between consecutive elements, and the spacing factor σ, usually 0.15 to 0.25, which governs the relative distance between them. These parameters enable bandwidth ratios of 10:1 or greater, allowing operation from VHF to UHF frequencies without significant degradation in performance. In operation, the active region of an LPDA—the portion where the nearest half-wavelength dipole dominates radiation—shifts toward the shorter elements as frequency increases, maintaining consistent impedance and radiation characteristics. The input impedance remains periodic and relatively stable, typically around 50-75 Ω, facilitating matching to standard transmission lines. LPDA designs exhibit unidirectional patterns with gains of 6-8 dBi, providing moderate directivity suitable for directional applications. Developed in the 1950s by Raymond H. DuHamel and Dwight E. Isbell at the University of Illinois, the LPDA evolved from early logarithmically periodic concepts to address broadband needs in communication systems. Variants include standard dipole configurations for VHF and UHF bands, as well as toothed or trapezoidal tooth designs for planar implementations in higher frequencies. These antennas are widely used in television broadcast reception and electromagnetic compatibility (EMC) testing due to their reliable wideband performance.

Planar and Microstrip Antennas

Patch Antennas

Patch antennas, also known as microstrip patch antennas, consist of a metallic patch mounted on a dielectric substrate with a ground plane beneath it, forming a low-profile, planar structure suitable for integration into compact devices. These antennas operate by exciting resonant modes in the patch, where the dominant radiation mechanism arises from fringing electric fields at the patch edges, particularly along the length, leading to broadside radiation patterns. The fundamental transverse magnetic (TM_{10}) mode is typically dominant, producing linear polarization normal to the patch surface. For a rectangular patch, the resonant length L is approximately given by L = \frac{\lambda}{2 \sqrt{\epsilon_r}}, where \lambda is the free-space wavelength at the operating frequency and \epsilon_r is the relative permittivity of the substrate, commonly ranging from 2 to 10 to balance size and performance. The width W is selected to achieve a characteristic impedance near 50 \Omega, typically around W \approx 1.5 L for efficient matching, enabling direct integration with standard transmission lines like microstrip feeds. Basic designs exhibit narrow bandwidths of 1-5%, limited by the low Q-factor of the resonant cavity formed by the patch and ground plane, though enhancements such as stacked patches or slots can extend this to about 20% while maintaining compact form factors. Circular polarization can be achieved in rectangular or square patches through techniques like truncating opposite corners to perturb the degenerate modes, or by employing dual orthogonal feeds excited with 90° phase difference to generate quadrature fields. The concept of microstrip patch antennas originated in the early 1970s, with key developments patented by Robert E. Munson in 1975 for conformal array structures. Due to their lightweight, conformal nature, and ease of fabrication using printed circuit technology, patch antennas have become ubiquitous in modern wireless systems, including cellular phones for multiband operation and WiFi modules for 2.4/5 GHz connectivity. Patch antennas lend themselves well to array configurations, where multiple elements arranged in linear or planar grids can achieve higher directive gain, up to 20 dBi or more, by constructive interference of broadside lobes, making them ideal for base stations and radar applications.

Slot and Dielectric Resonator Antennas

Slot antennas operate on principles derived from Babinet's principle, which establishes a duality between a slot cut in a conducting sheet and its complementary dipole antenna. According to this principle, the electric field of the dipole corresponds to the magnetic field of the slot, and vice versa, with the radiation patterns rotated by 90 degrees and the impedances related by Z_{\text{slot}} \cdot Z_{\text{dipole}} = \left( \frac{\eta_0}{2} \right)^2, where \eta_0 is the intrinsic impedance of free space (approximately 377 Ω). This duality implies an inversion in impedance characteristics: a low-impedance dipole (typically around 73 Ω for a thin half-wave dipole) complements a high-impedance slot (often exceeding 500 Ω). The seminal application of Babinet's principle to slot antennas was detailed by H. G. Booker in 1946, linking optical diffraction to antenna behavior. A basic slot antenna consists of a narrow rectangular aperture, typically with a length of \lambda/2 (half-wavelength at the operating frequency) cut into a conducting ground plane or waveguide wall, where \lambda is the wavelength in the medium. This length ensures resonance, with the slot's narrow width (much less than \lambda/2) minimizing higher-order effects and controlling the radiation pattern. Radiation arises from an equivalent magnetic current density along the slot's edges, induced by the electric field across the aperture, which contrasts with the electric current in a dipole. For integration with waveguides, slots are offset or inclined to couple energy efficiently, achieving impedance matching through position adjustments. To enhance bandwidth, variations like the U-slot antenna introduce a U-shaped cutout within a patch or ground plane, creating coupled resonant modes that broaden the impedance bandwidth to approximately 30%. This design leverages the interaction between the fundamental patch mode and a slot-induced LC resonator, analyzed via coupled mode theory, enabling operation over wider frequencies without multilayer complexity. Dielectric resonator antennas (DRAs) utilize high-permittivity dielectric materials (\varepsilon_r > 10, often ranging from 10 to 100) to form compact radiating structures, typically cylindrical or rectangular in shape, mounted on a ground plane. These geometries support low-loss operation at microwave and millimeter-wave frequencies, with the dielectric concentrating fields to achieve high radiation efficiency. The quality factor Q of DRAs typically ranges from 100 to 1000, scaling with \varepsilon_r, which influences stored energy and thus bandwidth. The fundamental operating mode for cylindrical DRAs is often the \text{TE}_{11\delta} mode (transverse electric with no variation in the azimuthal direction and a small longitudinal component), excited via probe, slot, or aperture coupling to radiate broadside patterns. Rectangular DRAs similarly excite \text{TE}_{xy\delta} modes, offering flexibility in aspect ratios for impedance control. Basic DRAs exhibit fractional bandwidths of 5-15%, limited by the high Q, but suitable for narrowband applications where size and efficiency are prioritized over wide tuning. Theoretical foundations for DRAs were established in the 1980s through systematic studies by S. A. Long, M. W. McAllister, and S. A. Shen, who analyzed the resonant cylindrical dielectric cavity and its radiation properties, confirming efficient operation with minimal conductor losses. In modern contexts, DRAs are increasingly applied in millimeter-wave 5G systems, where their compact size and high efficiency support array configurations for beamforming, achieving bandwidths around 7-10% in the 28 GHz band. Polarization in DRAs can be tailored for linear or circular operation by perturbing the resonator geometry, such as asymmetric cuts or notches that degenerate degenerate modes to produce orthogonal components with phase quadrature. This flexibility, often via simple uniaxial perturbations, enables circular polarization without complex feeds, enhancing performance in multipath environments.

Specialized and Emerging Antenna Types

Broadband and Frequency-Independent Antennas

Broadband antennas are engineered to operate effectively across a wide range of frequencies, often achieving bandwidths exceeding an octave, which is essential for applications requiring versatility without retuning. Frequency-independent antennas, a subset of these designs, rely on self-similar geometries that maintain performance invariance with frequency scaling, as theorized by Victor Rumsey in the 1950s. Rumsey's principle of scale invariance posits that antennas with shapes unaltered by linear scaling will exhibit radiation characteristics independent of frequency, provided the structure is large compared to the wavelength at the lowest operating frequency. This theory laid the foundation for antennas used in military surveillance and broadband communication systems, where consistent performance over decades of frequency is critical. A prominent example of frequency-independent antennas is the equiangular spiral antenna, characterized by its self-similar, logarithmic spiral arms that expand at a constant angle. These antennas achieve bandwidths greater than 10:1, supporting circular polarization suitable for wideband radar and electronic warfare applications. The equiangular spiral's radiation pattern remains stable across its operational band due to the progressive excitation of current modes along the arms, with the active region shifting outward as frequency increases. Developed in the mid-20th century, such spirals have been integral to airborne surveillance systems, offering balanced bidirectional radiation with axial ratios typically below 3 dB over the band. The discone antenna represents another key broadband design, featuring an omnidirectional pattern in the horizontal plane and a conical shape that approximates a ground plane for vertical polarization. It consists of a metal disc atop a cone with apex angle α typically between 45° and 60°, enabling low VSWR (<2:1) from approximately 0.1 times the maximum frequency to the maximum frequency, often spanning 10:1 bandwidths. This structure supports TEM mode propagation, minimizing impedance variations and making it ideal for VHF/UHF monitoring and broadcast reception. Discones are widely deployed in fixed-site applications due to their simplicity and ruggedness, though they exhibit a null in the vertical pattern near the horizon. Biconical antennas, similar in broadband capability, employ two opposing cones to form a balanced TEM transmission line, radiating with gains of 2-4 dBi across wide bands. These are favored in EMC testing and calibration setups for their uniform field illumination and ease of polarization adjustment. Unlike spirals, bicones provide linear polarization and are often filled with dielectric to extend low-frequency performance. While these antennas excel in bandwidth, they face limitations such as large physical size at lower frequencies—often requiring diameters on the order of a wavelength—and inherently low gain due to their widebeam patterns. For instance, equiangular spirals may need absorbers to suppress back radiation, adding complexity. Log-periodic antennas, another frequency-independent design using periodic scaling, typically achieve bandwidths around 10:1. Overall, these passive geometric approaches remain foundational for scenarios demanding untuned, wideband operation.

Metamaterial and Reconfigurable Antennas

Metamaterial antennas exploit artificially structured materials exhibiting unconventional electromagnetic properties, such as negative permittivity and permeability, to surpass traditional antenna limitations in size, efficiency, and directivity. The conceptual groundwork originated with Veselago's 1968 analysis of substances possessing simultaneously negative values of dielectric constant ε and magnetic permeability μ, predicting phenomena like negative refraction and reversed Doppler shift. Building on this, Pendry's 2000 proposal of a "perfect lens" using negative-index materials demonstrated the potential for subwavelength focusing, igniting experimental realizations in the early 2000s through split-ring resonators and wire arrays that achieved effective negative refractive indices at microwave frequencies. These foundations enabled antennas with enhanced performance, particularly in compact designs where conventional materials fall short. Negative-index metamaterials facilitate superdirectivity by enabling phase compensation that concentrates radiation into narrower beams beyond the conventional directivity limit for electrically small antennas. For instance, transmission-line-based negative-refractive-index structures have been used to design compact monopole antennas exhibiting superdirectivity in dual bands, with directivities exceeding 10 dB in subwavelength apertures. Similarly, epsilon-near-zero (ENZ) metamaterials, where the real part of permittivity approaches zero, support cloak-inspired low-profile designs by creating impedance-matched boundaries that minimize scattering and enhance aperture efficiency. ENZ slabs, for example, enable multi-beam antennas with directive emission from line sources, achieving beam angles controlled by material thickness without bulky reflectors. These properties allow for low-profile antennas suitable for integration into conformal surfaces. Reconfigurable antennas incorporate active elements like varactor diodes or microelectromechanical systems (MEMS) to provide tunability in frequency, pattern, or polarization, offering agility without mechanical movement. Varactor diodes enable continuous tuning via capacitance variation, as seen in metasurface-loaded designs that achieve beam steering over ±30° in the Ka-band by altering reflection phases electronically, bypassing the need for phase shifters. MEMS switches, with their low insertion loss (<0.5 dB) and high isolation (>20 dB), support discrete reconfiguration for pattern diversity, such as switching between broadside and endfire modes in compact arrays. Since the 2010s, prototypes integrating these with metamaterials have advanced 5G applications, enabling more compact designs such as sub-5 mm elements at mmWave frequencies while maintaining fractional bandwidths over 10%. Ongoing research as of 2025 extends these technologies to 6G, including reconfigurable intelligent surfaces (RIS) based on metamaterials for dynamic beam management and interference mitigation in dense networks. Recent innovations also include shape-morphing antennas that adjust frequency ranges for versatile sensing and communication. Despite these advances, challenges persist, including ohmic losses in metamaterial structures where loss tangents often exceed 0.01, degrading efficiency to below 70% at high frequencies, and limited power handling capacities typically under 10 W due to dielectric breakdown in subwavelength elements. These issues necessitate low-loss fabrication techniques, such as cryogenic etching for resonators. Applications include cognitive radio systems, where frequency reconfigurability via varactors allows dynamic spectrum access across 1-6 GHz bands, and satellite communications, enabling adaptive beams for tracking via MEMS-tuned patterns to mitigate interference.

References

  1. [1]
    Antenna Classification - MATLAB & Simulink - MathWorks
    Antennas are classified based on the radiation pattern or the feeding mechanism. Antenna radiation pattern is the angular variation of signal strength around ...
  2. [2]
    Antenna Basics and Radio Wave Propagation - IEEE Xplore
    Antennas can be classified in terms of their structures such as wire antennas, aperture antennas, and microstrip antennas. In a transmitter or receiver ...
  3. [3]
    An Introduction to Common Antenna Designs - Ansys
    Feb 21, 2023 · Common Antenna Types · Half-wave Dipole Antennas · PIFA Antennas · Horn Antennas · Yagi-Uda Antennas · Slot Antennas · Rectangular Patch Antennas.
  4. [4]
    Types of Antennas in Modern Technology | Arrow.com
    Jul 12, 2023 · There are three main types of polarization: vertical (linear), horizontal (linear), and circular. Vertical polarization antennas are commonly ...Types Of Antennas And How... · Antenna Bandwidth · Array Antennas
  5. [5]
    [PDF] CHAPTER 3: ANTENNAS
    Mar 3, 2022 · Antennas couple propagating electromagnetic waves to and from circuits and devices, typically using wires (treated in Section 3.2) or apertures ...
  6. [6]
    Microwaves101 | Microstrip Patch Antennas - Microwave Encyclopedia
    The microstrip antenna was first proposed by G.A. Deschamps in 1953, but didn't become practical until the 1970s when it was developed further by researchers ...
  7. [7]
    [PDF] History of Antennas
    Early antennas included wire antennas (19th century), vertical top-loaded (1875), dipole (1888), and the first directive antenna (1901). Later, wave tilt (1923 ...
  8. [8]
    Planar Integrated Antenna Technology - Microwave Journal
    Jul 1, 1999 · During the past several decades, much attention has been focused on the development of planar antennas. These antennas are highly desirable for ...
  9. [9]
    [PDF] Overview of Antennas
    4. Effective Aperture & Effective Length: If an antenna is able to extract a certain amount of power from a passing wave of known ...
  10. [10]
    [PDF] Chapter 11: Common Antennas and Applications
    Jan 11, 2011 · Equation (11.1. 18) applies to any uniformly illuminated aperture antenna, and such antennas have on-axis effective areas A(θ,φ) that approach ...
  11. [11]
    Revisiting the 1888 Hertz experiment | American Journal of Physics
    We revisit the original experiment performed by Hertz in 1888 and use a simple setup to produce electromagnetic waves with a frequency in the range of 3 MHz.
  12. [12]
    The Helical Antenna | IEEE Journals & Magazine
    Depending on the helix geometry, the radiation may, in theory, be elliptically, plane, or circularly polarized. When the helix circumference is about 1 ...
  13. [13]
    None
    ### Summary of Half-Wave Dipole Antenna
  14. [14]
    What is the radiation resistance of a short dipole of length L?
    Feb 18, 2022 · Then short dipole is of length l/2, so the radiation resistance is given by R=80π^2((frac{L/2}{lambda})^2=20pi^2 (frac{L}{lambda})^2.) ...
  15. [15]
    Folded Dipole Antenna / Aerial - Electronics Notes
    In free space, this gives an increase in feed impedance from 73Ω to around 300Ω ohms. Additionally the RF antenna has a wider bandwidth. Folded dipole impedance ...
  16. [16]
    [PDF] Six great pioneers of wireless - EBU tech
    1888: The English physicist Oliver Lodge (later Sir Oliver. Lodge) demonstrated the importance of ´resonanceµ where oscillatory currents were concerned. Page ...
  17. [17]
    Blanco, Jeffrey L - OhioLINK ETD
    The first array antennas for producing shaped directive beam patterns were introduced during World War II for early radar systems using an array of dipole ...
  18. [18]
    [PDF] Chapter 10: Antennas and Radiation - MIT OpenCourseWare
    The effective length deff of a short dipole is approximately half its physical length [see (10.2.25) and Figure 10.2.3]. The reactance X of a short dipole ...
  19. [19]
    Radiation pattern and impedance of a quarter wavelength monopole antenna above a finite ground plane
    **Summary of Radiation Pattern and Impedance of a Quarter Wavelength Monopole Antenna:**
  20. [20]
    [PDF] 8 Antennas
    If a monopole antenna is placed above a ground plane, the induced image charge will match the radiation from a dipole. This can be used to synthesize extra ...
  21. [21]
  22. [22]
    Fessenden and the Early History of Radio Science
    Sep 3, 1994 · Marconi, following the lead of Hertz, employed such a spark transmitter, but his antenna was a monopole type, a wire fed against ground. Since ...
  23. [23]
    Review on 5G Small Cell Base Station Antennas - IEEE Xplore
    Jun 26, 2024 · The following section explains the multiple bandwidth-enhanced antennas proposed for 5G base station applications. Several designing ...
  24. [24]
    [PDF] Wideband Compact Antennas for Wireless ... - VTechWorks
    ... H-plane [28]......................... 34. Figure 2 ... is an end-fed half-wave dipole. ... The antenna has an omnidirectional pattern in the x-y plane.
  25. [25]
    [PDF] L12_Loop.pdf - Electrical & Computer Engineering
    The small loops, regardless of their shape, have a far-field pattern very similar ... This is expected because the small loops are effectively magnetic dipoles.
  26. [26]
    E-Field & H-Field Antennas: what are they, what are the differences
    For example, a horizontal loop radiates a vertically polarized E-field (or horizontally polarized H-field). Radiation Pattern: Small loop antennas typically ...
  27. [27]
    History of Wireless Communications | 2015-07-22 - Microwave Journal
    Jul 22, 2015 · 1883 – Irish physicist and chemist George Francis FitzGerald published a formula for the power radiated by a small loop antenna. 1884 ...
  28. [28]
    Small Loop Antennas: Part 1 - Simulations and Applied Theory
    Mar 26, 2006 · In this application note we will explore antenna theory and design using EZNEC, an easy-to-use antenna simulator.
  29. [29]
    Horn Antenna Theory: Equations & Formulas - Electronics Notes
    The difference in phase between the centre point and the edges is called the phase error. This increases with the flare angle reducing the gain, but increasing ...Missing: minimization | Show results with:minimization
  30. [30]
    [PDF] Horn Antennas (Rectangular horn antennas. Circular apertures.)
    The E-plane principal pattern of the pyramidal horn is the same as the E-plane principal pattern of the E-plane sectoral horn. The same holds for the H-plane.
  31. [31]
    Antenna Gain and Effective Area: Formulas and Types
    Explore the relationship between antenna gain and effective area with formulas for parabolic dishes, isotropic radiators, and horn antennas.
  32. [32]
  33. [33]
    Schwarzbeck- Broadband Horn Antennas - Reliantemc
    The usable bandwidth of horn antennas is typically of the order of 10:1, and can be up to 20:1 (for example allowing it to operate from 1 GHz to 20 GHz).
  34. [34]
    Transforming Linear to Circular Polarization on Horn Antennas by ...
    Although cheap and highly directional, pyramidal horn antennas provide a linear polarization. Polarization transformation can be achieved by three main methods.
  35. [35]
    Parabolic Reflector Antenna | Theory & Formulas - Electronics Notes
    ... focal point. To determine this it is necessary to know the focal length. f=D216c. Where: f is the focal length. D is the diameter of the reflector c is the ...
  36. [36]
    Parabolic Reflector Antenna Gain: Formula Calculation
    Antenna efficiency: The efficiency of the antenna has a significant effect on the overall parabolic reflector gain. Typical figures are between 50 and 70%. The ...
  37. [37]
    [PDF] Large Antennas of the Deep Space Network - DESCANSO
    Feb 13, 2002 · Reflector antennas have existed since the days of Heinrich Hertz. ... thus, nearly equivalent performance to a 70-m parabolic reflector.
  38. [38]
    1964 Adoption of Our Parabolic Antenna for a Syncom Satellite
    Our parabolic antenna for a Syncom satellite was adopted as part of the antenna manufactured by NEC Corporation, helping realize the dreams and hopes of people ...
  39. [39]
    [PDF] LECTURE 12: Reflector Antennas - Electrical & Computer Engineering
    They have almost no spillover toward the noisy ground, as do the single-feed reflector antennas. Their spillover (if any) is directed toward the much less noisy ...
  40. [40]
  41. [41]
    Orthomode transducer - Radartutorial.eu
    An orthomode transducer (OMT) or Orthomode Junction (OMJ) is a waveguide component that acts as a bidirectional polarization switch for orthogonally polarized ...
  42. [42]
    Edmond Bruce - Engineering and Technology History Wiki
    Dec 4, 2019 · Edmond Bruce was best known for his inventions of the rhombic antenna and Bruce Array. He was born on September 28, 1899 in St. Louis.
  43. [43]
    [PDF] PHASED ARRAY ANTENNA DEVELOPMENT AT THE APPLIED ...
    During the late 1950s, the ability to reposition radar beams within microseconds became a necessity owing to the increasing demands on weapons systems. The ...
  44. [44]
    Effects of Mutual Coupling on Gain and Beam Width of a Linear ...
    Sep 9, 2022 · It is shown that for both the E and H planes, mutual coupling has an adverse effect on the performance of both phased array configurations. The ...
  45. [45]
    Navy Electricity and Electronics Training Series (NEETS), Module 10 ...
    Size for size, both the gain and directivity of parasitic arrays are greater than those of driven arrays. The disadvantage of parasitic arrays is that their ...
  46. [46]
    An Array of Antenna Arrays - Pasternack Blog
    Sep 13, 2018 · An array antenna is known as a driven or connected when all of the elements in an array are driven. Interestingly, if one or more elements ...
  47. [47]
    (PDF) YAGI-UDA ANTENNA DESIGN - ResearchGate
    Mar 13, 2022 · The paper is focused on principles and characteristics of Yagi-Uda antenna, and its performance parameters are examined.
  48. [48]
    [PDF] Yagi antenna design - NIST Technical Series Publications
    Directors of Different Length and Diameter Spaced 0.35A. Yagi Antenna Design Data Showing the Relationship Between Element. Diameter to Wavelength Ratio and ...
  49. [49]
    The History Column: The Yagi Antenna - IEEE AESS
    In 1928 at the Tohoku Imperial University at Sendai, Japan, Yagi and his student Shintaro Uda developed a novel antenna array consisting of one active element ...
  50. [50]
    Optimization of Gain, Impedance, and Bandwidth of Yagi-Uda Array ...
    Aug 7, 2025 · A method of using particle swarm optimisation (PSO) algorithms to optimise the element spacing and lengths of Yagi-Uda antennas is presented.<|control11|><|separator|>
  51. [51]
    Multifunctional twelve port frequency agile diversity antenna ... - Nature
    May 17, 2023 · In this paper, a frequency reconfigurable polarization and pattern diverse Multiple-Input-Multiple-Output (MIMO) antenna is presented for indoor scenarios.Introduction · Antenna Design · Antenna Gain And Efficiency<|control11|><|separator|>
  52. [52]
    [PDF] Log-Periodic Dipole Array (LPDA)
    DuHamel & Isbell- Planar and wire trapezoidal toothed log-periodic antennas [Balanis, Figure 11.7, p. 553]. Page 7. 7 / 14. Trapezoidal wire antenna in Vee ...
  53. [53]
    Log periodic dipole arrays - Semantic Scholar
    A compact planar log-periodic dipole array (LPDA) antenna operating from 0.55 to 9 GHz with high measured gain is proposed to be implemented in a ...
  54. [54]
    [PDF] DESIGN AND ANALYSIS OF LOG PERIODIC DIPOLE ARRAY ...
    Log periodic antenna was invented by Dwight E. Isbell and Raymond DuHamel was licensed by University of Illinois at. Urbana Champaign. The main invention of ...
  55. [55]
    The Secrets of the Log Periodic Antenna
    Log periodic antennas are a group of varying dipole antennas with uniform characteristics over a wide frequency range. The active region shifts with frequency.
  56. [56]
  57. [57]
    3144 Log-Periodic Dipole Array - ETS-Lindgren
    The 3144 LPDA is a medium gain (6 dBi) antenna with a frequency range of 80 MHz to 2 GHz, high power input, and a tough powder coat finish.Missing: typical | Show results with:typical
  58. [58]
    Practical Log-Periodic Antenna Designs, May 4, 1964 Electronics ...
    Jul 11, 2019 · However, when Dwight Isbell and Raymond DuHamel of the University of Illinois came up with the log periodic concept in 1958, they did not ...<|control11|><|separator|>
  59. [59]
  60. [60]
    Slot Antenna
    The dual antenna is similar to a dipole antenna. The voltage source is applied at the center of the dipole, so that the voltage source is rotated.
  61. [61]
    [PDF] U-Slot Patch Antenna Principle and Design Methodology Using ...
    Abstract—Patch antennas incorporating a U-shaped slot are well-known to have relatively large (about 30%) impedance bandwidths.
  62. [62]
    [PDF] Dielectric Resonator Antennas: Designs and Advances
    Jun 27, 2014 · The DRA is a resonant antenna, fabricated from a high-permittivity (from about 6 to 100) dielectric material mounted on a ground plane and fed ...Missing: 100-1000 TE11δ
  63. [63]
  64. [64]
  65. [65]
    A Review of Circularly Polarized Dielectric Resonator Antennas - NIH
    Various DRAs with circular polarization have been designed by exploiting different dielectric resonator (DR) geometries and feeding configurations. Feeding ...
  66. [66]
    THE ELECTRODYNAMICS OF SUBSTANCES WITH ... - IOP Science
    THE ELECTRODYNAMICS OF SUBSTANCES WITH SIMULTANEOUSLY NEGATIVE VALUES OF AND μ. Viktor G Veselago. © 1968 American Institute of Physics
  67. [67]
    Negative Refraction Makes a Perfect Lens | Phys. Rev. Lett.
    Oct 30, 2000 · An unconventional alternative to a lens, a slab of negative refractive index material, has the power to focus all Fourier components of a 2D image.
  68. [68]
    Design of multi-beam antennas based on Epsilon-Near-Zero ...
    A novel Epsilon-Near-Zero (ENZ) metamaterial-assisted multi-beam antenna is presented. Simulation results show that the cylindrical wave emitted from a line ...Missing: inspired | Show results with:inspired
  69. [69]
    Continuous beam steering with a varactor-based reconfigurable ...
    Sep 19, 2023 · Varactor diodes have been proposed as active tuning elements of reconfigurable metasurfaces in the Ka-band.
  70. [70]
    [PDF] Reconfigurable Antennas: A Review of Recent Progress and Future ...
    Oct 11, 2021 · Abstract—Reconfigurable antennas are devices that can dynamically alter their geometry and/or electromagnetic properties to facilitate ...
  71. [71]
    Using Metamaterials in mmWave 5G Antennas - Microwave Journal
    Jun 13, 2024 · This article delves into the application of the sub-wavelength MTM structures for compact, high gain antennas.
  72. [72]
    Experimental analysis of metamaterial integrated applicator for ...
    May 29, 2025 · A metamaterial integrated applicator is fabricated on an FR4 substrate (having a dielectric constant (εr) of 4.4, tangent loss (tan δ) of 0.018, ...
  73. [73]
    [PDF] Reconfigurable Antenna and Cognitive Radio for Space Applications
    This study proposes a communication system with Reconfigurable Antenna (RA) and Cognitive Radio (CR) to mitigate the RFI impact. A cognitive radio is an ...
  74. [74]
    Reconfigurable Antennas for Wireless and Space Applications
    Apr 5, 2012 · This paper discusses the different reconfigurable components that can be used in an antenna to modify its structure and function.