Flight simulator

Last updated

F/A-18 Hornet flight simulator aboard the USS Independence aircraft carrier 980310-N-7355H-003 Simulator Training.jpg
F/A-18 Hornet flight simulator aboard the USS Independence aircraft carrier

A flight simulator is a device that artificially re-creates aircraft flight and the environment in which it flies, for pilot training, design, or other purposes. It includes replicating the equations that govern how aircraft fly, how they react to applications of flight controls, the effects of other aircraft systems, and how the aircraft reacts to external factors such as air density, turbulence, wind shear, cloud, precipitation, etc. Flight simulation is used for a variety of reasons, including flight training (mainly of pilots), the design and development of the aircraft itself, and research into aircraft characteristics and control handling qualities. [1]

Contents

The term "flight simulator" may carry slightly different meaning in general language and technical documents. In past regulations, it referred specifically to devices which can closely mimic the behavior of aircraft throughout various procedures and flight conditions. [2] In more recent definitions, this has been named "full flight simulator". [3] The more generic term "flight simulation training device" (FSTD) is used to refer to different kinds of flight training devices, and that corresponds more closely to meaning of the phrase "flight simulator" in general English. [4]

History of flight simulation

In 1910, on the initiative of the French commanders Clolus and Laffont and Lieutenant Clavenad, the first ground training aircraft for military aircraft were built. The "Tonneau Antoinette" (Antoinette barrel), created by the Antoinette company, seems to be the precursor of flight simulators.

World War I (1914–1918)

An area of training was for air gunnery handled by the pilot or a specialist air gunner. Firing at a moving target requires aiming ahead of the target (which involves the so-called lead angle) to allow for the time the bullets require to reach the vicinity of the target. This is sometimes also called "deflection shooting" and requires skill and practice. During World War I, some ground-based simulators were developed to teach this skill to new pilots. [5]

The 1920s and 1930s

Link Trainer patent drawing, 1930 Edlink pt1930.jpg
Link Trainer patent drawing, 1930

The best-known early flight simulation device was the Link Trainer, produced by Edwin Link in Binghamton, New York, United States, which he started building in 1927. He later patented his design, which was first available for sale in 1929. The Link Trainer was a basic metal frame flight simulator usually painted in its well-known blue color. Some of these early war era flight simulators still exist, but it is becoming increasingly difficult to find working examples. [6]

The Link family firm in Binghamton manufactured player pianos and organs, and Ed Link was therefore familiar with such components as leather bellows and reed switches. He was also a pilot, but dissatisfied with the amount of real flight training that was available, he decided to build a ground-based device to provide such training without the restrictions of weather and the availability of aircraft and flight instructors. His design had a pneumatic motion platform driven by inflatable bellows which provided pitch and roll cues. A vacuum motor similar to those used in player pianos rotated the platform, providing yaw cues. A generic replica cockpit with working instruments was mounted on the motion platform. When the cockpit was covered, pilots could practice flying by instruments in a safe environment. The motion platform gave the pilot cues as to real angular motion in pitch (nose up and down), roll (wing up or down) and yaw (nose left and right). [7]

Initially, aviation flight schools showed little interest in the "Link Trainer". Link also demonstrated his trainer to the U.S. Army Air Force (USAAF), but with no result. However, the situation changed in 1934 when the Army Air Force was given a government contract to fly the postal mail. This included having to fly in bad weather as well as good, for which the USAAF had not previously carried out much training. During the first weeks of the mail service, nearly a dozen Army pilots were killed. The Army Air Force hierarchy remembered Ed Link and his trainer. Link flew in to meet them at Newark Field in New Jersey, and they were impressed by his ability to arrive on a day with poor visibility, due to practice on his training device. The result was that the USAAF purchased six Link Trainers, and this can be said to mark the start of the world flight simulation industry. [7]

World War II (1939–1945)

Military Personnel Using Link Trainer, Pepperell Manufacturing Co., 1943 Military Personnel Using Link Trainer, Pepperell Manufacturing Company (11327128056).jpg
Military Personnel Using Link Trainer, Pepperell Manufacturing Co., 1943

The principal pilot trainer used during World War II was the Link Trainer. Some 10,000 were produced to train 500,000 new pilots from allied nations, many in the US and Canada because many pilots were trained in those countries before returning to Europe or the Pacific to fly combat missions. [7] Almost all US Army Air Force pilots were trained in a Link Trainer. [8]

A different type of World War II trainer was used for navigating at night by the stars. The Celestial Navigation Trainer of 1941 was 13.7 m (45 ft) high and capable of accommodating the navigation team of a bomber crew. It enabled sextants to be used for taking "star shots" from a projected display of the night sky. [7]

1945 to the 1960s

In 1954 United Airlines bought four flight simulators at a cost of $3 million from Curtiss-Wright that were similar to the earlier models, with the addition of visuals, sound and movement. This was the first of today's modern flight simulators for commercial aircraft. [9]

A simulator for helicopters existed as the Jacobs Jaycopter as means of “Cutting helicopter training cost.”. [10] [11] [12] The simulator was later sold as a funfair ride in the 1964-65 New York World's Fair. [13]

Today

Cockpit of a twinjet flight simulator AC97-0295-13 a.jpeg
Cockpit of a twinjet flight simulator

The simulator manufacturers are consolidating and integrate vertically as training offers double-digit growth: CAE forecast 255,000 new airline pilots from 2017 to 2027 (70 a day), and 180,000 first officers evolving to captains. The largest manufacturer is Canadian CAE Inc. with a 70% market share and $2.8 billion annual revenues, manufacturing training devices for 70 years but moved into training in 2000 with multiple acquisitions. Now CAE makes more from training than from producing the simulators. Crawley-based L3 CTS entered the market in 2012 by acquiring Thales Training & Simulation's manufacturing plant near Gatwick Airport where it assembles up to 30 devices a year, then UK CTC training school in 2015, Aerosim in Sanford, Florida in 2016, and Portuguese academy G Air in October 2017. [14]

With a 20% market share, equipment still accounts for more than half of L3 CTS turnover but that could soon be reversed as it educates 1,600 commercial pilots each year, 7% of the 22,000 entering the profession annually, and aims for 10% in a fragmented market. The third largest is TRU Simulation + Training, created in 2014 when parent Textron Aviation merged its simulators with Mechtronix, OPINICUS and ProFlight, focusing on simulators and developing the first full-flight simulators for the 737 MAX and the 777X. The fourth is FlightSafety International, focused on general, business and regional aircraft. Airbus and Boeing have invested in their own training centres, aiming for higher margins than aircraft manufacturing like MRO, competing with their suppliers CAE and L3. [14]

In June 2018, there were 1,270 commercial airline simulators in service, up by 50 over a year: 85% FFSs and 15% FTDs. CAE supplied 56% of this installed base, L3 CTS 20% and FlightSafety International 10%, while CAE's training centres are the largest operator, with a 13% share. North America has 38% of the world's training devices, Asia-Pacific 25% and Europe 24%. Boeing types represent 45% of all simulated aircraft, followed by Airbus with 35%, then Embraer at 7%, Bombardier at 6% and ATR at 3%. [15]

Applications

Pilot training

Interior of a flight simulator in Estonia, for a Piper Seneca PA-34
(view as a 360deg interactive panorama) Eesti Lennuakadeemia 360-8.jpg
Interior of a flight simulator in Estonia, for a Piper Seneca PA-34
( view as a 360° interactive panorama )

Most flight simulators are used primarily for flight training. The simplest simulators are used to practice basic cockpit procedures, such as processing emergency checklists, and for cockpit familiarization. They are also used for instrument flight training, [16] [17] for which the outside view is less important. Certain aircraft systems may or may not be simulated, and the aerodynamic model is usually extremely generic if present at all. [18] Depending on the level of certification, instruments that would have moving indicators in a real aircraft may be implemented with a display. With more advanced displays, cockpit representation and motion systems, flight simulators can be used to credit different amount of flight hours towards a pilot license. [19]

Specific classes of simulators are also used for training other than obtaining initial license such as instrument rating revalidation, or most commonly [20] obtaining type rating for specific kind of aircraft.

Other uses

During the aircraft design process, flight simulators can be used instead of performing some flight tests. Such "engineering flight simulators" can provide a fast way to find errors, reducing both the risks and the cost of development. [21] Additionally, this allows use of extra measurement equipment that might be too large or otherwise impractical to include during onboard a real aircraft. Throughout different phases of the design process, different engineering simulators with various level of complexity are used. [22] :13

Flight simulators may include training tasks for crew other than pilots. Examples include gunners on a military aircraft [23] or hoist operators. [24] Separate simulators have also been used for tasks related to flight, like evacuating the aircraft in case of a crash in water. [25] With high complexity of many systems composing contemporary aircraft, aircraft maintenance simulators are increasingly popular. [26] [27]

Qualification and approval

Full flight simulator of a Boeing 737 Baltic Aviation Academy Airbus B737 Full Flight Simulator (FFS).jpg
Full flight simulator of a Boeing 737
A spherical display with multiple projectors visible above the cockpit SIMAER.jpg
A spherical display with multiple projectors visible above the cockpit

Procedure

Before September 2018, [28] when a manufacturer wished to have an ATD model approved, a document that contains the specifications for the model line and that proves compliance with the appropriate regulations is submitted to the FAA. Once this document, called a Qualification Approval Guide (QAG), has been approved, all future devices conforming to the QAG are automatically approved and individual evaluation is neither required nor available. [29]

The actual procedure accepted by all CAAs (Civil Aviation Authorities) around the world is to propose 30 days prior qualification date (40 days for CAAC) a MQTG document (Master Qualification Test Guide), which is proper to a unique simulator device and will live along the device itself, containing objective, and functional and subjective tests to demonstrate the representativeness of the simulator compare to the airplane. The results will be compared to Flight Test Data provided by aircraft OEMs or from test campaign ordered by simulator OEMs or also can be compared by POM (Proof Of Match) data provided by aircraft OEMs development simulators. Some of the QTGs will be rerun during the year to prove during continuous qualification that the simulator is still in the tolerances approved by the CAA. [30] [16] [31]

US Federal Aviation Administration (FAA) categories

Aviation Training Device (ATD) [32]
Flight Training Devices (FTD) [33]
Full Flight Simulators (FFS) [34]

European Aviation Safety Agency (EASA, ex JAA) categories

These definitions apply to both airplanes [3] and helicopters [35] unless specified otherwise. Training devices briefly compared below are all different subclasses of Flight simulation training device (FSTD).

Basic instrument training device (BITD)airplanes only : A basic student station for instrument flight procedures; can use spring loaded flight controls, and instruments displayed on a screen

Flight Navigation and Procedures Trainer (FNPT) : Representation of cockpit with all equipment and software to replicate function of aircraft systems

Flight Training Devices (FTD)

Full Flight Simulators (FFS)

Technology

Simulator structure

Flight simulator block diagram Flight simulator block diagram.svg
Flight simulator block diagram

Flight simulators are an example of a human-in-the-loop system, in which interaction with a human user is constantly happening. From perspective of the device, the inputs are primary flight controls, instrument panel buttons and switches and the instructor's station, if present. Based on these, the internal state is updated, and equations of motion solved for the new time step. [37] The new state of the simulated aircraft is shown to the user through visual, auditory, motion and touch channels.

To simulate cooperative tasks, the simulator can be suited for multiple users, as is the case with multi-crew cooperation simulators. Alternatively, more simulators can be connected, what is known as "parallel simulation" or "distributed simulation". [38] As military aircraft often need to cooperate with other craft or military personnel, wargames are a common use for distributed simulation. Because of that, numerous standards for distributed simulation including aircraft have been developed with military organisations. Some examples include SIMNET, DIS and HLA .

Simulation models

The central element of simulation model are the equations of motion for the aircraft. [37] As the aircraft moves through atmosphere it can exhibit both translational and rotational degrees of freedom. To achieve perception of fluent movement, these equations are solved 50 or 60 times per second. [22] :16 The forces for motion are calculated from aerodynamical models, which in turn depend on state of control surfaces, driven by specific systems, with their avionics, etc. As is the case with modelling, depending on the required level of realism, there are different levels of detail, with some sub-models omitted in simpler simulators.

If a human user is part of the simulator, which might not be the case for some engineering simulators, there is a need to perform the simulation in real-time. Low refresh rates not only reduce realism of simulation, but they have also been linked with increase in simulator sickness. [39] The regulations place a limit on maximum latency between pilot input and aircraft reaction. Because of that, tradeoffs are made to reach the required level of realism with a lower computational cost. Flight simulators typically don't include full computational fluid dynamics models for forces or weather, but use databases of prepared results from calculations and data acquired in real flights. As an example, instead of simulating flow over the wings, lift coefficient may be defined in terms of motion parameters like angle of attack. [22] :17

While different models need to exchange data, most often they can be separated into a modular architecture, for better organisation and ease of development. [40] [41] Typically, gear model for ground handling would be separate input to the main equations of motion. Each engine and avionics instrument is also a self-contained system with well-defined inputs and outputs.

Instruments

Simulator with primary flight instruments replicated with flat displays Simpic.jpg
Simulator with primary flight instruments replicated with flat displays

All classes of FSTD require some form of replicating the cockpit. As they are the primary means of interaction between the pilot and the aircraft special importance is assigned to cockpit controls. To achieve good transfer of skills, there are very specific requirements in the flight simulator regulations [16] that determine how closely they must match the real aircraft. These requirements in case of full flight simulators are so detailed, that it may be cost-effective to use the real part certified to fly, rather than manufacture a dedicated replica. [22] :18 Lower classes of simulators may use springs to mimic forces felt when moving the controls. When there is a need to better replicate the control forces or dynamic response, many simulators are equipped with actively driven force feedback systems. Vibration actuators may also be included, either due to helicopter simulation requirements, or for aircraft equipped with a stick shaker.

Another form of tactile input from the pilot are instruments located on the panels in the cockpit. As they are used to interact with various aircraft systems, just that may be sufficient for some forms of procedure training. Displaying them on a screen is sufficient for the most basic BITD simulators [3] and amateur flight simulation, however most classes of certified simulators need all buttons, switches and other inputs to be operated in the same way as in the aircraft cockpit. The necessity for a physical copy of a cockpit contributes to the cost of simulator construction, and ties the hardware to a specific aircraft type. Because of these reasons, there is ongoing research on interactions in virtual reality, however lack of tactile feedback negatively affects users' performance when using this technology. [42] [43]

Visual system

A wide angle cylindrical display DA42-Simulator at Horizon SFA.jpg
A wide angle cylindrical display

Outside view from the aircraft is an important cue for flying the aircraft, and is the primary means of navigation for visual flight rules operation. [44] One of the primary characteristics of a visual system is the field of view. Depending on the simulator type it may be sufficient to provide only a view forward using a flat display. However, some types of craft, e.g. fighter aircraft, require a very large field of view, preferably almost full sphere, due to the manoeuvres that are performed during air combat. [45] Similarly, since helicopters can perform hover flight in any direction, some classes of helicopter flight simulators require even 180 degrees of horizontal field of view. [46]

There are many parameters in visual system design. For a narrow field of view, a single display may be sufficient, however typically multiple projectors are required. This arrangement needs additional calibration, both in terms of distortion from not projecting on a flat surface, as well as brightness in regions with overlapping projections. [47] There are also different shapes of screens used, including cylindrical, [48] spherical [47] or ellipsoidal. The image can be projected on the viewing side of the projection screen, or alternatively "back-projection" onto a translucent screen. [49] Because the screen is much closer than objects outside aircraft, for simulators with multiple pilots, there are special collimated displays that eliminate the parallax effect between the pilots' point of view. [50]

An alternative to large-scale displays are virtual reality simulators using a head-mounted display. This approach allows for a complete field of view, and makes the simulator size considerably smaller. There are examples of use in research, [41] as well as certified FSTD. [51]

Contribution to modern Computer Graphics. Visual simulation science applied from the visual systems developed in flight simulators were also an important precursor to three dimensional computer graphics and Computer Generated Imagery (CGI) systems today. Namely because the object of flight simulation is to reproduce on the ground the behavior of an aircraft in flight. Much of this reproduction had to do with believable visual synthesis that mimicked reality. [52] Combined with the need to pair virtual synthesis with military level training requirements, graphics technologies applied in flight simulation were often years ahead of what would have been available in commercial products. When CGI was first used to train pilots, early systems proved effective for certain simple training missions but needed further development for sophisticated training tasks as terrain following and other tactical maneuvers. Early CGI systems could depict only objects consisting of planar polygons. Advances in algorithms and electronics in flight simulator visual systems and CGI in the 1970s and 1980s influenced many technologies still used in modern graphics. Over time CGI systems were able to superimpose texture over the surfaces and transition from one level of image detail to the next one in a smooth manner. [53] Real-time computer graphics visualization of virtual worlds makes some aspects of flight simulator visual systems very similar to game engines, sharing some techniques like different levels of details or libraries like OpenGL. [22] :343 Many computer graphics visionaries began their careers at Evans & Sutherland and Link Flight Simulation, Division of Singer Company, two leading companies in flight simulation before today's modern computing era. For example, the Singer Link Digital Image Generator (DIG) created in 1978 was considered one of the worlds first CGI system. [54]

Motion system

Stewart platform Hexapod general Anim.gif
Stewart platform

Initially, the motion systems used separate axes of movement, similar to a gimbal. After the invention of Stewart platform [55] simultaneous operation of all actuators became the preferred choice, with some FFS regulations specifically requiring "synergistic" 6 degrees of freedom motion. [56] In contrast to real aircraft, the simulated motion system has a limited range in which it is able to move. That especially affects the ability to simulate sustained accelerations, and requires a separate model to approximate the cues to the human vestibular system within the given constraints. [22] :451

Motion system is a major contributor to overall simulator cost [22] :423, but assessments of skill transfer based on training on a simulator and leading to handling an actual aircraft are difficult to make, particularly where motion cues are concerned. Large samples of pilot opinion are required and many subjective opinions tend to be aired, particularly by pilots not used to making objective assessments and responding to a structured test schedule. For many years, it was believed that 6 DOF motion-based simulation gave the pilot closer fidelity to flight control operations and aircraft responses to control inputs and external forces and gave a better training outcome for students than non-motion-based simulation. This is described as "handling fidelity", which can be assessed by test flight standards such as the numerical Cooper-Harper rating scale for handling qualities. Recent scientific studies have shown that the use of technology such as vibration or dynamic seats within flight simulators can be equally effective in the delivery of training as large and expensive 6-DOF FFS devices. [57] [58]

Modern high-end flight simulators

Vertical Motion Simulator (VMS) at NASA/Ames

The largest flight simulator in the world is the Vertical Motion Simulator (VMS) at NASA Ames Research Center, south of San Francisco. This has a very large-throw motion system with 60 feet (+/- 30 ft) of vertical movement (heave). The heave system supports a horizontal beam on which are mounted 40 ft rails, allowing lateral movement of a simulator cab of +/- 20 feet. A conventional 6-degree of freedom hexapod platform is mounted on the 40 ft beam, and an interchangeable cabin is mounted on the platform. This design permits quick switching of different aircraft cabins. Simulations have ranged from blimps, commercial and military aircraft to the Space Shuttle. In the case of the Space Shuttle, the large Vertical Motion Simulator was used to investigate a longitudinal pilot-induced oscillation (PIO) that occurred on an early Shuttle flight just before landing. After identification of the problem on the VMS, it was used to try different longitudinal control algorithms and recommend the best for use in the Shuttle program. [59]

Disorientation training

AMST Systemtechnik GmbH (AMST) of Austria and Environmental Tectonics Corporation (ETC) of Philadelphia, US, manufacture a range of simulators for disorientation training, that have full freedom in yaw. The most complex of these devices is the Desdemona simulator at the TNO Research Institute in The Netherlands, manufactured by AMST. This large simulator has a gimballed cockpit mounted on a framework which adds vertical motion. The framework is mounted on rails attached to a rotating platform. The rails allow the simulator cab to be positioned at different radii from the centre of rotation and this gives a sustained G capability up to about 3.5. [60] [61]

Amateur and video game flight simulation

See also

Related Research Articles

<span class="mw-page-title-main">Simulation</span> Imitation of the operation of a real-world process or system over time

A simulation is an imitative representation of a process or system that could exist in the real world. In this broad sense, simulation can often be used interchangeably with model. Sometimes a clear distinction between the two terms is made, in which simulations require the use of models; the model represents the key characteristics or behaviors of the selected system or process, whereas the simulation represents the evolution of the model over time. Another way to distinguish between the terms is to define simulation as experimentation with the help of a model. This definition includes time-independent simulations. Often, computers are used to execute the simulation.

<span class="mw-page-title-main">Stick shaker</span> Mechanical device in an aircraft cockpit to warn the pilot of an imminent stall

A stick shaker is a mechanical device designed to rapidly and noisily vibrate the control yoke of an aircraft, warning the flight crew that an imminent aerodynamic stall has been detected. It is typically present on the majority of large civil jet aircraft, as well as most large military planes.

X-Plane is a flight simulation software initially launched by Laminar Research in 1995. Commercial desktop versions are sold for macOS, Windows, and Linux. In addition, Laminar Research also distributes FAA-certified versions for professional use. A mobile version has been available for Android, iOS, and webOS since 2009 as well.

<span class="mw-page-title-main">Simulation cockpit</span> Cockpit used for training pilots with a flight simulator

A simulation cockpit, simpit or sim rig is an environment designed to replicate a vehicle cockpit. Although many pits commonly designed around an aircraft cockpit, the term is equally valid for train, spacecraft or car projects.

<span class="mw-page-title-main">Garmin G1000</span> Electronic flight instrument system

The Garmin G1000 is an electronic flight instrument system (EFIS) typically composed of two display units, one serving as a primary flight display, and one as a multi-function display. Manufactured by Garmin Aviation, it serves as a replacement for most conventional flight instruments and avionics. Introduced in June 2004, the system has since become one of the most popular integrated glass cockpit solutions for general aviation and business aircraft.

Combat flight simulators are vehicle simulation games, amateur flight simulation computer programs used to simulate military aircraft and their operations. These are distinct from dedicated flight simulators used for professional pilot and military flight training which consist of realistic physical recreations of the actual aircraft cockpit, often with a full-motion platform.

<span class="mw-page-title-main">Motion simulator</span> Type of mechanism

A motion simulator or motion platform is a mechanism that creates the feelings of being in a real motion environment. In a simulator, the movement is synchronised with a visual display of the outside world (OTW) scene. Motion platforms can provide movement in all of the six degrees of freedom (DOF) that can be experienced by an object that is free to move, such as an aircraft or spacecraft:. These are the three rotational degrees of freedom and three translational or linear degrees of freedom.

Thales Training & Simulation Ltd. is a multinational company which manufactures simulators, including full flight simulators and military simulators, and provides related training and support services. It is a wholly owned subsidiary of the Thales Group.

<span class="mw-page-title-main">John D. Odegard School of Aerospace Sciences</span>

The John D. Odegard School of Aerospace Sciences is a multidisciplinary college within the University of North Dakota (UND) in Grand Forks, North Dakota. The school was formed in 1968. The majority of the school's fleet of over 120 aircraft is based at nearby Grand Forks International Airport and is the largest fleet of civilian flight training aircraft in North America. UND Aerospace also operates a flight training center at Phoenix–Mesa Gateway Airport in Mesa, Arizona. Today, the school has many aerospace-related programs including commercial aviation, Unmanned aircraft systems operations, air traffic control, airport management, Space Studies, Computer Science, Atmospheric Sciences, and Earth System Science & Policy. Currently, the school has over 500 faculty and 2,000 students making it the second largest of UND's degree-granting colleges. The present dean of the school is Robert Kraus.

<span class="mw-page-title-main">Type rating</span> Certification of an airplane pilot to fly a certain type of aircraft

A type rating is an authorization entered on or associated with a pilot license and forming part thereof, stating the pilot's privileges or limitations pertaining to certain aircraft type. Such qualification requires additional training beyond the scope of the initial license and aircraft class training.

Frasca International, Inc., is an American manufacturer of flight simulation training devices, with over 3000 training devices delivered in approximately 70 countries throughout the world. Based in Urbana, Illinois, Frasca International was founded in Champaign, Illinois in 1958 by Rudy Frasca. It is owned by FlightSafety International, a division of Berkshire Hathaway, following an acquisition in 2022.
Frasca flight simulators are used in all segments of the aviation industry, including their extensive use in many prominent college aviation programs, including Purdue University, Indiana State University, Embry-Riddle Aeronautical University, University of North Dakota, Liberty University, Louisiana Tech University, University of Illinois, and Western Michigan University.

<span class="mw-page-title-main">Full flight simulator</span>

Full flight simulator (FFS) is a term used by national (civil) aviation authorities (NAA) for a high technical level of flight simulator. Such authorities include the Federal Aviation Administration (FAA) in the United States and the European Aviation Safety Agency (EASA).

<span class="mw-page-title-main">Acceleration onset cueing</span>

Acceleration onset cueing is a term for the cueing principle used by a simulator motion platform.

<i>Digital Combat Simulator</i> Combat flight simulation game developed by Eagle Dynamics

Digital Combat Simulator, or DCS, is a combat flight simulation game developed primarily by Eagle Dynamics and The Fighter Collection.

A Control Loading System, is used to provide pilots with realistic flight control forces in a flight simulator or training device. These are used in both commercial and military training applications.

<span class="mw-page-title-main">Flight simulation video game</span> Video game genre

A flight simulation video game refers to the simulation of various aspects of flight or the flight environment for purposes other than flight training or aircraft development. A significant community of simulation enthusiasts is supported by several commercial software packages, as well as commercial and homebuilt hardware. Open-source software that is used by the aerospace industry like FlightGear, whose flight dynamics engine (JSBSim) is used in a 2015 NASA benchmark to judge new simulation code to space industry standards, is also available for private use. A popular type of flight simulators video games are combat flight simulators, which simulate combat air operations from the pilot and crew's point of view. Combat flight simulation titles are more numerous than civilian flight simulators due to variety of subject matter available and market demand.

ATP Flight School is the largest flight training company in the United States. The curriculum focuses on airline-oriented pilot programs at locations across the country. ATP is the leading supplier of professionally trained pilots to the nation's regional airlines.

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

A pilot logbook is a record of a pilot's flying hours. It contains every flight a pilot has flown, including flight time, number of landings, and types of instrument approaches made. Pilots also log simulator time, as it counts towards training.

Unmanned aircraft system simulation focuses on training pilots to control an unmanned aircraft or its payload from a control station. Flight simulation involves a device that artificially re-creates aircraft flight and the environment in which it flies for pilot training, design, or other purposes. It includes replicating the equations that govern how aircraft fly, how they react to applications of flight controls, the effects of other aircraft systems, and how the aircraft reacts to external factors such as air density, turbulence, wind shear, cloud, precipitation, etc.

<span class="mw-page-title-main">European Helicopter Safety Team</span>

The European Helicopter Safety Team (EHEST) was a European aviation safety improvement initiative focusing on improving helicopter safety in Europe and worldwide. It was established in 2006 as part of the European Strategic Safety Initiative (ESSI) of the European Aviation Safety Agency (EASA). The goal of the European Helicopter Safety Team was to contribute to reducing the worldwide helicopter accident rate by 80% in the time-span 2006-2016, which was set as a goal by the International Helicopter Safety Team (IHST) in 2006. Focusing on European helicopter operators and manufacturers, the European Helicopter Safety Team conducted helicopter accident analyses, provided technology potential studies, and published safety management and training documents.

References

Notes

  1. Federal Aviation Administration (25 April 2013). "FAR 121 Subpart N—Training Program" . Retrieved 28 April 2013.
  2. "AC 120-40 Airplane Simulator and Visual System Evaluation" (PDF). Federal Aviation Administration.
  3. 1 2 3 CS FSTD(A).200: Terminology
  4. "Definition of flight simulator from the Cambridge Advanced Learner's Dictionary & Thesaurus". Cambridge University Press.
  5. Bonnier Corporation (January 1919). "Dry Shooting for Airplane Gunners". Popular Science Monthly. Bonnier Corporation. pp. 13–14.
  6. Fly Away Simulation (12 July 2010). "Flight Simulator Technology Through the Years". Archived from the original on 12 October 2011. Retrieved 20 April 2011.
  7. 1 2 3 4 "ASME Landmarks: The Link Flight Trainer." Archived 17 December 2011 at the Wayback Machine American Society of Mechanical Engineers. Retrieved: 18 December 2011.
  8. "U.S. Air Force Fact Sheet: Link Trainer." National Museum of the United States Air Force. Retrieved: 12 October 2016.
  9. Hearst Magazines (September 1954). "Airline Pilots Fly Anywhere in the world – Without Leaving the Ground". Popular Mechanics. Hearst Magazines. p. 87.
  10. Canadian Aviation February 1961
  11. Fortier, Rénald. "A helicopter simulator with a difference: it flies – Canada's Jacobs Jaycopter". Ingenium Canadian National Museums of Science and Innovation.
  12. "Jaycopter, Edmonton, Alberta". Provincial Archives of Alberta - A Question.
  13. "The 1964-1965 New York World's Fair - Jaycopter Ride". WorldsFairPhotos.com.
  14. 1 2 Murdo Morrison (25 June 2018). "Civil simulator manufacturer strategies compared". FlightGlobal.
  15. Antoine Fafard (26 June 2018). "Analysis: Civil simulator fleet nears 1,300 mark". FlightGlobal.
  16. 1 2 3 "EASA CS-FSTD(A) Issue 2" (PDF).
  17. Leonard Ross; Paul Slotten; Louise Yeazel (1990). "Pilot's Evaluation of the Usefulness of Full Mission IFR Simulator Flights for General Aviation Pilot Training". Journal of Aviation/Aerospace Education & Research. 1 (2). doi:10.15394/JAAER.1990.1024. ISSN   1065-1136. Wikidata   Q112800809.
  18. "Navy CPT". www.navair.navy.mil. U.S. Navy. Archived from the original on 8 August 2014. Retrieved 4 August 2014.
  19. "14 CFR Appendix D to Part 141 4.(c)". Archived from the original on 24 April 2022. Retrieved 1 July 2022.
  20. European Helicopter Safety Team (EHEST). "Advantages of Simulators (FSTDs) in Helicopter Flight Training" (PDF). European Union Aviation Safety Agency (EASA). p. 6. Retrieved 29 June 2022.
  21. David J Allerton (December 2010). "The impact of flight simulation in aerospace". The Aeronautical Journal. 114: 6. doi:10.1017/S0001924000004231. ISSN   0001-9240. Wikidata   Q112813532.
  22. 1 2 3 4 5 6 7 David J Allerton (2009). Principles of flight simulation. Wiley. doi:10.2514/4.867033. ISBN   978-0-470-75436-8. Wikidata   Q112813340.
  23. Susan T. Heers; Patricia A. Casper (October 1998). "Subjective Measurement Assessment in a Full Mission Scout-Attack Helicopter Simulation". Proceedings of the Human Factors and Ergonomics Society Annual Meeting . 42 (1): 26–30. doi:10.1177/154193129804200107. ISSN   1071-1813. Wikidata   Q112800993.
  24. Michael King; Stephen Lenser; D Rogers; H Carnahan (2 January 2022). "Novice and experienced hoist operators in a helicopter hoist virtual reality simulator". International Journal of Training Research. 20: 1–13. ISSN   1448-0220. Wikidata   Q112805528.
  25. Karsten Hytten (November 1989). "Helicopter crash in water: Effects of simulator Escape training". Acta Psychiatrica Scandinavica . 80: 73–78. doi:10.1111/J.1600-0447.1989.TB05256.X. ISSN   0001-690X. Wikidata   Q112805503.
  26. André Pinheiro; Paulo Fernandes; Ana Maia; et al. (2012). "Development of a Mechanical Maintenance Training Simulator in OpenSimulator for F-16 Aircraft Engines". Procedia Computer Science. 15: 248–255. doi:10.1016/J.PROCS.2012.10.076. ISSN   1877-0509. Wikidata   Q57592005.
  27. Francesca De Crescenzio; Massimiliano Fantini; Franco Persiani; Luigi Di Stefano; Pietro Azzari; Samuele Salti (1 January 2011). "Augmented reality for aircraft maintenance training and operations support". IEEE Computer Graphics and Applications . 31 (1): 96–101. doi:10.1109/MCG.2011.4. ISSN   0272-1716. PMID   24807975. Wikidata   Q87833678.
  28. "FAA AC 61-136B" (PDF).
  29. "FAA AC 61-136A" (PDF).
  30. "FAA CFR Part 60" (PDF).
  31. "CAAC CCAR-60" (PDF).
  32. AC-61-136A Appendix 1 and 2
  33. 14 CFR Part 60, Appendices B and D
  34. 14 CFR Part 60, Appendices A and C
  35. CS FSTD(H).200: Terminology
  36. Appendix 1 to CS FSTD(H).300, Appendix 1 to CS FSTD(A).300
  37. 1 2 Baarspul, M. (1990) A review of flight simulation techniques. Progress in Aerospace Sciences, 22, 1–20.
  38. Richard Fujimoto (December 2015), Parallel and distributed simulation, doi:10.1109/WSC.2015.7408152, Wikidata   Q63321790
  39. Randy Pausch; Thomas Crea; Matthew Conway (January 1992). "A Literature Survey for Virtual Environments: Military Flight Simulator Visual Systems and Simulator Sickness". Presence: Teleoperators & Virtual Environments . 1 (3): 344–363. doi:10.1162/PRES.1992.1.3.344. ISSN   1054-7460. Wikidata   Q112822678.
  40. C. A. Ippolito; Amy Ruth Pritchett (14 August 2000), Software architecture for a Reconfigurable Flight Simulator, American Institute of Aeronautics and Astronautics, doi:10.2514/6.2000-4501, Wikidata   Q112822781
  41. 1 2 Matthias Oberhauser; Daniel Dreyer (1 September 2017). "A virtual reality flight simulator for human factors engineering". Cognition, Technology and Work. 19 (2): 263–277. doi:10.1007/S10111-017-0421-7. ISSN   1435-5558. Wikidata   Q112822831.
  42. Turgay Aslandere; Daniel Dreyer; Frieder Pankratz (March 2015). Virtual hand-button interaction in a generic virtual reality flight simulator. pp. 1–8. doi:10.1109/AERO.2015.7118876. ISBN   978-1-4799-5379-0. Wikidata   Q112826446.{{cite book}}: |journal= ignored (help)
  43. Markus Tatzgern; Cristoph Birgmann (March 2021). "Exploring Input Approximations for Control Panels in Virtual Reality". Virtual Reality and 3D User Interfaces: 1–9. doi:10.1109/VR50410.2021.00092. Wikidata   Q112826551.
  44. Section 91.155 14 CFR Part 91 - General Operating and Flight Rules - FAA
  45. R. Barette; A. Morris; J. Baribeau (22 July 1985), A modern air combat dome visual system, American Institute of Aeronautics and Astronautics, doi:10.2514/6.1985-1747, Wikidata   Q112840484
  46. Appendix 1 to CS FSTD(H).300, 1.3 Visual system, requirement b.3
  47. 1 2 Brian Reno (14 August 1989), Full field of view dome display system, Boston: American Institute of Aeronautics and Astronautics, doi:10.2514/6.1989-3316, Wikidata   Q112790735
  48. Brent Cameron; Hooman Rajaee; Bradley Jung; Robert Langlois (May 2016), Development and Implementation of Cost-effective Flight Simulator Technologies, doi:10.11159/CDSR16.126, Wikidata   Q112812641
  49. Leonard G. Best; Don R. Wight; Philip W. Peppler (16 August 1999), M2DART: a real image rear-projection display, pp. 348–355, doi:10.1117/12.357610, Wikidata   Q112840621
  50. Byron J. Pierce; George A. Geri (October 1998). "The Implications of Image Collimation for Flight Simulator Training". Proceedings of the Human Factors and Ergonomics Society Annual Meeting . 42 (20): 1383–1387. doi:10.1177/154193129804202004. ISSN   1071-1813. Wikidata   Q112793062.
  51. "EASA approves the first virtual reality (VR) based flight simulation training device". European Union Aviation Safety Agency (EASA). 26 April 2021. Retrieved 30 June 2022.
  52. Rolfe, JM and Staples, KJ (27 May 1988). Flight Simulation Cambridge Aerospace Series No 1. Cambridge University Press. ISBN   978-0521357517.{{cite book}}: CS1 maint: multiple names: authors list (link)
  53. Yan, Johnson (August 1985). "Advances in Computer-Generated Imagery for Flight Simulation". IEEE. 5 (8): 37–51. doi:10.1109/MCG.1985.276213. S2CID   15309937.
  54. Carlson, Wayne (20 June 2017). "Computer Graphics and Animation: a retrospective review". p. 13.2.
  55. Stewart, D. (1965–1966). "A Platform with Six Degrees of Freedom". Proceedings of the Institution of Mechanical Engineers. 180 (1, No 15): 371–386. doi:10.1243/pime_proc_1965_180_029_02.
  56. Appendix 1 to CS FSTD(H).300, 1.2 Motion system, requirement b.1
  57. Andrea L. Sparko; Judith Bürki-Cohen; Tiauw H. Go (2010). Transfer of Training from a Full-Flight Simulator vs. a High Level Flight Training Device with a Dynamic Seat. AIAA Modeling and Simulation Technologies Conference. doi:10.2514/6.2010-8218.
  58. Peter John Davison. "A summary of studies conducted on the effect of motion in flight simulator pilot training" (PDF). MPL Simulator Solutions. Retrieved 12 November 2019.
  59. Beard, Steven; et al. "Space Shuttle Landing and Rollout Training at the Vertical Motion Simulator" (PDF). AIAA. Archived from the original (PDF) on 20 January 2009. Retrieved 5 February 2014.
  60. "DESDEMONA: The next generation in movement simulation" Nederlandse Organisatie voor Toegepast Natuurwetenschappelijk Onderzoek Retrieved: 5 July 2012.
  61. Roza, M., M. Wentink and Ph. Feenstra. "Performance Testing of the Desdemona Motion System." AIAA MST, Hilton Head, South Carolina, 20–23 August 2007.

Bibliography

  • Kelly, Lloyd L. as told to Robert B. Parke. The Pilot Maker. New York: Grosset & Dunlap, 1979, First edition 1970. ISBN   0-448-02226-5.