Remote sensing

Last updated

Synthetic aperture radar image of Death Valley colored using polarimetry Death-valley-sar.jpg
Synthetic aperture radar image of Death Valley colored using polarimetry

Remote sensing is the acquisition of information about an object or phenomenon without making physical contact with the object, in contrast to in situ or on-site observation. The term is applied especially to acquiring information about Earth and other planets. Remote sensing is used in numerous fields, including geophysics, geography, land surveying and most Earth science disciplines (e.g. exploration geophysics, hydrology, ecology, meteorology, oceanography, glaciology, geology). It also has military, intelligence, commercial, economic, planning, and humanitarian applications, among others.

Contents

In current usage, the term remote sensing generally refers to the use of satellite- or aircraft-based sensor technologies to detect and classify objects on Earth. It includes the surface and the atmosphere and oceans, based on propagated signals (e.g. electromagnetic radiation). It may be split into "active" remote sensing (when a signal is emitted by a satellite or aircraft to the object and its reflection is detected by the sensor) and "passive" remote sensing (when the reflection of sunlight is detected by the sensor). [1] [2] [3] [4]

Overview

This video is about how Landsat was used to identify areas of conservation in the Democratic Republic of the Congo, and how it was used to help map an area called MLW in the north.

Remote sensing can be divided into two types of methods: Passive remote sensing and Active remote sensing. Passive sensors gather radiation that is emitted or reflected by the object or surrounding areas. Reflected sunlight is the most common source of radiation measured by passive sensors. Examples of passive remote sensors include film photography, infrared, charge-coupled devices, and radiometers. Active collection, on the other hand, emits energy in order to scan objects and areas whereupon a sensor then detects and measures the radiation that is reflected or backscattered from the target. RADAR and LiDAR are examples of active remote sensing where the time delay between emission and return is measured, establishing the location, speed and direction of an object.

Illustration of remote sensing Remote Sensing Illustration.jpg
Illustration of remote sensing

Remote sensing makes it possible to collect data of dangerous or inaccessible areas. Remote sensing applications include monitoring deforestation in areas such as the Amazon Basin, glacial features in Arctic and Antarctic regions, and depth sounding of coastal and ocean depths. Military collection during the Cold War made use of stand-off collection of data about dangerous border areas. Remote sensing also replaces costly and slow data collection on the ground, ensuring in the process that areas or objects are not disturbed.

Orbital platforms collect and transmit data from different parts of the electromagnetic spectrum, which in conjunction with larger scale aerial or ground-based sensing and analysis, provides researchers with enough information to monitor trends such as El Niño and other natural long and short term phenomena. Other uses include different areas of the earth sciences such as natural resource management, agricultural fields such as land usage and conservation, [5] [6] greenhouse gas monitoring, [7] oil spill detection and monitoring, [8] and national security and overhead, ground-based and stand-off collection on border areas. [9]

Types of data acquisition techniques

The basis for multispectral collection and analysis is that of examined areas or objects that reflect or emit radiation that stand out from surrounding areas. For a summary of major remote sensing satellite systems see the overview table.

Applications of remote sensing

Radar image of Aswan Dam, Egypt taken by Umbra 25cm radar image of Aswan Dam, Egypt from Umbra Lab Inc.tif
Radar image of Aswan Dam, Egypt taken by Umbra
Examples of remote sensing equipment deployed by
or interfaced with oceanographic research vessels. Deployment of oceanographic research vessels.png
Examples of remote sensing equipment deployed by
or interfaced with oceanographic research vessels.

Geodetic

Acoustic and near-acoustic

To coordinate a series of large-scale observations, most sensing systems depend on the following: platform location and the orientation of the sensor. High-end instruments now often use positional information from satellite navigation systems. The rotation and orientation are often provided within a degree or two with electronic compasses. Compasses can measure not just azimuth (i. e. degrees to magnetic north), but also altitude (degrees above the horizon), since the magnetic field curves into the Earth at different angles at different latitudes. More exact orientations require gyroscopic-aided orientation, periodically realigned by different methods including navigation from stars or known benchmarks.

Data characteristics

The quality of remote sensing data consists of its spatial, spectral, radiometric and temporal resolutions.

Spatial resolution
The size of a pixel that is recorded in a raster image – typically pixels may correspond to square areas ranging in side length from 1 to 1,000 metres (3.3 to 3,280.8 ft).
Spectral resolution
The wavelength of the different frequency bands recorded – usually, this is related to the number of frequency bands recorded by the platform. Current Landsat collection is that of seven bands, including several in the infrared spectrum, ranging from a spectral resolution of 0.7 to 2.1 μm. The Hyperion sensor on Earth Observing-1 resolves 220 bands from 0.4 to 2.5 μm, with a spectral resolution of 0.10 to 0.11 μm per band.
Radiometric resolution
The number of different intensities of radiation the sensor is able to distinguish. Typically, this ranges from 8 to 14 bits, corresponding to 256 levels of the gray scale and up to 16,384 intensities or "shades" of colour, in each band. It also depends on the instrument noise.
Temporal resolution
The frequency of flyovers by the satellite or plane, and is only relevant in time-series studies or those requiring an averaged or mosaic image as in deforesting monitoring. This was first used by the intelligence community where repeated coverage revealed changes in infrastructure, the deployment of units or the modification/introduction of equipment. Cloud cover over a given area or object makes it necessary to repeat the collection of said location.

Data processing

In order to create sensor-based maps, most remote sensing systems expect to extrapolate sensor data in relation to a reference point including distances between known points on the ground. This depends on the type of sensor used. For example, in conventional photographs, distances are accurate in the center of the image, with the distortion of measurements increasing the farther you get from the center. Another factor is that of the platen against which the film is pressed can cause severe errors when photographs are used to measure ground distances. The step in which this problem is resolved is called georeferencing and involves computer-aided matching of points in the image (typically 30 or more points per image) which is extrapolated with the use of an established benchmark, "warping" the image to produce accurate spatial data. As of the early 1990s, most satellite images are sold fully georeferenced.

In addition, images may need to be radiometrically and atmospherically corrected.

Radiometric correction
Allows avoidance of radiometric errors and distortions. The illumination of objects on the Earth's surface is uneven because of different properties of the relief. This factor is taken into account in the method of radiometric distortion correction. [28] Radiometric correction gives a scale to the pixel values, e. g. the monochromatic scale of 0 to 255 will be converted to actual radiance values.
Topographic correction (also called terrain correction)
In rugged mountains, as a result of terrain, the effective illumination of pixels varies considerably. In a remote sensing image, the pixel on the shady slope receives weak illumination and has a low radiance value, in contrast, the pixel on the sunny slope receives strong illumination and has a high radiance value. For the same object, the pixel radiance value on the shady slope will be different from that on the sunny slope. Additionally, different objects may have similar radiance values. These ambiguities seriously affected remote sensing image information extraction accuracy in mountainous areas. It became the main obstacle to the further application of remote sensing images. The purpose of topographic correction is to eliminate this effect, recovering the true reflectivity or radiance of objects in horizontal conditions. It is the premise of quantitative remote sensing application.
Atmospheric correction
Elimination of atmospheric haze by rescaling each frequency band so that its minimum value (usually realised in water bodies) corresponds to a pixel value of 0. The digitizing of data also makes it possible to manipulate the data by changing gray-scale values.

Interpretation is the critical process of making sense of the data. The first application was that of aerial photographic collection which used the following process; spatial measurement through the use of a light table in both conventional single or stereographic coverage, added skills such as the use of photogrammetry, the use of photomosaics, repeat coverage, Making use of objects' known dimensions in order to detect modifications. Image Analysis is the recently developed automated computer-aided application that is in increasing use.

Object-Based Image Analysis (OBIA) is a sub-discipline of GIScience devoted to partitioning remote sensing (RS) imagery into meaningful image-objects, and assessing their characteristics through spatial, spectral and temporal scale.

Old data from remote sensing is often valuable because it may provide the only long-term data for a large extent of geography. At the same time, the data is often complex to interpret, and bulky to store. Modern systems tend to store the data digitally, often with lossless compression. The difficulty with this approach is that the data is fragile, the format may be archaic, and the data may be easy to falsify. One of the best systems for archiving data series is as computer-generated machine-readable ultrafiche, usually in typefonts such as OCR-B, or as digitized half-tone images. Ultrafiches survive well in standard libraries, with lifetimes of several centuries. They can be created, copied, filed and retrieved by automated systems. They are about as compact as archival magnetic media, and yet can be read by human beings with minimal, standardized equipment.

Generally speaking, remote sensing works on the principle of the inverse problem : while the object or phenomenon of interest (the state) may not be directly measured, there exists some other variable that can be detected and measured (the observation) which may be related to the object of interest through a calculation. The common analogy given to describe this is trying to determine the type of animal from its footprints. For example, while it is impossible to directly measure temperatures in the upper atmosphere, it is possible to measure the spectral emissions from a known chemical species (such as carbon dioxide) in that region. The frequency of the emissions may then be related via thermodynamics to the temperature in that region.

Data processing levels

To facilitate the discussion of data processing in practice, several processing "levels" were first defined in 1986 by NASA as part of its Earth Observing System [29] and steadily adopted since then, both internally at NASA (e. g., [30] ) and elsewhere (e. g., [31] ); these definitions are:

LevelDescription
0Reconstructed, unprocessed instrument and payload data at full resolution, with any and all communications artifacts (e. g., synchronization frames, communications headers, duplicate data) removed.
1aReconstructed, unprocessed instrument data at full resolution, time-referenced, and annotated with ancillary information, including radiometric and geometric calibration coefficients and georeferencing parameters (e. g., platform ephemeris) computed and appended but not applied to the Level 0 data (or if applied, in a manner that level 0 is fully recoverable from level 1a data).
1bLevel 1a data that have been processed to sensor units (e. g., radar backscatter cross section, brightness temperature, etc.); not all instruments have Level 1b data; level 0 data is not recoverable from level 1b data.
2Derived geophysical variables (e. g., ocean wave height, soil moisture, ice concentration) at the same resolution and location as Level 1 source data.
3Variables mapped on uniform spacetime grid scales, usually with some completeness and consistency (e. g., missing points interpolated, complete regions mosaicked together from multiple orbits, etc.).
4Model output or results from analyses of lower level data (i. e., variables that were not measured by the instruments but instead are derived from these measurements).

A Level 1 data record is the most fundamental (i. e., highest reversible level) data record that has significant scientific utility, and is the foundation upon which all subsequent data sets are produced. Level 2 is the first level that is directly usable for most scientific applications; its value is much greater than the lower levels. Level 2 data sets tend to be less voluminous than Level 1 data because they have been reduced temporally, spatially, or spectrally. Level 3 data sets are generally smaller than lower level data sets and thus can be dealt with without incurring a great deal of data handling overhead. These data tend to be generally more useful for many applications. The regular spatial and temporal organization of Level 3 datasets makes it feasible to readily combine data from different sources.

While these processing levels are particularly suitable for typical satellite data processing pipelines, other data level vocabularies have been defined and may be appropriate for more heterogeneous workflows.

Applications

Satellite images provide very useful information to produce statistics on topics closely related to the territory, such as agriculture, forestry or land cover in general. The first large project to apply Landsata 1 images for statistics was LACIE (Large Area Crop Inventory Experiment), run by NASA, NOAA and the USDA in 1974–77. [32] [33] Many other application projects on crop area estimation have followed, including the Italian AGRIT project and the MARS project of the Joint Research Centre (JRC) of the European Commission. [34] Forest area and deforestation estimation have also been a frequent target of remote sensing projects, [35] [36] the same as land cover and land use [37]

Ground truth or reference data to train and validate image classification require a field survey if we are targetting annual crops or individual forest species, but may be substituted by photointerpretation if we look at wider classes that can be reliably identified on aerial photos or satellite images. It is relevant to highlight that probabilistic sampling is not critical for the selection of training pixels for image classification, but it is necessary for accuracy assessment of the classified images and area estimation. [38] [39] [40] Additional care is recommended to ensure that training and validation datasets are not spatially correlated. [41]

We suppose now that we have classified images or a land cover map produced by visual photo-interpretation, with a legend of mapped classes that suits our purpose, taking again the example of wheat. The straightforward approach is counting the number of pixels classified as wheat and multiplying by the area of each pixel. Many authors have noticed that estimator is that it is generally biased because commission and omission errors in a confusion matrix do not compensate each other [42] [43] [44]

The main strength of classified satellite images or other indicators computed on satellite images is providing cheap information on the whole target area or most of it. This information usually has a good correlation with the target variable (ground truth) that is usually expensive to observe in an unbiased and accurate way. Therefore it can be observed on a probabilistic sample selected on an area sampling frame. Traditional survey methodology provides different methods to combine accurate information on a sample with less accurate, but exhaustive, data for a covariable or proxy that is cheaper to collect. For agricultural statistics, field surveys are usually required, while photo-interpretation may better for land cover classes that can be reliably identified on aerial photographs or high resolution satellite images. Additional uncertainty can appear because of imperfect reference data (ground truth or similar). [45] [46]

Some options are: ratio estimator, regression estimator, [47] calibration estimators [48] and small area estimators [37]

If we target other variables, such as crop yield or leaf area, we may need different indicators to be computed from images, such as the NDVI, a good proxy to chlorophyll activity. [49]

History

The TR-1 reconnaissance/surveillance aircraft Usaf.u2.750pix.jpg
The TR-1 reconnaissance/surveillance aircraft
The 2001 Mars Odyssey used spectrometers and imagers to hunt for evidence of past or present water and volcanic activity on Mars. 2001 mars odyssey wizja.jpg
The 2001 Mars Odyssey used spectrometers and imagers to hunt for evidence of past or present water and volcanic activity on Mars.

The modern discipline of remote sensing arose with the development of flight. The balloonist G. Tournachon (alias Nadar) made photographs of Paris from his balloon in 1858. [50] Messenger pigeons, kites, rockets and unmanned balloons were also used for early images. With the exception of balloons, these first, individual images were not particularly useful for map making or for scientific purposes.

Systematic aerial photography was developed for military surveillance and reconnaissance purposes beginning in World War I. [51] After WWI, remote sensing technology was quickly adapted to civilian applications. [52] This is demonstrated by the first line of a 1941 textbook titled "Aerophotography and Aerosurverying," which stated the following:

"There is no longer any need to preach for aerial photography-not in the United States- for so widespread has become its use and so great its value that even the farmer who plants his fields in a remote corner of the country knows its value."

James Bagley, [52]

The development of remote sensing technology reached a climax during the Cold War with the use of modified combat aircraft such as the P-51, P-38, RB-66 and the F-4C, or specifically designed collection platforms such as the U2/TR-1, SR-71, A-5 and the OV-1 series both in overhead and stand-off collection. [53] A more recent development is that of increasingly smaller sensor pods such as those used by law enforcement and the military, in both manned and unmanned platforms. The advantage of this approach is that this requires minimal modification to a given airframe. Later imaging technologies would include infrared, conventional, Doppler and synthetic aperture radar. [54]

The development of artificial satellites in the latter half of the 20th century allowed remote sensing to progress to a global scale as of the end of the Cold War. [55] Instrumentation aboard various Earth observing and weather satellites such as Landsat, the Nimbus and more recent missions such as RADARSAT and UARS provided global measurements of various data for civil, research, and military purposes. Space probes to other planets have also provided the opportunity to conduct remote sensing studies in extraterrestrial environments, synthetic aperture radar aboard the Magellan spacecraft provided detailed topographic maps of Venus, while instruments aboard SOHO allowed studies to be performed on the Sun and the solar wind, just to name a few examples. [56] [57]

Recent developments include, beginning in the 1960s and 1970s, the development of image processing of satellite imagery. The use of the term "remote sensing" began in the early 1960s when Evelyn Pruitt realized that advances in science meant that aerial photography was no longer an adequate term to describe the data streams being generated by new technologies. [58] [59] With assistance from her fellow staff member at the Office of Naval Research, Walter Bailey, she coined the term "remote sensing". [60] [61] Several research groups in Silicon Valley including NASA Ames Research Center, GTE, and ESL Inc. developed Fourier transform techniques leading to the first notable enhancement of imagery data. In 1999 the first commercial satellite (IKONOS) collecting very high resolution imagery was launched. [62]

Training and education

Remote Sensing has a growing relevance in the modern information society. It represents a key technology as part of the aerospace industry and bears increasing economic relevance – new sensors e.g. TerraSAR-X and RapidEye are developed constantly and the demand for skilled labour is increasing steadily. Furthermore, remote sensing exceedingly influences everyday life, ranging from weather forecasts to reports on climate change or natural disasters. As an example, 80% of the German students use the services of Google Earth; in 2006 alone the software was downloaded 100 million times. But studies have shown that only a fraction of them know more about the data they are working with. [63] There exists a huge knowledge gap between the application and the understanding of satellite images. Remote sensing only plays a tangential role in schools, regardless of the political claims to strengthen the support for teaching on the subject. [64] A lot of the computer software explicitly developed for school lessons has not yet been implemented due to its complexity. Thereby, the subject is either not at all integrated into the curriculum or does not pass the step of an interpretation of analogue images. In fact, the subject of remote sensing requires a consolidation of physics and mathematics as well as competences in the fields of media and methods apart from the mere visual interpretation of satellite images.

Many teachers have great interest in the subject "remote sensing", being motivated to integrate this topic into teaching, provided that the curriculum is considered. In many cases, this encouragement fails because of confusing information. [65] In order to integrate remote sensing in a sustainable manner organizations like the EGU or Digital Earth [66] encourage the development of learning modules and learning portals. Examples include: FIS – Remote Sensing in School Lessons, [67] Geospektiv, [68] Ychange, [69] or Spatial Discovery, [70] to promote media and method qualifications as well as independent learning.

Software

Remote sensing data are processed and analyzed with computer software, known as a remote sensing application. A large number of proprietary and open source applications exist to process remote sensing data.

Remote Sensing with gamma rays

There are applications of gamma rays to mineral exploration through remote sensing. In 1972 more than two million dollars were spent on remote sensing applications with gamma rays to mineral exploration. Gamma rays are used to search for deposits of uranium. By observing radioactivity from potassium, porphyry copper deposits can be located. A high ratio of uranium to thorium has been found to be related to the presence of hydrothermal copper deposits. Radiation patterns have also been known to occur above oil and gas fields, but some of these patterns were thought to be due to surface soils instead of oil and gas. [71]

Satellites

Six Earth observation satellites comprising the A-train satellite constellation as of 2014. A-Train w-Time2013 Web.jpg
Six Earth observation satellites comprising the A-train satellite constellation as of 2014.

An Earth observation satellite or Earth remote sensing satellite is a satellite used or designed for Earth observation (EO) from orbit, including spy satellites and similar ones intended for non-military uses such as environmental monitoring, meteorology, cartography and others. The most common type are Earth imaging satellites, that take satellite images, analogous to aerial photographs; some EO satellites may perform remote sensing without forming pictures, such as in GNSS radio occultation.

The first occurrence of satellite remote sensing can be dated to the launch of the first artificial satellite, Sputnik 1, by the Soviet Union on October 4, 1957. [72] Sputnik 1 sent back radio signals, which scientists used to study the ionosphere. [73] The United States Army Ballistic Missile Agency launched the first American satellite, Explorer 1, for NASA's Jet Propulsion Laboratory on January 31, 1958. The information sent back from its radiation detector led to the discovery of the Earth's Van Allen radiation belts. [74] The TIROS-1 spacecraft, launched on April 1, 1960, as part of NASA's Television Infrared Observation Satellite (TIROS) program, sent back the first television footage of weather patterns to be taken from space. [72]

In 2008, more than 150 Earth observation satellites were in orbit, recording data with both passive and active sensors and acquiring more than 10 terabits of data daily. [72] By 2021, that total had grown to over 950, with the largest number of satellites operated by US-based company Planet Labs. [75]

Most Earth observation satellites carry instruments that should be operated at a relatively low altitude. Most orbit at altitudes above 500 to 600 kilometers (310 to 370 mi). Lower orbits have significant air-drag, which makes frequent orbit reboost maneuvers necessary. The Earth observation satellites ERS-1, ERS-2 and Envisat of European Space Agency as well as the MetOp spacecraft of EUMETSAT are all operated at altitudes of about 800 km (500 mi). The Proba-1, Proba-2 and SMOS spacecraft of European Space Agency are observing the Earth from an altitude of about 700 km (430 mi). The Earth observation satellites of UAE, DubaiSat-1 & DubaiSat-2 are also placed in Low Earth Orbits (LEO) orbits and providing satellite imagery of various parts of the Earth. [76] [77]

To get global coverage with a low orbit, a polar orbit is used. A low orbit will have an orbital period of roughly 100 minutes and the Earth will rotate around its polar axis about 25° between successive orbits. The ground track moves towards the west 25° each orbit, allowing a different section of the globe to be scanned with each orbit. Most are in Sun-synchronous orbits.

A geostationary orbit, at 36,000 km (22,000 mi), allows a satellite to hover over a constant spot on the earth since the orbital period at this altitude is 24 hours. This allows uninterrupted coverage of more than 1/3 of the Earth per satellite, so three satellites, spaced 120° apart, can cover the whole Earth. This type of orbit is mainly used for meteorological satellites.

See also

Related Research Articles

<span class="mw-page-title-main">Lidar</span> Method of spatial measurement using laser

Lidar is a method for determining ranges by targeting an object or a surface with a laser and measuring the time for the reflected light to return to the receiver. Lidar may operate in a fixed direction or it may scan multiple directions, in which case it is known as lidar scanning or 3D laser scanning, a special combination of 3-D scanning and laser scanning. Lidar has terrestrial, airborne, and mobile applications.

<span class="mw-page-title-main">Landsat program</span> American network of Earth-observing satellites for international research purposes

The Landsat program is the longest-running enterprise for acquisition of satellite imagery of Earth. It is a joint NASA / USGS program. On 23 July 1972, the Earth Resources Technology Satellite was launched. This was eventually renamed to Landsat 1 in 1975. The most recent, Landsat 9, was launched on 27 September 2021.

<span class="mw-page-title-main">Moderate Resolution Imaging Spectroradiometer</span> Payload imaging sensor

The Moderate Resolution Imaging Spectroradiometer (MODIS) is a satellite-based sensor used for earth and climate measurements. There are two MODIS sensors in Earth orbit: one on board the Terra satellite, launched by NASA in 1999; and one on board the Aqua satellite, launched in 2002. MODIS has now been replaced by the VIIRS, which first launched in 2011 aboard the Suomi NPP satellite.

<span class="mw-page-title-main">Satellite imagery</span> Images taken from an artificial satellite

Satellite images are images of Earth collected by imaging satellites operated by governments and businesses around the world. Satellite imaging companies sell images by licensing them to governments and businesses such as Apple Maps and Google Maps.

India's remote sensing program was developed with the idea of applying space technologies for the benefit of humankind and the development of the country. The program involved the development of three principal capabilities. The first was to design, build and launch satellites to a Sun-synchronous orbit. The second was to establish and operate ground stations for spacecraft control, data transfer along with data processing and archival. The third was to use the data obtained for various applications on the ground.

<span class="mw-page-title-main">Multispectral imaging</span> Capturing image data across multiple electromagnetic spectrum ranges

Multispectral imaging captures image data within specific wavelength ranges across the electromagnetic spectrum. The wavelengths may be separated by filters or detected with the use of instruments that are sensitive to particular wavelengths, including light from frequencies beyond the visible light range, i.e. infrared and ultra-violet. It can allow extraction of additional information the human eye fails to capture with its visible receptors for red, green and blue. It was originally developed for military target identification and reconnaissance. Early space-based imaging platforms incorporated multispectral imaging technology to map details of the Earth related to coastal boundaries, vegetation, and landforms. Multispectral imaging has also found use in document and painting analysis.

<span class="mw-page-title-main">MERIS</span>

MEdium Resolution Imaging Spectrometer (MERIS) was one of the main instruments on board the European Space Agency (ESA)'s Envisat platform. The sensor was in orbit from 2002 to 2012. ESA formally announced the end of Envisat's mission on 9 May 2012.

<span class="mw-page-title-main">Normalized difference vegetation index</span> Graphical indicator of remotely sensed live green vegetation

The normalized difference vegetation index (NDVI) is a widely-used metric for quantifying the health and density of vegetation using sensor data. It is calculated from spectrometric data at two specific bands: red and near-infrared. The spectrometric data is usually sourced from remote sensors, such as satellites.

<span class="mw-page-title-main">Advanced very-high-resolution radiometer</span>

The Advanced Very-High-Resolution Radiometer (AVHRR) instrument is a space-borne sensor that measures the reflectance of the Earth in five spectral bands that are relatively wide by today's standards. AVHRR instruments are or have been carried by the National Oceanic and Atmospheric Administration (NOAA) family of polar orbiting platforms (POES) and European MetOp satellites. The instrument scans several channels; two are centered on the red (0.6 micrometres) and near-infrared (0.9 micrometres) regions, a third one is located around 3.5 micrometres, and another two the thermal radiation emitted by the planet, around 11 and 12 micrometres.

<span class="mw-page-title-main">Interferometric synthetic-aperture radar</span> Geodesy and remote sensing technique

Interferometric synthetic aperture radar, abbreviated InSAR, is a radar technique used in geodesy and remote sensing. This geodetic method uses two or more synthetic aperture radar (SAR) images to generate maps of surface deformation or digital elevation, using differences in the phase of the waves returning to the satellite or aircraft. The technique can potentially measure millimetre-scale changes in deformation over spans of days to years. It has applications for geophysical monitoring of natural hazards, for example earthquakes, volcanoes and landslides, and in structural engineering, in particular monitoring of subsidence and structural stability.

<span class="mw-page-title-main">Ocean color</span> Explanation of the color of oceans and ocean color remote sensing

Ocean color is the branch of ocean optics that specifically studies the color of the water and information that can be gained from looking at variations in color. The color of the ocean, while mainly blue, actually varies from blue to green or even yellow, brown or red in some cases. This field of study developed alongside water remote sensing, so it is focused mainly on how color is measured by instruments.

<span class="mw-page-title-main">Landsat 8</span> American Earth-observing satellite launched in 2013 as part of the Landsat program

Landsat 8 is an American Earth observation satellite launched on 11 February 2013. It is the eighth satellite in the Landsat program; the seventh to reach orbit successfully. Originally called the Landsat Data Continuity Mission (LDCM), it is a collaboration between NASA and the United States Geological Survey (USGS). NASA Goddard Space Flight Center in Greenbelt, Maryland, provided development, mission systems engineering, and acquisition of the launch vehicle while the USGS provided for development of the ground systems and will conduct on-going mission operations. It comprises the camera of the Operational Land Imager (OLI) and the Thermal Infrared Sensor (TIRS), which can be used to study Earth surface temperature and is used to study global warming.

<span class="mw-page-title-main">Sentinel-3</span> Earth observation satellite series

Sentinel-3 is an Earth observation heavy satellite series developed by the European Space Agency as part of the Copernicus Programme. It currently consists of 2 satellites: Sentinel-3A and Sentinel-3B. After initial commissioning, each satellite was handed over to EUMETSAT for the routine operations phase of the mission. Two recurrent satellites— Sentinel-3C and Sentinel-3D— will follow in approximately 2025 and 2028 respectively to ensure continuity of the Sentinel-3 mission.

Multispectral remote sensing is the collection and analysis of reflected, emitted, or back-scattered energy from an object or an area of interest in multiple bands of regions of the electromagnetic spectrum. Subcategories of multispectral remote sensing include hyperspectral, in which hundreds of bands are collected and analyzed, and ultraspectral remote sensing where many hundreds of bands are used. The main purpose of multispectral imaging is the potential to classify the image using multispectral classification. This is a much faster method of image analysis than is possible by human interpretation.

Sea ice concentration is a useful variable for climate scientists and nautical navigators. It is defined as the area of sea ice relative to the total at a given point in the ocean. This article will deal primarily with its determination from remote sensing measurements.

<span class="mw-page-title-main">Gaofen</span> Chinese satellites

Gaofen is a series of Chinese high-resolution Earth imaging satellites launched as part of the China High-resolution Earth Observation System (CHEOS) program. CHEOS is a state-sponsored, civilian Earth-observation program used for agricultural, disaster, resource, and environmental monitoring. Proposed in 2006 and approved in 2010, the CHEOS program consists of the Gaofen series of space-based satellites, near-space and airborne systems such as airships and UAVs, ground systems that conduct data receipt, processing, calibration, and taskings, and a system of applications that fuse observation data with other sources to produce usable information and knowledge.

<span class="mw-page-title-main">Remote sensing in geology</span> Data acquisition method for earth sciences

Remote sensing is used in the geological sciences as a data acquisition method complementary to field observation, because it allows mapping of geological characteristics of regions without physical contact with the areas being explored. About one-fourth of the Earth's total surface area is exposed land where information is ready to be extracted from detailed earth observation via remote sensing. Remote sensing is conducted via detection of electromagnetic radiation by sensors. The radiation can be naturally sourced, or produced by machines and reflected off of the Earth surface. The electromagnetic radiation acts as an information carrier for two main variables. First, the intensities of reflectance at different wavelengths are detected, and plotted on a spectral reflectance curve. This spectral fingerprint is governed by the physio-chemical properties of the surface of the target object and therefore helps mineral identification and hence geological mapping, for example by hyperspectral imaging. Second, the two-way travel time of radiation from and back to the sensor can calculate the distance in active remote sensing systems, for example, Interferometric synthetic-aperture radar. This helps geomorphological studies of ground motion, and thus can illuminate deformations associated with landslides, earthquakes, etc.

Remote sensing in oceanography is a widely used observational technique which enables researchers to acquire data of a location without physically measuring at that location. Remote sensing in oceanography mostly refers to measuring properties of the ocean surface with sensors on satellites or planes, which compose an image of captured electromagnetic radiation. A remote sensing instrument can either receive radiation from the Earth’s surface (passive), whether reflected from the Sun or emitted, or send out radiation to the surface and catch the reflection (active). All remote sensing instruments carry a sensor to capture the intensity of the radiation at specific wavelength windows, to retrieve a spectral signature for every location. The physical and chemical state of the surface determines the emissivity and reflectance for all bands in the electromagnetic spectrum, linking the measurements to physical properties of the surface. Unlike passive instruments, active remote sensing instruments also measure the two-way travel time of the signal; which is used to calculate the distance between the sensor and the imaged surface. Remote sensing satellites often carry other instruments which keep track of their location and measure atmospheric conditions.

<span class="mw-page-title-main">Hyperspectral Imager for the Coastal Ocean</span> Observation sensor on the International Space Station

The Hyperspectral Imager for the Coastal Ocean (HICO) was a hyperspectral earth observation sensor that operated on the International Space Station (ISS) from 2009 to 2014. HICO collected hyperspectral satellite imagery of the Earth's surface from the ISS.

<span class="mw-page-title-main">Thermal remote sensing</span>

Thermal remote sensing is a branch of remote sensing in the thermal infrared region of the electromagnetic spectrum. Thermal radiation from ground objects is measured using a thermal band in satellite sensors.

References

  1. Schowengerdt, Robert A. (2007). Remote sensing: models and methods for image processing (3rd ed.). Academic Press. p. 2. ISBN   978-0-12-369407-2. Archived from the original on 1 May 2016. Retrieved 15 November 2015.
  2. Schott, John Robert (2007). Remote sensing: the image chain approach (2nd ed.). Oxford University Press. p. 1. ISBN   978-0-19-517817-3. Archived from the original on 24 April 2016. Retrieved 15 November 2015.
  3. Guo, Huadong; Huang, Qingni; Li, Xinwu; Sun, Zhongchang; Zhang, Ying (2013). "Spatiotemporal analysis of urban environment based on the vegetation–impervious surface–soil model" (PDF). Journal of Applied Remote Sensing . 8: 084597. Bibcode:2014JARS....8.4597G. doi: 10.1117/1.JRS.8.084597 . S2CID   28430037. Archived (PDF) from the original on 19 July 2018. Retrieved 27 October 2021.
  4. Liu, Jian Guo & Mason, Philippa J. (2009). Essential Image Processing for GIS and Remote Sensing. Wiley-Blackwell. p. 4. ISBN   978-0-470-51032-2 . Retrieved 2 April 2023.
  5. "Saving the monkeys". SPIE Professional. Archived from the original on 4 February 2016. Retrieved 1 January 2016.
  6. Howard, A.; et al. (19 August 2015). "Remote sensing and habitat mapping for bearded capuchin monkeys (Sapajus libidinosus): landscapes for the use of stone tools". Journal of Applied Remote Sensing. 9 (1): 096020. doi:10.1117/1.JRS.9.096020. S2CID   120031016.
  7. Innocenti, Fabrizio; Robinson, Rod; Gardiner, Tom; Finlayson, Andrew; Connor, Andy (2017). "Differential Absorption Lidar (DIAL) Measurements of Landfill Methane Emissions". Remote Sensing. 9 (9): 953. Bibcode:2017RemS....9..953I. doi: 10.3390/rs9090953 .
  8. C. Bayindir; J. D. Frost; C. F. Barnes (January 2018). "Assessment and enhancement of SAR noncoherent change detection of sea-surface oil spills". IEEE J. Ocean. Eng. 43 (1): 211–220. Bibcode:2018IJOE...43..211B. doi:10.1109/JOE.2017.2714818. S2CID   44706251.
  9. "Science@nasa - Technology: Remote Sensing". Archived from the original on 29 September 2006. Retrieved 18 February 2009.
  10. Zhao, Kaiguang; Suarez, Juan C; Garcia, Mariano; Hu, Tongxi; Wang, Cheng; Londo, Alexis (2018). "Utility of multitemporal lidar for forest and carbon monitoring: Tree growth, biomass dynamics, and carbon flux". Remote Sensing of Environment. 204: 883–897. Bibcode:2018RSEnv.204..883Z. doi:10.1016/j.rse.2017.09.007.
  11. Just Sit Right Back and You'll Hear a Tale, a Tale of a Plankton Trip Archived 10 August 2021 at the Wayback Machine NASA Earth Expeditions, 15 August 2018.
  12. Levin, Noam; Kyba, Christopher C.M.; Zhang, Qingling; Sánchez de Miguel, Alejandro; Román, Miguel O.; Li, Xi; Portnov, Boris A.; Molthan, Andrew L.; Jechow, Andreas; Miller, Steven D.; Wang, Zhuosen; Shrestha, Ranjay M.; Elvidge, Christopher D. (February 2020). "Remote sensing of night lights: A review and an outlook for the future". Remote Sensing of Environment. 237: 111443. Bibcode:2020RSEnv.23711443L. doi:10.1016/j.rse.2019.111443. hdl: 10871/40052 . S2CID   214254543.
  13. Corradino, Claudia; Ganci, Gaetana; Bilotta, Giuseppe; Cappello, Annalisa; Del Negro, Ciro; Fortuna, Luigi (January 2019). "Smart Decision Support Systems for Volcanic Applications". Energies. 12 (7): 1216. doi: 10.3390/en12071216 .
  14. Corradino, Claudia; Ganci, Gaetana; Cappello, Annalisa; Bilotta, Giuseppe; Hérault, Alexis; Del Negro, Ciro (January 2019). "Mapping Recent Lava Flows at Mount Etna Using Multispectral Sentinel-2 Images and Machine Learning Techniques". Remote Sensing. 11 (16): 1916. Bibcode:2019RemS...11.1916C. doi: 10.3390/rs11161916 .
  15. Goldberg, A.; Stann, B.; Gupta, N. (July 2003). "Multispectral, Hyperspectral, and Three-Dimensional Imaging Research at the U.S. Army Research Laboratory" (PDF). Proceedings of the International Conference on International Fusion [6th]. 1: 499–506.
  16. Makki, Ihab; Younes, Rafic; Francis, Clovis; Bianchi, Tiziano; Zucchetti, Massimo (1 February 2017). "A survey of landmine detection using hyperspectral imaging". ISPRS Journal of Photogrammetry and Remote Sensing. 124: 40–53. Bibcode:2017JPRS..124...40M. doi:10.1016/j.isprsjprs.2016.12.009. ISSN   0924-2716.
  17. Mills, J.P.; et al. (1997). "Photogrammetry from Archived Digital Imagery for Seal Monitoring". The Photogrammetric Record. 15 (89): 715–724. doi:10.1111/0031-868X.00080. S2CID   140189982.
  18. Twiss, S.D.; et al. (2001). "Topographic spatial characterisation of grey seal Halichoerus grypus breeding habitat at a sub-seal size spatial grain". Ecography. 24 (3): 257–266. doi: 10.1111/j.1600-0587.2001.tb00198.x .
  19. Stewart, J.E.; et al. (2014). "Finescale ecological niche modeling provides evidence that lactating gray seals (Halichoerus grypus) prefer access to fresh water in order to drink" (PDF). Marine Mammal Science. 30 (4): 1456–1472. Bibcode:2014MMamS..30.1456S. doi:10.1111/mms.12126. Archived (PDF) from the original on 13 July 2021. Retrieved 27 October 2021.
  20. Zhang, Chuanrong; Li, Xinba (September 2022). "Land Use and Land Cover Mapping in the Era of Big Data". Land. 11 (10): 1692. doi: 10.3390/land11101692 .
  21. "Begni G. Escadafal R. Fontannaz D. and Hong-Nga Nguyen A.-T. (2005). Remote sensing: a tool to monitor and assess desertification. Les dossiers thématiques du CSFD. Issue 2. 44 pp". Archived from the original on 26 May 2019. Retrieved 27 October 2021.
  22. Wang, Ran; Gamon, John A. (15 September 2019). "Remote sensing of terrestrial plant biodiversity". Remote Sensing of Environment. 231: 111218. Bibcode:2019RSEnv.23111218W. doi:10.1016/j.rse.2019.111218. ISSN   0034-4257. S2CID   197567301.
  23. Rocchini, Duccio; Boyd, Doreen S.; Féret, Jean-Baptiste; Foody, Giles M.; He, Kate S.; Lausch, Angela; Nagendra, Harini; Wegmann, Martin; Pettorelli, Nathalie (February 2016). Skidmore, Andrew; Chauvenet, Alienor (eds.). "Satellite remote sensing to monitor species diversity: potential and pitfalls". Remote Sensing in Ecology and Conservation. 2 (1): 25–36. Bibcode:2016RSEC....2...25R. doi:10.1002/rse2.9. hdl: 11585/720672 . ISSN   2056-3485. S2CID   59446258.
  24. Schweiger, Anna K.; Cavender-Bares, Jeannine; Townsend, Philip A.; Hobbie, Sarah E.; Madritch, Michael D.; Wang, Ran; Tilman, David; Gamon, John A. (June 2018). "Plant spectral diversity integrates functional and phylogenetic components of biodiversity and predicts ecosystem function". Nature Ecology & Evolution. 2 (6): 976–982. Bibcode:2018NatEE...2..976S. doi:10.1038/s41559-018-0551-1. ISSN   2397-334X. PMID   29760440. S2CID   256718584.
  25. Cerrejón, Carlos; Valeria, Osvaldo; Marchand, Philippe; Caners, Richard T.; Fenton, Nicole J. (18 February 2021). "No place to hide: Rare plant detection through remote sensing". Diversity and Distributions. 27 (6): 948–961. Bibcode:2021DivDi..27..948C. doi: 10.1111/ddi.13244 . ISSN   1366-9516. S2CID   233886263.
  26. Carfagna, E. (2005). "Using remote sensing for agricultural statistics". International Statistical Review. 73 (3): 389–404. doi:10.1111/j.1751-5823.2005.tb00155.x.
  27. "Geodetic Imaging". Archived from the original on 2 October 2016. Retrieved 29 September 2016.
  28. Grigoriev А.N. (2015). "Мethod of radiometric distortion correction of multispectral data for the earth remote sensing". Scientific and Technical Journal of Information Technologies, Mechanics and Optics. 15 (4): 595–602. doi: 10.17586/2226-1494-2015-15-4-595-602 .
  29. NASA (1986), Report of the EOS data panel, Earth Observing System, Data and Information System, Data Panel Report, Vol. IIa., NASA Technical Memorandum 87777, June 1986, 62 pp. Available at http://hdl.handle.net/2060/19860021622 Archived 27 October 2021 at the Wayback Machine
  30. C. L. Parkinson, A. Ward, M. D. King (Eds.) Earth Science Reference Handbook – A Guide to NASA's Earth Science Program and Earth Observing Satellite Missions, National Aeronautics and Space Administration Washington, D. C. Available at http://eospso.gsfc.nasa.gov/ftp_docs/2006ReferenceHandbook.pdf Archived 15 April 2010 at the Wayback Machine
  31. GRAS-SAF (2009), Product User Manual, GRAS Satellite Application Facility, Version 1.2.1, 31 March 2009. Available at http://www.grassaf.org/general-documents/products/grassaf_pum_v121.pdf Archived 26 July 2011 at the Wayback Machine
  32. Houston, A.H. "Use of satellite data in agricultural surveys". Communications in Statistics. Theory and Methods (23): 2857–2880.
  33. Allen, J.D. "A Look at the Remote Sensing Applications Program of the National Agricultural Statistics Service". Journal of Official Statistics. 6 (4): 393–409.
  34. Taylor, J (1997). Regional Crop Inventories in Europe Assisted by Remote Sensing: 1988-1993. Synthesis Report. Luxembourg: Office for Pubblications of the EC.
  35. Foody, G.M. (1994). "Estimation of tropical forest extent and regenerative stage using remotely sensed data". Journal of Biogeography. 21 (3): 223–244. Bibcode:1994JBiog..21..223F. doi:10.2307/2845527. JSTOR   2845527.
  36. Achard, F (2002). "Determination of deforestation rates of the world's humid tropical forests". Science. 297 (5583): 999–1002. Bibcode:2002Sci...297..999A. doi:10.1126/science.1070656. PMID   12169731.
  37. 1 2 Ambrosio Flores, L (2000). "Land cover estimation in small areas using ground survey and remote sensing". Remote Sensing of the Environment. 74 (2): 240–248. Bibcode:2000RSEnv..74..240F. doi:10.1016/S0034-4257(00)00114-0.
  38. Green, Russell G. Congalton, Kass (25 January 2019). Assessing the Accuracy of Remotely Sensed Data: Principles and Practices, Third Edition (3 ed.). Boca Raton: CRC Press. doi:10.1201/9780429052729. ISBN   978-0-429-05272-9.{{cite book}}: CS1 maint: multiple names: authors list (link)
  39. Stehman, S. (2013). "Estimating Area from an Accuracy Assessment Error Matrix". Remote Sensing of Environment. 132 (132): 202–211. Bibcode:2013RSEnv.132..202S. doi:10.1016/j.rse.2013.01.016.
  40. Stehman, S. (2019). "Key issues in rigorous accuracy assessment of land cover products". Remote Sensing of Environment. 231 (231). Bibcode:2019RSEnv.23111199S. doi:10.1016/j.rse.2019.05.018.
  41. Zhen, Z (2013). "Impact of training and validation sample selection on classification accuracy and accuracy assessment when using reference polygons in object-based classification". International Journal of Remote Sensing. 34 (19): 6914–6930. Bibcode:2013IJRS...34.6914Z. doi:10.1080/01431161.2013.810822.
  42. Czaplewski, R.L. "Misclassification bias in areal estimates". Photogrammetric Engineering and Remote Sensing (39): 189–192.
  43. Bauer, M.E. (1978). "Area estimation of crops by digital analysis of Landsat data". Photogrammetric Engineering and Remote Sensing (44): 1033–1043.
  44. Olofsson, P. (2014). "Good practices for estimating area and assessing accuracy of land change". Remote Sensing of Environment. 148 (148): 42–57. Bibcode:2014RSEnv.148...42O. doi:10.1016/j.rse.2014.02.015.
  45. Mcroberts, R (2018). "The effects of imperfect reference data on remote sensing-assisted estimators of land cover class proportions". ISPRS Journal of Photogrammetry and Remote Sensing. 142 (142): 292–300. Bibcode:2018JPRS..142..292M. doi: 10.1016/j.isprsjprs.2018.06.002 .
  46. Foody, G.M. (2010). "Assessing the accuracy of land cover change with imperfect ground reference data". Remote Sensing of Environment. 114 (10): 2271–2285. Bibcode:2010RSEnv.114.2271F. doi:10.1016/j.rse.2010.05.003.
  47. Sannier, C (2014). "Using the regression estimator with landsat data to estimate proportion forest cover and net proportion deforestation in gabon". Remote Sensing of Environment. 151 (151): 138–148. Bibcode:2014RSEnv.151..138S. doi:10.1016/j.rse.2013.09.015.
  48. Gallego, F.J. (2004). "Remote sensing and land cover area estimation". International Journal of Remote Sensing. 25 (5): 3019–3047. Bibcode:2004IJRS...25.3019G. doi:10.1080/01431160310001619607.
  49. Carfagna, E. (2005). "Using remote sensing for agricultural statistics". International Statistical Review. 73 (3): 389–404. doi:10.1111/j.1751-5823.2005.tb00155.x.
  50. Maksel, Rebecca. "Flight of the Giant". Air & Space Magazine. Archived from the original on 18 August 2021. Retrieved 19 February 2019.
  51. IWM, Alan Wakefield
    Head of photographs at (4 April 2014). "A bird's-eye view of the battlefield: aerial photography". The Daily Telegraph. ISSN   0307-1235. Archived from the original on 18 April 2014. Retrieved 19 February 2019.
  52. 1 2 Bagley, James (1941). Aerophotography and Aerosurverying (1st ed.). York, PA: The Maple Press Company.
  53. "Air Force Magazine". www.airforcemag.com. Archived from the original on 19 February 2019. Retrieved 19 February 2019.
  54. "Military Imaging and Surveillance Technology (MIST)". www.darpa.mil. Archived from the original on 18 August 2021. Retrieved 19 February 2019.
  55. The Indian Society of International Law - Newsletter: VOL. 15, No. 4, October - December 2016 (Report). doi:10.1163/2210-7975_hrd-9920-2016004.
  56. "In Depth | Magellan". Solar System Exploration: NASA Science. Archived from the original on 19 October 2021. Retrieved 19 February 2019.
  57. Garner, Rob (15 April 2015). "SOHO - Solar and Heliospheric Observatory". NASA. Archived from the original on 18 September 2021. Retrieved 19 February 2019.
  58. Campbell, James B.; Wynne, Randolph H. (21 June 2011). Introduction to Remote Sensing (5th ed.). New York London: The Guilford Press. ISBN   978-1-60918-176-5.
  59. Ryerson, Robert A. (2010). Why 'where' matters : understanding and profiting from GPS, GIS, and remote sensing : practical advice for individuals, communities, companies and countries. Internet Archive. Manotick, ON : Kim Geomatics Corp. ISBN   978-0-9866376-0-5.
  60. Fussell, Jay; Rundquist, Donald; Harrington, John A. (September 1986). "On defining remote sensing" (PDF). Photogrammetric Engineering and Remote Sensing. 52 (9): 1507–1511. Archived from the original (PDF) on 4 October 2021.
  61. Pruitt, Evelyn L. (1979). "The Office of Naval Research and Geography". Annals of the Association of American Geographers. 69 (1): 103–108. doi:10.1111/j.1467-8306.1979.tb01235.x. ISSN   0004-5608. JSTOR   2569553.
  62. Colen, Jerry (8 April 2015). "Ames Research Center Overview". NASA. Archived from the original on 28 September 2021. Retrieved 19 February 2019.
  63. Ditter, R., Haspel, M., Jahn, M., Kollar, I., Siegmund, A., Viehrig, K., Volz, D., Siegmund, A. (2012) Geospatial technologies in school – theoretical concept and practical implementation in K-12 schools. In: International Journal of Data Mining, Modelling and Management (IJDMMM): FutureGIS: Riding the Wave of a Growing Geospatial Technology Literate Society; Vol. X
  64. Stork, E.J., Sakamoto, S.O., and Cowan, R.M. (1999) "The integration of science explorations through the use of earth images in middle school curriculum", Proc. IEEE Trans. Geosci. Remote Sensing 37, 1801–1817
  65. Bednarz, S.W.; Whisenant, S.E. (July 2000). "Mission Geography: Linking national geography standards, innovative technologies and NASA". IGARSS 2000. IEEE 2000 International Geoscience and Remote Sensing Symposium. Taking the Pulse of the Planet: The Role of Remote Sensing in Managing the Environment. Proceedings (Cat. No.00CH37120). Vol. 6. pp. 2780–2782 vol.6. doi:10.1109/IGARSS.2000.859713. ISBN   0-7803-6359-0. S2CID   62414447.
  66. Digital Earth
  67. "FIS – Remote Sensing in School Lessons". Archived from the original on 26 October 2012. Retrieved 25 October 2012.
  68. "geospektiv". Archived from the original on 2 May 2018. Retrieved 1 June 2018.
  69. "YCHANGE". Archived from the original on 17 August 2018. Retrieved 1 June 2018.
  70. "Landmap – Spatial Discovery". Archived from the original on 29 November 2014. Retrieved 27 October 2021.
  71. Grasty, R (1976). Applications of Gamma Radiation in Remote Sensing (1st ed.). Berlin: Springer-Verlag. p. 267. ISBN   978-3-642-66238-6.
  72. 1 2 3 Tatem, Andrew J.; Goetz, Scott J.; Hay, Simon I. (2008). "Fifty Years of Earth-observation Satellites". American Scientist. 96 (5): 390–398. doi:10.1511/2008.74.390. PMC   2690060 . PMID   19498953.
  73. Kuznetsov, V.D.; Sinelnikov, V.M.; Alpert, S.N. (June 2015). "Yakov Alpert: Sputnik-1 and the first satellite ionospheric experiment". Advances in Space Research. 55 (12): 2833–2839. Bibcode:2015AdSpR..55.2833K. doi:10.1016/j.asr.2015.02.033.
  74. "James A. Van Allen". nmspacemuseum.org. New Mexico Museum of Space History. Retrieved 14 May 2018.
  75. "How many Earth observation satellites are orbiting the planet in 2021?". 18 August 2021.
  76. "DubaiSat-2, Earth Observation Satellite of UAE". Mohammed Bin Rashid Space Centre. Archived from the original on 17 January 2019. Retrieved 4 July 2016.
  77. "DubaiSat-1, Earth Observation Satellite of UAE". Mohammed Bin Rashid Space Centre. Archived from the original on 4 March 2016. Retrieved 4 July 2016.

Further reading