Fact-checked by Grok 2 weeks ago

Radio propagation

Radio propagation refers to the behavior of radio waves—electromagnetic waves with frequencies typically ranging from 3 kHz to 300 GHz—as they travel from a transmitter to a through various , including , the atmosphere, and terrestrial environments. This process is fundamental to communication systems, such as , mobile networks, and , where the propagation characteristics determine signal strength, coverage, and reliability. The primary mechanisms governing radio propagation include reflection, where waves bounce off surfaces like buildings or the ground; refraction, the bending of waves due to changes in the medium's refractive index, often in the troposphere; diffraction, which allows waves to bend around obstacles such as hills; scattering, the dispersion of waves by small particles or rough surfaces; and absorption, the loss of energy as waves interact with atmospheric gases or precipitation. These phenomena can lead to multipath propagation, where signals arrive via multiple paths, causing interference, fading, or constructive/destructive effects at the receiver. Several factors influence propagation, including frequency (higher frequencies experience greater , following the P_r = P_t G_t G_r \left( \frac{\lambda}{4\pi r} \right)^2, where P_r is received power, P_t is transmitted power, G_t and G_r are gains, \lambda is , and r is distance), terrain irregularity, atmospheric conditions (e.g., ionospheric reflection for high-frequency sky waves), and . Propagation modes are categorized as ground waves (dominant at low frequencies below 2 MHz, following the Earth's curvature), sky waves (reflected by the for long-distance communication), and line-of-sight space waves (prevalent above 30 MHz for VHF/UHF, limited by the horizon but extendable via tropospheric ducting). In mobile and urban settings, these effects are modeled using tools like the Irregular Terrain Model (ITM) to predict coverage over varied landscapes. Understanding radio propagation is crucial for designing efficient systems, mitigating losses (e.g., 20 log(d) dB in free space), and addressing challenges like signal in non-line-of-sight scenarios. Advances in modeling continue to support applications from cellular networks to communications, ensuring reliable performance across diverse environments.

Fundamentals

Definition and Principles

Radio propagation is the study of how electromagnetic waves in the radio frequency range, from 3 kHz to 300 GHz, travel from a transmitter to a , undergoing effects such as off surfaces, around obstacles, by particles or irregularities, and absorption by the medium. These processes determine the signal's strength, direction, and reliability in various environments. The core principles governing radio propagation derive from , which unify electricity and magnetism by describing how varying s generate magnetic fields and vice versa, leading to self-sustaining electromagnetic waves that propagate through space at the , c \approx 3 \times 10^8 m/s. A key characteristic of these waves is , which refers to the time-varying orientation of the vector; occurs when the field oscillates along a fixed axis, while arises when the field rotates in a helical , either (right-hand) or counterclockwise (left-hand) relative to the direction of . The \lambda, a fundamental parameter influencing behavior, is calculated as \lambda = c / f, where f is the wave's . Path loss quantifies the reduction in signal power during propagation, with free-space path loss (FSPL) providing a theoretical baseline for unobstructed conditions: \text{FSPL} = \left( \frac{4\pi d f}{c} \right)^2, where d is the ; this quadratic dependence on and highlights the inherent even in vacuum. Early insights into radio propagation emerged in the 1880s when experimentally verified predictions by generating and detecting electromagnetic waves using spark-gap apparatus, observing their , , and over distances up to several meters. This work laid the groundwork for practical applications, exemplified by Guglielmo Marconi's 1901 reception of a transatlantic signal in Newfoundland from a transmitter in , which was later explained by reflection from the —a layer in the upper atmosphere proposed by and Arthur Kennelly in 1902 to account for the unexpected long-distance propagation beyond the Earth's curvature.

Electromagnetic Spectrum Relevance

Radio propagation pertains to the portion of the electromagnetic spectrum known as radio waves, which span frequencies from 3 kHz to 300 GHz, corresponding to wavelengths from 100 km to 1 mm. This range is defined by the International Telecommunication Union (ITU) as the radio-frequency spectrum, distinguishing it from higher-frequency portions like microwaves beyond 300 GHz or lower-frequency extremely low frequencies below 3 kHz in some classifications. The ITU further subdivides this spectrum into standardized bands to facilitate international coordination and usage, including very low frequency (VLF: 3–30 kHz), low frequency (LF: 30–300 kHz), medium frequency (MF: 300 kHz–3 MHz), high frequency (HF: 3–30 MHz), very high frequency (VHF: 30–300 MHz), ultra high frequency (UHF: 300 MHz–3 GHz), super high frequency (SHF: 3–30 GHz), and extremely high frequency (EHF: 30–300 GHz). Note that extremely low frequency (ELF: 3–30 Hz) is sometimes included in broader definitions but is less common in standard propagation studies due to its specialized applications. These bands exhibit varying propagation characteristics tied to their wavelengths and frequencies; for instance, the HF band (3–30 MHz) has wavelengths of 10–100 m, enabling better around obstacles compared to higher bands like UHF (wavelengths of 10–100 cm), where signals attenuate more rapidly over irregular terrain. Lower-frequency bands such as VLF and LF generally experience less and diffract more effectively around physical barriers due to their longer wavelengths relative to obstacle sizes, supporting longer-range communications in challenging environments. In contrast, higher bands like SHF and EHF suffer increased atmospheric absorption but offer higher data rates for line-of-sight applications. The ITU manages these allocations through its Radio Regulations, which outline international frequency assignments to prevent interference and ensure efficient spectrum use, forming the basis for national spectrum management plans that account for propagation behaviors during planning. Countries develop National Tables of Frequency Allocations (NTFAs) aligned with ITU guidelines, incorporating propagation models to assign bands for services like broadcasting, mobile communications, and satellite links. A key attribute of radio waves across this spectrum is their non-ionizing nature, meaning they lack sufficient to remove electrons from atoms, allowing penetration through non-conductive materials such as , , and certain plastics—unlike higher-frequency or even visible light, which are more readily absorbed or reflected. This property underpins applications like networking indoors, where signals propagate through walls with minimal disruption.
BandSymbolFrequency RangeWavelength Range
Very Low FrequencyVLF3–30 kHz100–10 km
Low FrequencyLF30–300 kHz10–1 km
Medium FrequencyMF0.3–3 MHz1,000–100 m
High FrequencyHF3–30 MHz100–10 m
Very High FrequencyVHF30–300 MHz10–1 m
Ultra High FrequencyUHF300–3,000 MHz1–0.1 m
Super High FrequencySHF3–30 GHz0.1–0.01 m
Extremely High FrequencyEHF30–300 GHz0.01–0.001 m

Propagation Modes

Free-Space Propagation

Free-space propagation describes the idealized transmission of radio waves through a or uniform medium devoid of obstacles, atmospheric influences, variations, or multipath effects, where electromagnetic waves emanate spherically from a . In this scenario, an —a theoretical that emits power uniformly in all directions—serves as the baseline model, resulting in a that diminishes inversely with the square of the distance from the source due to the expanding wavefront. This model assumes far-field conditions, where the wavefront approximates a locally, but globally spreads over a spherical surface. The fundamental relationship governing power transfer in free-space propagation is the Friis transmission equation, which quantifies the received power at a distant antenna. The equation is expressed as P_r = P_t G_t G_r \left( \frac{\lambda}{4 \pi d} \right)^2 where P_r is the power available at the receiving antenna, P_t is the power supplied to the transmitting antenna, G_t and G_r are the gains of the transmitting and receiving antennas respectively, \lambda is the wavelength, and d is the separation distance between the antennas. This formula, originally derived for microwave frequencies, assumes matched polarization, far-field operation, and no losses beyond geometric spreading. The derivation of the Friis equation begins with the , which represents the directional density of the , with magnitude equal to the time-averaged S for a . For a spherical wavefront from an isotropic radiator, the power spreads over the surface of a of d, yielding S = \frac{P_t}{4 \pi d^2}; incorporating the transmitting antenna's gain G_t (which concentrates power in preferred directions) modifies this to S = \frac{P_t G_t}{4 \pi d^2}. The receiving antenna captures power proportional to its effective A_e = \frac{G_r \lambda^2}{4 \pi}, so P_r = S \cdot A_e, leading directly to the Friis form upon substitution. This outline highlights the geometric dilution of power and the role of antenna properties in free space./10%3A_Antennas/10.14%3A_Friis_Transmission_Equation) In free-space conditions, antenna directivity measures the concentration of radiated power in a specific direction relative to an isotropic source, defined as D = \frac{4 \pi U(\theta, \phi)}{P_{\rm rad}}, where U(\theta, \phi) is the radiation intensity and P_{\rm rad} the total radiated power; gain G equals directivity for lossless antennas, accounting for ohmic efficiency. These parameters are independent of distance in the far field, enabling predictable link budgets. For radar applications in free space, the maximum detection range follows from applying the Friis equation twice—once to the target and once back—yielding the radar range equation R_{\max} = \left[ \frac{P_t G^2 \lambda^2 \sigma}{(4\pi)^3 S_{\min}} \right]^{1/4} where \sigma is the target's radar cross-section and S_{\min} the minimum detectable power; this assumes a monostatic radar with identical transmit and receive antennas. While free space provides the theoretical foundation, real propagations deviate due to environmental factors like atmospheric absorption, setting the stage for more complex models.

Line-of-Sight Propagation

Line-of-sight (LOS) propagation describes the direct transmission of radio waves from a transmitter to a along a straight path unobstructed by significant or structures, limited primarily by the curvature of the to the radio horizon distance. In this mode, signals travel essentially as in free space but are constrained by geometric and refractive effects in the , distinguishing it from ideal vacuum propagation by incorporating real atmospheric bending. The optical horizon represents the visual line-of-sight limit based on pure , while the radio horizon extends approximately 4/3 times farther due to tropospheric , which bends waves downward toward the 's surface; this extension is modeled using a standard of 4/3 for the effective in predictions. For an at height h above a flat surface, the geometric to the horizon d is approximated by d = \sqrt{2 R h}, where R is the 's mean of approximately 6371 ; this formula provides a basic estimate under the flat- approximation, with total LOS range being the sum for transmitter and receiver heights. Even in nominally clear LOS paths, minor obstacles such as hills or buildings can intrude into the propagation path, leading to diffraction effects that allow some signal penetration. The knife-edge diffraction model is commonly used for such single, sharp-edged obstacles, where the blockage is analyzed relative to the s—ellipsoidal regions around the direct path critical for unobstructed propagation. The radius of the first Fresnel zone R_1 at the obstacle is R_1 = \sqrt{\frac{\lambda d_1 d_2}{d_1 + d_2}}, with \lambda as the wavelength and d_1, d_2 as distances from the transmitter and to the obstacle, respectively; clear LOS requires no significant intrusion into this zone (typically 60% clearance). In the knife-edge model, the diffraction loss depends on the Fresnel parameter \nu = -h \sqrt{2(d_1 + d_2)/(\lambda d_1 d_2)}, where h is the obstacle height relative to the direct path; for partial blockage near incidence (\nu around 0 to 2), ranges from about 6 to 20 , enabling marginal communication beyond strict geometric . This model assumes a perfectly sharp edge, with additional corrections for rounded obstacles to account for further loss. propagation dominates in the VHF (30–300 MHz) and UHF (300–3000 MHz) frequency bands, supporting applications like television and radio , where high-power transmitters provide coverage up to the radio horizon, and cellular networks, which rely on base stations with elevated antennas for direct urban and suburban links. In urban scenarios for these bands, multipath reflections from buildings and vehicles can superimpose on the direct signal, causing rapid signal fluctuations characterized briefly by statistics when scattered components dominate. Unlike free-space propagation models, real incorporates these horizon limits and for practical estimation.

Ground-Wave Propagation

Ground-wave propagation refers to the mode in which radio waves travel along the surface of the , guided by the boundary between the atmosphere and the ground. This mechanism primarily involves surface waves that induce currents in the conducting , allowing the signal to diffract around the planet's curvature and extend beyond the line-of-sight horizon. At lower frequencies such as (LF, 30–300 kHz) and (MF, 300 kHz–3 MHz), ground waves dominate due to their ability to follow the without significant or . Vertical polarization is essential for efficient ground-wave propagation, as it allows the to extend to the conducting surface, minimizing losses compared to horizontal polarization, which experiences rapid near the ground. The effective range of ground waves varies with , transmitter power, and ground . For MF signals used in , reliable coverage typically extends 100–200 km over land during daytime, with greater distances possible over more conductive surfaces like . In the LF and (VLF, 3–30 kHz) bands, ranges can exceed 1,000 km, particularly over , where high (around 4–5 S/m) reduces losses compared to land (typically 0.001–0.01 S/m). This conductivity dependence arises because waves propagate more efficiently over better conductors, enabling global-scale transmission in VLF applications. Attenuation in ground-wave propagation is modeled using the Sommerfeld-Norton theory, which accounts for the interaction between the wave and the 's surface. The theory describes the electric field strength over a homogeneous plane as E_z = \frac{60 I \, dl}{r} F, where I \, dl is the current element, r is the , and F is the function that incorporates and surface effects. A key parameter is the surface impedance Z = \sqrt{\frac{j \omega \mu}{\sigma + j \omega \varepsilon}}, which approximates to Z \approx \sqrt{\frac{j \omega \mu}{\sigma}} for conductive grounds where is negligible; here, \sigma is , \omega is , \mu is permeability, and \varepsilon is . Higher \sigma lowers |Z|, reducing and extending range. Ground-wave propagation finds primary applications in at , providing stable local coverage, and in navigation systems like at around 100 kHz, where ground waves offer precise timing over distances up to 1,700 nautical miles. These systems rely on the mode's reliability during , when it is the dominant signal path. In VLF, ground waves enable long-range submarine communication due to their low over —approximately 2–3 dB per 1,000 km—and ability to penetrate to depths of 10–40 meters, allowing submerged receivers to detect signals without surfacing.

Non-Ionospheric Non-Line-of-Sight Propagation

Non-ionospheric (NLOS) propagation occurs primarily through tropospheric mechanisms that enable radio signals to bypass direct line-of-sight paths without involving the , such as over obstacles, by atmospheric irregularities, and reflections from terrain or structures. These processes are dominant at VHF and UHF frequencies and above, where signals interact with the lower atmosphere and surface features to achieve transhorizon coverage. Unlike ground-wave propagation, which hugs the Earth's surface at lower frequencies, NLOS tropospheric modes rely on volume and anomalous for range extension in urban, rural, and over-water environments. Tropospheric scatter, a key NLOS mechanism, involves forward of radio waves by fluctuations in the , particularly effective at VHF (30-300 MHz) and UHF (300 MHz-3 GHz) bands for paths up to several hundred kilometers. This random reflection and from atmospheric turbulence allows transhorizon communication, with basic transmission modeled empirically to account for frequency, distance, and refractivity. via reflections contributes to NLOS by creating multiple signal paths, often dominated by a ray and a ground-reflected ray in the two-ray model. In this model, the path length difference \delta between the and reflected paths approximates \delta = \frac{2 h_t h_r}{d}, where h_t and h_r are transmitter and heights, and d is the separation distance; this difference leads to phase interference that causes , with received power falling off as d^4 at large distances. In urban environments, NLOS propagation encounters significant building penetration losses at UHF frequencies, typically ranging from 10-30 dB depending on wall materials like brick or concrete, which attenuate signals entering indoor spaces. Rural NLOS paths experience additional foliage attenuation, where vegetation layers cause exponential signal loss; extensions to the Okumura-Hata model incorporate these effects by adding correction factors for wooded areas, improving path loss predictions in suburban and forested terrains. Tropospheric ducting further enhances NLOS range through temperature inversions that form atmospheric waveguides, trapping signals via super-refraction where the gradient bends waves downward more sharply than normal. These ducts, often occurring in stable boundary layers over land or sea, can extend distances to hundreds of kilometers, enabling reliable beyond-horizon links at frequencies. Applications of non-ionospheric NLOS include urban systems, where multipath supports cellular coverage in cities despite obstacles, and links employing antennas to mitigate from reflections and ducting. or in these links selects or combines signals from multiple paths to maintain reliability over 20-50 km hops. Rain fade represents a critical impairment for microwave NLOS paths, with attenuation A approximated as A = \alpha f^{\beta} L, where \alpha and \beta are rain-specific coefficients dependent on f and , and L is the effective path length through ; this model predicts outages exceeding 0.01% of the time in zones.

Ionospheric Propagation

Ionospheric propagation refers to the transmission of radio waves through reflection and in the Earth's , a region of the upper atmosphere extending from approximately 50 to 1000 km altitude where solar radiation ionizes gas molecules, creating free s that interact with electromagnetic waves. This mode is particularly vital for high- (HF) signals, enabling beyond-line-of-sight communication over thousands of kilometers by bouncing signals off ionized layers, unlike direct ground-wave or line-of-sight paths. The ionosphere's ability to refract waves depends on the plasma , which arises from electron oscillations in the of the propagating wave. The is stratified into distinct layers based on altitude and ionization characteristics: the D layer (around 60-90 km), which is absorptive primarily during daytime due to higher densities from X-rays and radiation; the E layer (90-150 km), which provides sporadic reflection; and the , splitting into F1 (150-250 km) and (250-400 km) layers during daylight, both acting as primary reflective zones for signals. The f_c, the highest frequency reflected vertically back to from a layer, is given by f_c = 9 \sqrt{N_{\max}} in MHz, where N_{\max} is the maximum in electrons per cubic meter; for the layer, N_{\max} typically ranges from $10^{11} to $10^{12} m^{-3}, yielding f_c values of 5-15 MHz. Skywave propagation occurs via these reflections, with single-hop modes covering distances up to about 4000 km by one ionospheric bounce and ground reflection, while multi-hop modes extend global reach through repeated ionosphere-ground interactions, though signal increases with hops due to and . The maximum usable (MUF) for a given , the highest supporting reliable single-hop , is approximated as MUF = f_c / \cos \theta, where \theta is the angle of incidence to the ionosphere; for oblique paths, \theta exceeds the vertical case, reducing the effective MUF by the secant factor. Diurnal variations see the D layer forming only in daylight (disappearing at night), E layer peaking midday, and F2 layer maximum electron density often shifting to afternoon; seasonal effects include higher winter ionization in mid-latitudes due to geometric factors, while the 11-year sunspot cycle modulates N_{\max} by up to a factor of 10, with solar maximum enhancing reliability. Sudden ionospheric disturbances () from solar flares can abruptly increase D-layer , blacking out HF signals for minutes to hours. Key applications leverage these characteristics for shortwave broadcasting, which reaches international audiences via multi-hop F-layer reflections in the 3-30 MHz band, and operations that exploit variable conditions for global contacts. Oblique sounding, using transmitters and receivers separated by the path distance, measures real-time MUF and supports path prediction for reliable links. A historical extreme is the 1859 , a massive that induced geomagnetic currents disrupting telegraph systems worldwide, rendering lines inoperable and sparking fires, demonstrating ionospheric extremes' potential impact on early wired communications.

Influencing Factors

Frequency Dependence

Radio propagation characteristics exhibit significant variation with frequency, influencing the dominant modes, range, and reliability across different bands. At lower frequencies, such as (ELF) and (VLF), signals experience minimal atmospheric and excellent ground due to their long wavelengths, enabling reliable over-the-horizon via ground waves. This contrasts with higher frequencies, where line-of-sight (LOS) constraints tighten and environmental absorptions intensify, shifting reliance toward or direct paths. In the ELF (3–30 Hz) and VLF (3–30 kHz) bands, propagation benefits from low attenuation over conductive surfaces like , with ground waves capable of encircling the globe with minimal loss, typically under 10 dB per 1000 km under optimal conditions. These frequencies penetrate and effectively, up to several hundred meters depending on , making them suitable for subsurface applications such as communication. A prominent example is the NIST station at 60 kHz, which broadcasts time signals using ground-wave to achieve nationwide coverage in with low absorption losses, supporting of clocks and devices over distances exceeding 1500 km. For medium frequency (MF, 300 kHz–3 MHz) and high frequency (HF, 3–30 MHz) bands, propagation is predominantly influenced by ionospheric interactions, with the D-layer causing significant absorption during daylight hours. Absorption peaks around 5–10 MHz due to increased electron density from solar radiation, attenuating signals by 20–50 dB or more on trans-equatorial paths, severely limiting shortwave communications below this range during the day. At night, when the D-layer dissipates, sky-wave reflection enables long-distance propagation up to several thousand kilometers, though diurnal variations remain pronounced. In very high frequency (VHF, 30–300 MHz), ultra high frequency (UHF, 300 MHz–3 GHz), and super high frequency (SHF, 3–30 GHz) bands, propagation increasingly adheres to LOS limitations, with diffraction and multipath fading becoming negligible beyond a few tens of kilometers. Atmospheric absorption rises notably, with oxygen contributing a broad peak at approximately 60 GHz (up to 15 dB/km at sea level) and water vapor peaking around 22 GHz (about 0.1–1 dB/km depending on humidity). Rain further exacerbates attenuation in these bands, scaling roughly with frequency squared above 10 GHz, potentially exceeding 10 dB/km in heavy downpours at SHF, which constrains terrestrial microwave links but is mitigated in satellite applications through adaptive techniques. Quantitative trends in attenuation highlight these shifts: in free space, path loss \alpha follows \alpha \propto f^2, where f is , arising from the in the , resulting in a 6 dB increase per frequency doubling. For ground-wave modes below 30 MHz, attenuation behaves inversely, with \alpha decreasing as f lowers due to reduced surface wave losses over imperfectly conducting earth, enabling greater ranges at ELF/VLF compared to MF. These frequency dependencies drive band-specific applications: low frequencies like LF (30–300 kHz) excel in navigation systems such as , leveraging stable ground-wave propagation over 1000–2000 km with minimal ionospheric interference for precise hyperbolic positioning in maritime and contexts. Conversely, SHF bands above 10 GHz suit communications, where ionospheric effects become negligible ( <0.1 dB), allowing high-data-rate links with gaseous losses under 1 dB total for zenith paths outside absorption peaks, despite heightened rain vulnerability.

Atmospheric and Environmental Effects

The troposphere significantly influences radio wave propagation through variations in the refractive index, which causes bending or refraction of signals. The atmospheric radio refractive index n is given by n = 1 + N \times 10^{-6}, where N is the refractivity, computed as N = 77.6 \frac{P}{T} + 3.75 \times 10^5 \frac{e}{T^2}, with P as total atmospheric pressure in hPa, T as absolute temperature in K, and e as water vapor pressure in hPa. This gradient in n typically bends waves toward the Earth, extending beyond line-of-sight ranges, but anomalous gradients can lead to super-refraction or ducting. Turbulence in the troposphere induces scintillation, causing rapid fluctuations in signal amplitude and phase, particularly at microwave frequencies above 10 GHz, due to refractive index irregularities on scales of centimeters to meters. Precipitation, including rain, snow, and ice, attenuates radio waves by scattering and absorption, with effects most pronounced at frequencies above 5 GHz. The specific attenuation due to rain \gamma_R (in dB/km) follows the power-law relation \gamma_R = k R^\alpha, where R is the rain rate in mm/h, and k and \alpha are frequency- and polarization-dependent coefficients; for horizontal polarization at frequencies around 10-20 GHz, k approximates 0.1 f^{0.85} for moderate rain rates, leading to losses of several dB/km in heavy rain. Snow and ice particles cause less severe attenuation than rain due to their lower dielectric constant and larger, less efficient scattering sizes, typically resulting in 20-50% lower losses at the same equivalent water content. Solar and geomagnetic activity profoundly affect propagation, particularly through enhanced ionization. Auroral phenomena, driven by charged particle precipitation into the polar atmosphere, create ionized patches that scatter VHF signals (30-300 MHz), enabling sporadic long-distance propagation via auroral reflection but also causing fading and noise on affected paths. During solar flares, increased X-ray flux intensifies D-layer ionization, leading to non-deviative absorption that scales inversely with the square of frequency (\propto f^{-2}), severely attenuating HF signals below 15 MHz for durations of minutes to hours, with greater impact at lower frequencies. Variations in climate zones alter the likelihood of tropospheric ducting, where stable refractive index layers trap waves. Equatorial regions exhibit higher ducting probabilities (up to 30-40% annually) due to persistent evaporation ducts from warm, moist trade winds and convection, enhancing microwave propagation over sea paths. In contrast, polar regions show elevated ducting during winter nights (10-20% probability), driven by temperature inversions and radiative cooling, though overall rates are lower than in equatorial zones owing to drier conditions and stronger winds. During solar maxima, such as cycle 23 around 2000, heightened solar flux elevates the maximum usable frequency (MUF) for HF skywave propagation on mid-latitude paths, improving higher-band reliability but exacerbating disruptions from frequent flares and geomagnetic storms that caused widespread blackouts.

Terrain and Obstacle Effects

Terrain and obstacle effects significantly alter radio propagation paths by introducing diffraction, scattering, reflection, and absorption, particularly in non-line-of-sight (NLOS) scenarios where Earth's surface features obstruct direct waves. These effects are prominent in hilly, urban, rural, and vegetated environments, leading to increased path loss and multipath interference that must be accounted for in system design. In rural and hilly terrains, prominent obstacles like hills cause diffraction losses that can range from 10 to 40 dB depending on geometry and frequency, often modeled using methods that approximate irregular profiles. Urban settings introduce clutter from buildings, creating confined propagation channels with pronounced multipath, while vegetation adds attenuation that varies with density and season. Understanding these impacts is essential for applications such as cellular base station placement, where site-specific assessments mitigate coverage gaps. Diffraction over hills is a key mechanism in irregular terrain, where radio waves bend around obstacles to reach shadowed areas. The Bullington method approximates such terrain by identifying the dominant obstruction and treating it as an equivalent single knife-edge diffraction point, simplifying calculations for multiple irregular profiles along the path. This approach calculates loss by combining knife-edge diffraction with corrections for path curvature and distance, yielding typical losses of 10 to 40 dB for trans-horizon paths obstructed by hills. In rural environments, diffraction effects differ between sharp knife-edge obstacles and rounded hills. Knife-edge models assume an infinitely thin barrier, using the Fresnel diffraction parameter \nu = h \sqrt{\frac{2(d_1 + d_2)}{\lambda d_1 d_2}}, where h is the obstacle height relative to the line-of-sight, \lambda is the wavelength, and d_1, d_2 are transmitter-to-obstacle and obstacle-to-receiver distances. The diffraction loss J(\nu) is derived from Fresnel integrals involving the complex function V(\nu) = \int_0^\nu e^{j \pi t^2 / 2} dt, approximated for practical computation. For rounded hills, additional curvature loss is added, as the finite radius reduces bending efficiency compared to a knife-edge; for example, a rounded profile with radius R introduces a term T(m,n) that can increase total loss by several dB over knife-edge predictions. Vegetation, such as trees and foliage, causes absorption and scattering, particularly at UHF frequencies, with penetration loss proportional to depth and frequency. The total excess attenuation through dense foliage is approximated by A = 0.2 f^{0.3} d^{0.6} \quad \text{(dB)}, for penetration depths d < 400 m (with f in MHz), applicable for paths through leaves up to 400 m depth. Seasonal variations occur due to leaf density, with in-leaf conditions increasing attenuation by about 20% compared to leafless trees at around 1 GHz, based on measurements in mixed forests. In urban environments, clutter from buildings creates street canyon effects, where signals propagate via involving reflections off walls and diffraction over rooftops. Street canyons act as waveguides, confining waves and generating multiple delayed paths that cause ; models incorporate building height statistics, such as mean height h_b (typically 10-30 m in residential areas) and standard deviation \sigma_h (around 5-10 m), to predict over-rooftop loss and angular spread. These statistics, derived from urban morphology data, enable empirical adjustments for in non-line-of-sight urban links. Practical applications of these effects include site surveys for base stations, where terrain analysis identifies diffraction zones to optimize antenna placement and minimize outage. For instance, in a knife-edge scenario with 50% blockage of the first Fresnel zone, diffraction loss approximates 15 dB, guiding height adjustments to restore line-of-sight clearance and reduce signal degradation.

Modeling and Prediction

Theoretical Models

Theoretical models for radio propagation derive from fundamental principles of electromagnetism, primarily , to predict wave behavior in idealized or complex environments. These models provide analytical or numerical solutions for field distributions, enabling the calculation of signal strength, path loss, and interference without relying on experimental data fits. Unlike empirical approaches, they emphasize physical mechanisms such as reflection, refraction, diffraction, and scattering, often under high-frequency approximations where wavelengths are small compared to environmental features. Ray-tracing is a cornerstone theoretical method based on geometric optics, treating radio waves as rays that propagate in straight lines until interacting with boundaries or gradients. In line-of-sight (LOS) scenarios, rays follow the free-space path, while multipath effects arise from reflections off surfaces modeled using the image principle. Refraction occurs due to refractive index variations in the atmosphere, governed by , which states that the ratio of sines of incidence and refraction angles equals the ratio of refractive indices: \sin \theta_i / \sin \theta_r = n_2 / n_1. This law, adapted from optics to radio frequencies, allows tracing curved ray paths in layered media, such as tropospheric ducts. Seminal work by Keller extended geometric optics to include diffraction, laying the foundation for hybrid ray-based models. The parabolic equation (PE) method addresses propagation over irregular terrain and through inhomogeneous atmospheres by numerically solving a simplified form of the wave equation. Derived from the scalar under the narrow-angle paraxial approximation, it assumes forward-propagating waves and reduces the two-dimensional problem to a one-way marching algorithm, often implemented via the split-step for efficiency. The method excels in modeling terrain scattering and atmospheric refraction for frequencies from VHF to microwave bands, capturing effects like shadowing and ducting without ray singularities. The foundational , originally for surface waves, was adapted for broader propagation problems, providing a versatile tool for terrain-limited scenarios. The Geometrical Theory of Diffraction (GTD) and its uniform extension, the Uniform Theory of Diffraction (UTD), model wave bending around edges, corners, and obstacles, crucial for non-line-of-sight propagation. GTD introduces diffracted rays emanating from diffraction points, with diffraction coefficients derived from canonical problems like wedge scattering. Post-diffraction field strength decays as E \sim 1/\sqrt{s}, where s is the distance from the diffraction point, reflecting cylindrical wavefront spreading. UTD refines GTD by ensuring continuity across shadow and reflection boundaries, using Fresnel integrals to handle transition regions where standard GTD fails. Developed by Keller for general diffraction and extended by Kouyoumjian and Pathak for uniform asymptotics, these theories are applied at high frequencies to predict coverage in urban or obstructed environments. Full-wave solutions provide exact treatments for ground-wave propagation, solving boundary value problems for waves over a planar or spherical . The Sommerfeld integral represents the exact solution for a vertical dipole over a homogeneous lossy ground, expressing the field as an integral over plane waves that accounts for both direct and surface modes. This integral, while analytically intractable, is approximated numerically or asymptotically (e.g., via saddle-point methods) for practical computation, capturing attenuation due to ground conductivity and permittivity. Sommerfeld's 1909 formulation remains the benchmark for low-frequency ground waves (below 30 MHz), influencing standards for medium-wave broadcasting. These theoretical models, while physically rigorous, face limitations in computational demands, particularly for full-wave and PE methods requiring fine grids over large areas, often limiting real-time applications. They are most valid at high frequencies (typically f > 30 MHz), where asymptotic approximations hold, but break down at lower frequencies or in highly reverberant settings without extensions.

Empirical Models

Empirical models for radio propagation are derived from extensive field measurements and statistical analyses, providing practical predictions for in real-world environments where theoretical approximations fall short. These models prioritize accuracy in , suburban, and irregular scenarios by fitting curves to collected , often incorporating environmental without relying on physical derivations. They are widely used in wireless system design, spectrum allocation, and due to their computational efficiency and validated performance against measurements. The Okumura-Hata model, one of the earliest and most influential empirical formulations, was developed from drive tests conducted in during the , capturing median for land mobile radio services in urban and suburban areas. It applies to frequencies from 150 MHz to 1500 MHz, base station heights of 30 to 200 m, and mobile heights of 1 to 10 m, with distances up to 20 km. The PL in is given by: PL = 69.55 + 26.16 \log_{10} f - 13.82 \log_{10} h_b + (44.9 - 6.55 \log_{10} h_b) \log_{10} d - A_{hm} where f is the in MHz, h_b is the antenna height in meters, d is the link distance in km, and A_{hm} is a mobile antenna height correction factor that varies by environment (e.g., or suburban). This model has been extended in the late 2010s and 2020s for millimeter-wave bands through frequency scaling and additional loss terms, enabling predictions up to 30 GHz in microcells. The Longley-Rice model, also known as the Irregular Terrain Model (ITM), addresses propagation over varied topography, covering frequencies from 20 MHz to 20 GHz for point-to-point links up to 2000 km. It computes transmission loss relative to free space by integrating profile data, atmospheric refractivity, and clutter categories such as open area, urban, or forested, which add discrete loss values (e.g., 10-30 dB for ). The model uses a hybrid approach combining flat-earth , knife-edge losses, and empirical adjustments for irregular , making it suitable for broadcast and microwave planning. The COST-231 model extends the Hata formulation to higher frequencies (up to 2 GHz) for (WCDMA) systems in European urban environments, incorporating corrections for street widths and building heights. Its path loss equation modifies the Okumura-Hata base with an additional term C_m = 3 \log_{10} f - 4.97 for metropolitan areas, and includes optional add-ons for building penetration losses (typically 10-20 depending on wall materials). In recent years, (ML) techniques have emerged as a powerful extension to empirical modeling, leveraging large datasets from field measurements, simulations, or crowdsourced signals to predict , coverage, and channel characteristics in complex scenarios. These data-driven models, including neural networks, random forests, and (LSTM) networks, train on features like , , and environmental data to outperform traditional empirical models in and dynamic environments, with root-mean-square errors reduced by 20-50% in some applications as of 2024. Surveys highlight their integration with ray-tracing for hybrid predictions, supporting real-time optimization in beyond-5G networks. These models exhibit root-mean-square errors of approximately 5-10 when validated against field measurements in typical scenarios, reflecting their statistical nature and sensitivity to unmodeled variations like foliage or weather. The Longley-Rice ITM, in particular, is mandated for coverage predictions in U.S. (FCC) licensing processes for and land services.

Ionospheric Prediction Techniques

Ionospheric prediction techniques enable forecasting of radio wave propagation conditions in the , particularly for high-frequency () communications, by estimating key parameters such as critical frequencies, maximum usable frequencies (MUF), and signal-to-noise ratios (SNR). These methods rely on empirical data from soundings, observations, and numerical models to provide both short-term updates and longer-term climatological predictions, aiding in the optimization of paths and frequencies. Vertical and sounding systems offer direct measurements, while models like the International Reference Ionosphere (IRI) and propagation software integrate these data for broader applicability. Ionosondes perform vertical incidence by transmitting pulsed radio signals upward from a fixed and recording their reflections as ionograms, which are graphical traces displaying virtual height versus . From these ionograms, the (f_c), defined as the highest reflected vertically at a given ionospheric layer, is manually or automatically scaled, providing insights into profiles. For instance, the foF2 for the F2 layer is extracted to assess reflection capabilities for signals. This technique has been foundational since the mid-20th century, with modern digital ionosondes like the Digisonde enabling automated scaling and dissemination every few minutes. Oblique sounding extends prediction to specific propagation paths by using separated transmitter-receiver pairs, typically hundreds to thousands of kilometers apart, to measure real-time ionospheric parameters such as the MUF—the highest frequency supporting reliable one-hop propagation—and associated skip distances, which represent the minimum ground range for signal reflection. Ionograms from oblique sounders reveal trace delays and frequencies, allowing calculation of skip distances via geometric models relating path length to reflection height. Systems like the International Real-Time Assimilative Model (IRTAM) incorporate oblique data for enhanced MUF forecasting over mid-latitude regions, achieving accuracies within 10-20% of observed values during varying solar conditions. The IRI model serves as an international standard empirical representation of the , generating N(h) profiles across altitudes from 60 to 2000 km based on aggregated ground-based data, in-situ measurements, and incoherent scatter observations. IRI provides monthly median values for parameters including peak (NmF2) and its height (hmF2), enabling predictions of ionospheric for HF paths under average conditions. Updates to IRI, such as the 2016 and 2020 versions, incorporate storm-time corrections and topside profiles from satellites like COSMIC, improving N(h) accuracy to within 15-20% globally. Software tools like VOACAP, developed by the U.S. (NTIA), predict HF circuit performance by integrating ionospheric maps from the Comité Consultatif International des Radiocommunications (CCIR, now ) with user-specified parameters such as transmitter power, antenna gains, and solar activity indices. VOACAP computes median hourly SNR values for point-to-point paths, factoring in , , and noise, with outputs indicating reliable frequencies and bearing angles; for example, it forecasts SNR exceeding 10 dB for 90% of days on a 3000 km path during . Similarly, HF beacon networks, such as those using continuous transmissions for , support updates to these models by assimilating signal data into prediction algorithms. Recent advancements leverage Global Navigation Satellite Systems (GNSS) to derive (TEC)—the integrated column density of electrons along signal paths—for ionospheric corrections, with maps updated every 5-15 minutes from global receiver networks. GNSS-based TEC products, like those from the Laboratory's system, enable nowcasting of vertical TEC variations with root-mean-square errors below 5 TEC units, facilitating immediate adjustments for HF propagation delays and scintillations. Post-2020, artificial intelligence techniques, particularly deep learning models such as (LSTM) networks, have enhanced predictions by assimilating historical ionosonde and GNSS data to forecast TEC and foF2 up to 24 hours ahead, achieving improvements of 20-30% in accuracy over traditional empirical methods during geomagnetic storms.

Measurement and Applications

Propagation Measurement Methods

Field strength meters are essential instruments for measuring the intensity of radio signals along propagation paths, typically consisting of portable receivers equipped with loop antennas to detect components for directional sensitivity. These devices quantify signal levels in units such as dBμV, allowing for systematic logging of field variations over distance or time to assess and coverage in various environments. Calibration procedures ensure linearity across frequencies, as outlined in standards for professional measurements. Direction finding techniques complement field strength measurements by determining the bearing of incoming signals, often employing Adcock arrays—configurations of four vertical antennas arranged in a square to form orthogonal pairs for amplitude comparison and phase analysis. This setup, based on the Watson-Watt method, enables precise angle-of-arrival estimation with minimal ambiguity in the horizontal plane, particularly useful for low-frequency studies. Multiple direction finders at separated sites can then perform to locate signal sources or map anomalies, as demonstrated in tactical radio systems and research. Satellite beacons provide a global perspective on ionospheric propagation by transmitting continuous signals in VHF and UHF bands, which ground receivers use to monitor scintillation—rapid fluctuations in amplitude and phase caused by electron density irregularities. These beacons also facilitate total electron content (TEC) estimation through differential Doppler shift analysis between dual-frequency signals, revealing integrated electron densities along the ray path and aiding in the study of ionospheric dynamics. For HF propagation, ionosondes offer complementary vertical sounding measurements of reflection heights and critical frequencies. Drive tests in cellular networks involve vehicle-mounted equipment that logs signal parameters like (RSSI) alongside GPS positions, capturing spatiotemporal variations in due to and multipath effects. Post-processing includes statistical tools such as cumulative distribution functions (CDFs) of RSSI to evaluate coverage reliability and statistics, informing network optimization without relying on simulations. Pioneering global ionospheric measurements began with the satellite series in the , which transmitted beacon signals from low-Earth orbit to map electron content variations worldwide using Faraday rotation and Doppler techniques. Today, software-defined radios (SDRs) have democratized HF propagation monitoring by enabling flexible, low-cost reception and analysis of skywave signals through digital signal processing on commodity hardware.

Practical Effects on Systems

Radio propagation variations significantly impact the performance of communication and radar systems by introducing signal fluctuations, , and errors that degrade reliability and range. In communication systems, manifests in two primary forms: fast fading, caused by with coherence times on the order of milliseconds, and slow fading, or shadowing, due to obstacles with variations occurring over seconds. Fast fading leads to rapid signal fluctuations, potentially causing bit errors in transmissions, while slow fading results in longer-term signal that can interrupt service over extended periods. To counteract these effects, diversity techniques are employed, including space diversity (using multiple antennas), time diversity (repeating transmissions), and frequency diversity (switching carriers). These methods exploit the uncorrelated nature of fading channels to improve signal reliability by selecting or combining the strongest paths, often achieving diversity gains of several decibels. In receiver design, interference from propagation phenomena, such as co-channel skywave signals in medium and high-frequency bands, can overlap with desired groundwave signals, reducing signal-to-interference ratios and causing crosstalk in multi-user networks. Additionally, strong out-of-band signals can lead to receiver desensitization, where the dynamic range is compressed, elevating the noise floor and impairing weak signal detection by up to 20-30 dB in severe cases. System reliability is quantified through metrics like outage probability, defined as the likelihood that the instantaneous (BER) exceeds a , such as 10^{-5} for services, ensuring the link maintains acceptable performance. Link budgets incorporate margins of 10-20 to buffer against and other losses, allowing systems to operate below nominal conditions without outage; for instance, a 15 margin might limit outage to 1% in shadowed urban environments. In systems, propagation effects exacerbate clutter from ground and sea surfaces, where multipath reflections create false echoes that mask true , increasing rates. Super-refraction, often due to atmospheric inversions, bends beams toward the surface, generating anomalous that produces false at extended ranges, sometimes displacing apparent positions by kilometers and complicating range estimation. Mitigation strategies at the system level include adaptive modulation schemes, such as those in (OFDM) used in and networks, which dynamically adjust constellation sizes (e.g., from QPSK to 64-QAM) based on channel conditions to maintain throughput amid . Forward error correction (FEC) codes, like turbo or low-density parity-check codes, add redundancy to detect and correct errors induced by fast , achieving coding gains of 3-6 dB while complementing diversity for slow scenarios. These techniques collectively enhance link robustness without requiring detailed propagation modeling.

Real-World Applications and Case Studies

In broadcasting, the has long relied on high-frequency () skywave propagation for global reach, scheduling transmissions based on the maximum usable frequency (MUF) to optimize signal reliability over varying distances and times of day. Monthly MUF maps for transmissions from the are prepared by the , accounting for ionospheric conditions to select frequencies below the predicted MUF, ensuring effective coverage while avoiding or losses. In the , the service began shifting toward () to improve audio quality and efficiency on HF bands, launching dedicated DRM channels for regions like in collaboration with , which allowed multiplexing multiple programs within narrower bandwidths compared to analog shortwave. This transition addressed propagation variability while maintaining robustness against , with DRM transmissions continuing alongside analog HF to support areas with unreliable digital infrastructure. In mobile networks, 5G millimeter-wave (mmWave) deployments face significant challenges in urban non-line-of-sight (NLOS) environments due to high from buildings and foliage, limiting coverage to short ranges often under 200 meters without mitigation. techniques, such as multi-beam antenna combining, have proven effective in overcoming these issues; for instance, in measurements at 28 GHz, coherent quad-beam combining provided up to 24.9 dB improvement over single random beams at 100-meter transmitter-receiver separations, extending viable NLOS coverage and increasing range by 41% compared to single-beam operation. These gains stem from exploiting multipath reflections in dense urban canyons, enabling practical deployments like those in where NLOS links achieved reliable connectivity up to 150 meters through adaptive . Aviation and military applications highlight the vulnerabilities of satellite-based systems to ionospheric disturbances. During the 2003 Halloween geomagnetic storms, intense solar flares from active region 486 elevated total electron content (TEC) above 250 TEC units over —far exceeding typical daytime levels of 100-130—causing GPS signal degradation and positioning errors. The wide-area augmentation system (WAAS) experienced vertical errors exceeding 50 meters for over 15 hours on , disrupting precision approaches for and forcing operations, such as drilling surveys, to revert to backups like acoustic positioning. In contrast, troposcatter systems provide resilient over-the-horizon communication for use, scattering UHF signals off atmospheric irregularities to achieve beyond-line-of-sight links up to 300 kilometers without dependency, as demonstrated in U.S. evaluations of beyond-line-of-sight (BLOS) technology for tactical networks in remote theaters. Disaster response efforts underscore propagation's role in crisis communications and early warning. Researchers have investigated using (VLF) signals to detect potential ionospheric perturbations preceding , as subionospheric (3-30 kHz) is sensitive to D-region changes possibly induced by seismic electromagnetic emissions or atmospheric gravity waves. For example, studies reported shifts in VLF terminator times prior to the 1995 (M7.3), indicating ionospheric lowering, while the 2004 event (M9.3) showed increased nighttime fluctuations 6-2 days before, detectable via amplitude anomalies at distant receivers. However, the lacks consensus on the reliability of such precursors for . During in 2005, communications failures were exacerbated by atmospheric environmental variability, including disrupted tropospheric conditions that affected VHF/UHF , leading to incompatible systems and widespread inoperability among despite infrastructure damage. Emerging technologies like are constrained by (THz) propagation limits, where frequencies above 100 GHz suffer severe atmospheric absorption from and oxygen, resulting in path losses exceeding 100 / even in clear conditions and restricting to tens of without advanced . Satellite mega-constellations, such as , mitigate these by operating in Ku-band (10-14 GHz) with line-of-sight (LOS) models that assume direct from low-Earth orbit satellites to users, enabling global coverage through frequent handovers and minimal multipath in rural or oceanic areas.

References

  1. [1]
    [PDF] Chapter 2: Radio Wave Propagation Fundamentals - KIT - IHE
    Nov 12, 2018 · Institute of Radio Frequency Engineering and Electronics. Propagation Phenomena. Chapter 2: Radio Wave Propagation Fundamentals. 12.11.2018. zT.
  2. [2]
    [PDF] Radio Wave Propagation
    • Fundamentals. • Propagation Over Irregular Terrain. • MSAM Propagation Models. –ITM: Irregular Terrain Model. –LMS: Land Mobile Services Model. • ITS Models.
  3. [3]
    [PDF] Introduction to Mobile Radio Propagation and Characterization of ...
    Re f lection occurs when a radio wave collides with an object which has very large dimensions compared to the wavelength of the propagating wave. Reflections ...
  4. [4]
    None
    ### Summary of Definitions and Frequency Range from ITU-R P.310-10
  5. [5]
    [PDF] Chapter 13 Maxwell's Equations and Electromagnetic Waves - MIT
    13.3 Maxwell's Equations. We now have four equations which form the foundation of electromagnetic phenomena: 13-5. Page 6. Law. Equation. Physical ...
  6. [6]
  7. [7]
    Frequency and Wavelength Calculations - A.H. Systems
    Frequency Wavelength Formula. Freq is the Frequency in cycles per second. C is the velocity factor 299,792,458 meters per second λ represents wavelength in ...
  8. [8]
    Free Space Path Loss Formula Derivation - Reference Designer
    Free Space Path loss is directly proportional to the square of the distance ... $$ FSPL = \left(\frac{4pd}{λ}\right)^2$$ $$= \left(\frac{4πdf}{c}\right)^2$$
  9. [9]
    Heinrich Hertz - Magnet Academy
    ### Summary of Hertz's Discovery of Electromagnetic Waves in the 1880s
  10. [10]
    First radio transmission sent across the Atlantic Ocean - History.com
    In fact, Marconi's transatlantic radio signal had been headed into space when it was reflected off the ionosphere and bounced back down toward Canada.
  11. [11]
    Radio Wave Diffraction - Electronics Notes
    It is found that diffraction is more pronounced when the obstacle becomes sharper and more like a "knife edge". For a radio signal the definition of a knife ...
  12. [12]
    Managing the radio-frequency spectrum for the world - ITU-R
    The MIFR is a database which contains the spectrum characteristics (“frequency assignments") of the radio stations in operation throughout the world and confers ...
  13. [13]
    [PDF] National Table of Frequency Allocations - ITU
    Nov 8, 2016 · These guidelines focus on the detailed preparation of a National Table of Frequency Allocation (NTFA). In addition it is providing a brief ...
  14. [14]
    Radiation Basics | US EPA
    Sep 10, 2025 · Non-ionizing radiation has enough energy to move atoms in a molecule around or cause them to vibrate, but not enough to remove electrons from ...
  15. [15]
    Isotropic Radiator - an overview | ScienceDirect Topics
    An isotropic radiator is a point source that radiates energy uniformly in all directions, with the same intensity regardless of direction.
  16. [16]
    [PDF] A Note. on a Simple TransmissionFormula*
    FRIISt, FELLOW, I.R.E.. Summary-A simple transmission formula for a radio circuit is derived. The utility of the formula is emphasized and its ...
  17. [17]
    Directivity and antenna gain - Radartutorial.eu
    The directivity of an antenna is the ratio of the power density S (radiant intensity per unit area) of the real antenna in its main direction.
  18. [18]
    Module 3_1: Basic Refraction Principles
    Refraction is the term used for the bending of radio waves. In outer space, with no atmosphere, radio waves propagate in straight lines.Missing: extension | Show results with:extension
  19. [19]
    [PDF] OT Report 78-144: Radio Propagation in Urban Areas
    This report reviews radio propagation in urban areas, including multipath fading, location variability, and a computer model for calculating attenuation.
  20. [20]
    Electromagnetic Surface Waves - MIT
    Jun 15, 1996 · Hill and J. R. Wait, ``Ground wave attenuation function for a spherical earth with arbitrary surface impedance,'' Radio Sci. 15, 637--643 ...<|separator|>
  21. [21]
    [PDF] Handbook on Ground Wave Propagation - Engenharia Eletrica - UFPR
    Within the Sommerfeld-Norton theory this extra attenuation is given by the term, F, see equation (5), where for ground-based terminals w simplifies to become:.
  22. [22]
    [PDF] Ground-wave Analysis Model for MF Broadcast Systems
    The surface impedance is a function of the ground constants of the Earth's surface ... where A is the Norton approximation (Norton, 1941) to the Sommerfeld ...
  23. [23]
    Effect of Ground Conductivity on VLF Wave Propagation - Teysseyre
    Mar 18, 2025 · Very Low Frequency (VLF, 3–30 kHz) waves propagate long distances in the waveguide formed by the Earth and the lower ionosphere.
  24. [24]
    The Theory of Loran-C Ground Wave Propagation-A Review
    The space wave is made up of the direct wave (a signal traveling between the transmitter and the receiver via the direct path) and the ground- reflected wave (a ...
  25. [25]
    VLF communications - Australian Space Academy
    Propagation over an extensive ice sheet involves high attenuation due to the low conductivity of ice. Above 10 kHz VLF noise decreases by 20 dB per decade.
  26. [26]
    [PDF] Signal Propagation and Path Loss Models
    Reflected rays have power falloff proportional to d2 by free space path loss model. Scattered and refracted rays have power falloff that depends on exact ...
  27. [27]
    (PDF) An Integrated Terrain and Clutter Propagation Model for 1.7 ...
    Feb 10, 2022 · measurement data. Considering the Okumura-Hata ... Example of modeling foliage attenuation along a propagation path. The extracted foliage ...
  28. [28]
    Atmospheric Ducting Effect in Wireless Communications: Challenges and Opportunities
    **Summary of Tropospheric Ducting from IEEE Document (9475120):**
  29. [29]
    Wireless line-of-sight, non-line-of-sight, beyond-line-of-sight ...
    Sep 11, 2018 · In short range microwave transmission, the multipath phenomenon is dealt with by using diversity antennas and complex algorithms to combine ...<|separator|>
  30. [30]
    [PDF] handbook the ionosphere and its effects on radiowave propagation
    Meteor scatter provides a means of propagation at VHF, both for communication and interference. Two-way radiocommunications circuits have been operated with ...
  31. [31]
    [PDF] Ionospheric Propagation
    This requires the frequency to be less than the critical frequency fc, given by. 81Nmax f2 c. = 1 ⇒ fc = 9 p. Nmax. (15). Prof. Sean Victor Hum. Radio and ...
  32. [32]
    [PDF] Optimal Estimation Inversion of Ionospheric Electron Density from ...
    Aug 16, 2023 · values, especially when the critical frequency of the F1 layer (foF1) is relatively closer to the critical frequency of the F2 layer (foF2).Missing: formula | Show results with:formula
  33. [33]
    Lowest & Maximum Usable Frequency, Critical Frequency
    It is possible to calculate the relationship more exactly: MUF=CFcosθ. Where: MUF = Maximum Usable Frequency CF = critical frequency θ = the angle of incidence.
  34. [34]
    Diurnal, Seasonal, and 11-yr Solar Cycle Variation Effects on the ...
    Jul 1, 2016 · The electron density is primarily driven by the solar irradiance incident upon the earth's atmosphere, as photons of sufficient energy from the ...Missing: sunspot | Show results with:sunspot
  35. [35]
    [PDF] Introduction To Ionospheric Sounding
    The peaks of these layers usually form between 70 and 300 km altitude and are identified by the letters D, E, F1 and. F2, in order of their altitude. 111. By ...
  36. [36]
    (PDF) VLF Waveguide Propagation: The Basics - ResearchGate
    Feb 4, 2016 · This paper provides a brief tutorial of the relevant propagation dependencies for medium to long VLF paths best understood in terms of waveguide mode theory.
  37. [37]
    On the generation of ELF/VLF waves for long‐distance propagation ...
    Jul 29, 2010 · In this paper, we describe a theoretical model describing the HF heating and ionospheric responses, followed by a full-wave calculation of ELF/VLF propagation.Missing: WWVB | Show results with:WWVB
  38. [38]
    None
    ### Summary of D-Region Absorption in HF Radio Propagation
  39. [39]
    D Region Absorption Predictions (D-RAP)
    The D-Region Absorption Prediction model is used as guidance to understand the HF radio degradation and blackouts this can cause.Missing: daytime | Show results with:daytime
  40. [40]
  41. [41]
    [PDF] Loran for Required Navigation Performance 0.3 - Stanford University
    Jan 24, 2003 · Loran, is an attractive candidate to provide backup services for GPS because of its complementary area navigation (RNAV), stratum 1 timing, ...
  42. [42]
    None
    ### Summary of Radio Refractive Index Formula from ITU-R P.453-14
  43. [43]
    Survey of Polar and Auroral Region Effects on HF Propagation
    The effects of the auroral and polar zone ionosphere on forward propagation in the HF through low-VHF (3 to ∼30 MHz) portion of the spectrum are reviewed.<|separator|>
  44. [44]
    A ducting climatology derived from the European Centre for Medium ...
    Sep 17, 2004 · High ducting probabilities (≈100%) are found off the west coasts of the Americas, Africa, and Australia in typical stratocumulus conditions.
  45. [45]
    High‐Frequency Communications Response to Solar Activity in ...
    Jan 11, 2019 · Solar flares and geomagnetic storms disrupted high-frequency (HF, 3–30 MHz) emergency radio communications in September 2017 HF propagation ...
  46. [46]
    None
    ### Summary of Bullington Method and Related Diffraction Concepts from ITU-R P.526-14
  47. [47]
    None
    ### Summary on Vegetation Attenuation and Seasonal Changes
  48. [48]
    [PDF] Geometrical theory of diffraction. | Semantic Scholar
    Geometrical theory of diffraction. · J. Keller · Published in Journal of the Optical… 1 February 1962 · Physics · Journal of the Optical Society of America.Missing: radio seminal
  49. [49]
    On the generalization of Snell's law - CJ Coleman
    Snell's law can be regarded as a first integral of the ray tracing equations that are a key component in the geometric optics solution of Maxwell's equations.Missing: seminal | Show results with:seminal
  50. [50]
    Leontovich, M.A. and Fock, V.A. (1946) Solution of the Problem of ...
    (1946) Solution of the Problem of Propagation of Electromagnetic Waves along the Earth's Surface by Method of Parabolic Equations. Journal of Physics—USSR, 10, ...Missing: radio seminal
  51. [51]
    High Frequency Ground Wave Propagation | Request PDF
    Aug 5, 2025 · In his seminal 1909 paper, Sommerfeld presented the first rigorous solution for a vertical Hertzian dipole above a lossy ground, expressed ...
  52. [52]
    Empirical formula for propagation loss in land mobile radio services
    Abstract: An empirical formula for propagation loss is derived from Okumura's report in order to put his propagation prediction method to computational use.
  53. [53]
    New Empirical Path Loss Model for 28 GHz and 38 GHz Millimeter ...
    Nov 1, 2018 · Our model enhanced the prediction of path loss by 10% when compared to the Okumura and by 15% when compared to the COST-Hata. Keywords: path ...
  54. [54]
    [PDF] A Guide to the Use of the ITS Irregular Terrain Model in the Area ...
    Most radio propagation models, especially the general purpose ones, can be characterized as being either a "point-to-point" model or an "area prediction" model.Missing: cellular | Show results with:cellular
  55. [55]
    [PDF] Improved COST 231-WI Model for Irregular Built-Up Areas - URSI
    Aug 29, 2020 · This work copes with the estimation of electromagnetic field generated by UHF base stations in peculiar urban scenarios. To account for.Missing: original | Show results with:original
  56. [56]
    [PDF] Comparison of Empirical Propagation Path Loss Models for Mobile ...
    Abstract. Empirical propagation models have found favor in both research and industrial communities owing to their speed.
  57. [57]
    [PDF] oet bulletin - Federal Communications Commission
    Oct 26, 2015 · The Longley-Rice Fortran code implementing the Longley-Rice model is used in the. FCC's TVStudy software. As the Longley-Rice Fortran code is ...Missing: licensing | Show results with:licensing
  58. [58]
    [PDF] calibration of commercial radio field-strength meters at the National ...
    A calibration consists in part of measuring the over-all linearity of the field-strength meter at one or more frequencies and radio-frequency ini)ut voltage.
  59. [59]
    47 CFR 73.686 -- Field strength measurements. - eCFR
    Take all measurements with a horizontally polarized antenna. Use a shielded transmission line between the testing antenna and the field strength meter. Match ...
  60. [60]
    [PDF] Review of Conventional Tactical Radio Direction Finding Systems
    The directional antennas used are either two of four element Adcock antennas. The omni-directional antenna may be a single dipole positioned at the center of ...
  61. [61]
    [PDF] radio direction finding
    Directive loop antennas in which the horizontal conductors are shielded or removed are called ADCOCK antennas. Ad- cock antenna systems are used primarily for.
  62. [62]
    [PDF] Propagation studies using direction-finding techniques
    In this paper two techniques are described for the study of. Illulticomponent signals. One involves use of pulse transmissions to effect "time-of-arrival".
  63. [63]
    Using TEC and radio scintillation data from the CITRIS radio beacon ...
    Oct 15, 2011 · This paper is both a brief review of the CITRIS experiment and the first combined TEC and scintillation study of ionospheric irregularities ...
  64. [64]
    Beacon satellite receiver for ionospheric tomography - AGU Journals
    Oct 2, 2014 · We introduce a new coherent dual-channel beacon satellite receiver intended for ionospheric tomography. The measurement equation includes ...
  65. [65]
    Global HF - Ionospheric Map - SWS
    The above map can be used as a guide to NVIS ionospheric frequency support and to generate real time HF predictions (eg Hourly HAP charts) to assist the HF ...
  66. [66]
  67. [67]
    [PDF] Impact of Indoor-Outdoor Context on Crowdsourcing based Mobile ...
    Fig. 3 shows the CDF of differences in average RSSI be- tween indoors and outdoors across all 661 cell sectors with a minimum of 150 measurements ...<|separator|>
  68. [68]
    [PDF] Performance Evaluation of HSPA Network Through Drive Testing ...
    RSSI [dBm] = RSCP [dBm] - Ec/Io [dB]. Figure 23 shows the RSSI coverage as resulted from the CDT test. Figure 23 RSSI distribution. Figure 24 illustrate the ...
  69. [69]
    Satellite beacon contributions to studies of the structure of the ...
    The signals transmitted by radio beacons on board artificial earth satellites have been widely used for studies of the earth's ionosphere.
  70. [70]
    [PDF] Recent Results from Satellite Beacon Measurements
    The first beacons designed for ionospheric propagation studies were placed on low orbiting satellites and yielded a nearly instantaneous picture of the electron ...
  71. [71]
    Software-defined radio-based HF doppler receiving system
    Nov 26, 2021 · The new digital system enables radio waves to be received with much smaller amplitudes at four different frequencies.
  72. [72]
    [PDF] Receiver Interference Immunity: Issues and Recommendations
    Aug 12, 2025 · Achievable receiver interference performance is limited by size, weight, power, and cost (SWAP- C) constraints that vary widely across systems.
  73. [73]
    [PDF] Link Budgets - People
    The link budget includes all gains (and losses) from baseband input to baseband output.
  74. [74]
    [PDF] PREDICTING CLUTTER DURING ANOMALOUS PROPAGATION ...
    Under anomalous propagation conditions, ground and sea clutter can be caused by superrefraction (when rays are bent down with a radius of curvature ...
  75. [75]
    [PDF] BBC Monograph no 43
    MUF maps of Europe for transmissions from the United King- dom are prepared monthly by the BBC (Fig. 7), and, if necessary, similar MUF maps of the world can ...
  76. [76]
    Press Office - New Digital Radio Mondiale channel for South Asia
    Oct 26, 2010 · BBC World Service and Deutsche Welle (DW) are launching a new Digital Radio Mondiale (DRM) digital radio channel for South Asia.Missing: shift | Show results with:shift
  77. [77]
    DRM Is Part of the BBC World Service Story
    May 4, 2022 · BBC World Service has been interested and supportive of the DRM standard for over 20 years. It maintains shortwave transmissions from Middle East and the UK or ...
  78. [78]
  79. [79]
    Qualcomm Research demonstrates robust mmWave design for 5G
    Nov 18, 2015 · These challenges made utilizing mmWave for mobile ... urban deployment in Manhattan with NLOS coverage of approximately 150 meters.Missing: NYC gain
  80. [80]
    [PDF] Intense Space Weather Storms, October 19 – November 07, 2003
    Oct 30, 2025 · The storms are suspected to have caused the loss of the $640 million ADEOS-2 spacecraft. On board the ADEOS-2 was the $150 million NASA.
  81. [81]
    Communications demo for U.S. Army showcases Raytheon BLOS ...
    Troposcatter, or tropospheric scatter, technology uses particles that make up the Earth's atmosphere as a reflector for microwave radio signals, with the ...
  82. [82]
    VLF/LF Radio Sounding of Ionospheric Perturbations Associated ...
    Jul 10, 2007 · It is recently recognized that the ionosphere is very sensitive to seismic effects, and the detection of ionospheric perturbations associated ...
  83. [83]
    [PDF] Communications - H. Rpt. 109-377 - A Failure of Initiative: Final Report
    Dec 14, 2005 · Massive inoperability—failed, destroyed, or incompatible communications systems—was the biggest communications problem in the response to.Missing: ducting variability
  84. [84]
    6G THz-Band Facing Propagation and Atmospheric Absorption Losses
    6G will work at the THz band, starting from 0.1THZ to 10 THz, supporting Tbps-level data rate, micro-second-level latency, and connection density of 10^7 per km ...
  85. [85]
    The Parameters Comparison of the “Starlink” LEO Satellites ...
    The Starlink constellation is planned to consist of thousands of small LEO satellites, deployed in three shells (layers), dedicated to maximize broadband ...<|control11|><|separator|>