Study Copy Paste

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 132

The expected holding speed for many charted holding patterns is published on the associated enroute,
terminal or approach chart. In cases where a speed is not specified, holding patterns must be entered and
flown at or below the appropriate airspeed for the holding altitude. These speeds can vary from region to
region so pilots must be aware of the limitations in force for the area in which they are
operating. International Civil Aviation Organisation (ICAO) maximum holding speeds are as follows:
 Holding altitude 14000' or below - 230 KIAS
 Holding altitude above 14000' to 20000' - 240 KIAS
 Holding altitude above 20000' to 34000' - 265 KIAS
 Holding altitude above 34000' - Mach .83
 Pilots are to advise ATC immediately if airspeeds in excess of those specified above become necessary
for any reason, including turbulence. After such higher speed is no longer necessary, the aircraft should
be operated at or below the specified airspeeds and ATC notified
 Airspace protection for turbulent air holding is based on a maximum of 280 KIAS or Mach 0.8, whichever
is lower, from the Minimum Holding Altitude (MHA) to 34000' and Mach .83 above that altitude.
 Considerable impact on the flow of air traffic may result when aircraft hold at speeds which are higher than
those specified above. After departing a holding fix, pilots should resume normal speed subject to other
requirements, such as speed limitations in the vicinity of controlled airports, specific ATC requests, etc.
 Minimum holding speeds are defined by the airframe manufacturers (i.e. not by ICAO). The
values depend on the aircraft type and mass. Operators sometimes extend these definitions to include
specific weather conditions such as icing. Information about the minimum speeds can be found in relevant
aircraft manuals.

Holding Pattern

Inbound timing for a standard hold is one minute when at or below 14,000' and one and a half minutes when
above 14,000'. When the pilot receives ATC clearance.
Communication Failure: Guidance for Controllers
The aircraft shall comply with the voice communication failure procedures of Annex 10, Volume II, and
with those of the following procedures as are appropriate. The aircraft shall attempt to establish
communications with the appropriate air traffic control unit using all other available means. In addition, the
aircraft, when forming part of the aerodrome traffic at a controlled aerodrome, shall keep a watch for such
instructions as may be issued by visual signals.
An aircraft equipped with an SSR transponder is expected to operate the transponder on Mode A Code
7600 to indicate that it has experienced air-ground communication failure. An ADS-B equipped aircraft
experiencing radio communication failure may transmit the appropriate ADS-B emergency and/or urgency
mode. An aircraft equipped with other surveillance system transmitters, including ADS-C, might indicate
the loss of air-ground communication by all of the available means.
If the aircraft fails to indicate that it is able to receive and acknowledge transmissions, separation shall be
maintained between the aircraft having the communication failure and other aircraft, based on the
assumption that the aircraft will:
 In VMC:
1. Continue to fly in visual meteorological conditions;
2. Land at the nearest suitable aerodrome; and
3. Report the arrival by the most expeditious means to the appropriate air traffic control unit
 In IMC or when conditions are such that it does not appear likely that the pilot will complete the flight in
accordance with the prescribed VMC RCF procedures above:
1. Unless otherwise prescribed on the basis of a regional air navigation agreement, in airspace where
procedural separation is being applied, maintain the last assigned speed and level, or minimum flight
altitude if higher, for a period of 20 minutes following the aircraft’s failure to report its position over a
compulsory reporting point and thereafter adjust level and speed in accordance with the filed flight plan; or
2. In airspace where an ATS surveillance system is used in the provision of air traffic control, maintain the
last assigned speed and level, or minimum flight altitude if higher, for a period of 7 minutes following:
i) The time the last assigned level or minimum flight altitude is reached; or
ii) Time the transponder is set to Code 7600 or the ADS-B transmitter is set to indicate the loss of air-
ground communications; or
iii) The aircraft’s failure to report its position over a compulsory reporting point;
whichever is later and thereafter adjust level and speed in accordance with the filed flight plan;
3. When being vectored or having been directed by ATC to proceed offset using RNAV without a specified
limit, proceed in the most direct manner possible to rejoin the current flight plan route no later than the
next significant point, taking into consideration the applicable minimum flight altitude;
4. Proceed according to the current flight plan route to the appropriate designated navigation aid or fix
serving the destination aerodrome and, when required to ensure compliance with 5), hold over this aid or
fix until commencement of descent;
5. Commence descent from the navigation aid or fix specified in 4) at, or as close as possible to, the
expected approach time last received and acknowledged; or, if no expected approach time has been
received and acknowledged, at, or as close as possible to, the estimated time of arrival resulting from the
current flight plan;
6. Complete a normal instrument approach procedure as specified for the designated navigation aid or fix;
and
7. Land, if possible, within 30 minutes after the estimated time of arrival specified in 5) or the last
acknowledged expected approach time, whichever is later.
Runway Edge Light Spacing and Color

Both High Intensity Runway Lights (HIRLs) and Medium Intensity Runway Lights (MIRLs)
require a maximum spacing of 200 feet between each runway edge light. For runways with
intersecting taxiways or other runways, the maximum gap cannot exceed 400 feet.

Runway edge lights are white, until you start getting close to the departure end of the runway.
On instrument runways, edge lights are yellow on the last 2,000', or half the runway length,
whichever is less

Runway centerline lights are spaced at 50 feet apart.


Relation between IAS<CAS<TAS<MACH
Defining Each Runway Declared Distance

To define and calculate each declared distance, we must first look at what is meant by a clearway,
stopway, and displaced threshold. Each definition has been taken from ICAO’s Annex 14:

 Clearway (CWY) – A defined rectangular area on the ground or water under the control of the appropriate
authority selected or prepared as a suitable area over which an aeroplane may make a portion of its initial
climb to a specified height.
 Stopway (SWY) – A defined rectangular area on the ground at the end of take-off run available prepared as
a suitable area in which an aircraft can be stopped in the case of an abandoned take-off.
 Displaced threshold – A threshold not located at the extremity of a runway.

The definitions of each declared distance, also defined by ICAO, are presented below:

 TORA – The length of runway declared available and suitable for the ground run of an aircraft taking off.
 TODA – The length of TORA plus length of the CWY, if provided.
 ASDA – The length of TORA plus the length of the SWY, if provided.
 LDA – The length of the runway declared available and suitable for the ground run of an aircraft landing.
(The length of the TORA minus the length of the displaced threshold).
Vertical minima:
 CAT I Because the aircraft is unlikely to be flying over level ground at the same elevation as the touch-
down zone when passing the Missed Approach Point, the vertical minima used in a CAT I approach is
measured by reference to a barometric altimeter. In practice, this means that when flying a CAT I
approach either a Decision Altitude (DA) or Decision Height (DH) may be used.
 CAT II/III Because greater precision is required when flying a CAT II or CAT III approach, special attention
is given to the terrain in the runway undershoot area to enable a radio altimeter to be used. CAT II and
CAT III approaches are therefore always flown to a DH with reference to a radio altimeter.
CAT II and CAT III instrument approach and landing operations are not permitted unless RVR information
is provided.
On reaching the DH, the pilot may continue the approach to land provided that the required visual
references have been established. Otherwise the pilot must commence a missed approach procedure.
The DA or DH is a specified height in the Precision Approach or approach with vertical guidance at which
a Missed Approach must be initiated if the required visual reference to continue the approach has not
been established. - (ICAO Annex 6)
COMMUNICATION FAILURE
The aircraft shall comply with the voice communication failure procedures of Annex 10, Volume II, and
with those of the following procedures as are appropriate. The aircraft shall attempt to establish
communications with the appropriate air traffic control unit using all other available means. In addition, the
aircraft, when forming part of the aerodrome traffic at a controlled aerodrome, shall keep a watch for such
instructions as may be issued by visual signals.
An aircraft equipped with an SSR transponder is expected to operate the transponder on Mode A Code
7600 to indicate that it has experienced air-ground communication failure. An ADS-B equipped aircraft
experiencing radio communication failure may transmit the appropriate ADS-B emergency and/or urgency
mode. An aircraft equipped with other surveillance system transmitters, including ADS-C, might indicate
the loss of air-ground communication by all of the available means.
If the aircraft fails to indicate that it is able to receive and acknowledge transmissions, separation shall be
maintained between the aircraft having the communication failure and other aircraft, based on the
assumption that the aircraft will:
 In VMC:
1. Continue to fly in visual meteorological conditions;
2. Land at the nearest suitable aerodrome; and
3. Report the arrival by the most expeditious means to the appropriate air traffic control unit
 In IMC or when conditions are such that it does not appear likely that the pilot will complete the flight in
accordance with the prescribed VMC RCF procedures above:
1. Unless otherwise prescribed on the basis of a regional air navigation agreement, in airspace where
procedural separation is being applied, maintain the last assigned speed and level, or minimum flight
altitude if higher, for a period of 20 minutes following the aircraft’s failure to report its position over a
compulsory reporting point and thereafter adjust level and speed in accordance with the filed flight plan; or
2. In airspace where an ATS surveillance system is used in the provision of air traffic control, maintain the
last assigned speed and level, or minimum flight altitude if higher, for a period of 7 minutes following:
i) The time the last assigned level or minimum flight altitude is reached; or
ii) Time the transponder is set to Code 7600 or the ADS-B transmitter is set to indicate the loss of air-
ground communications; or
iii) The aircraft’s failure to report its position over a compulsory reporting point;
whichever is later and thereafter adjust level and speed in accordance with the filed flight plan;
3. When being vectored or having been directed by ATC to proceed offset using RNAV without a specified
limit, proceed in the most direct manner possible to rejoin the current flight plan route no later than the
next significant point, taking into consideration the applicable minimum flight altitude;
4. Proceed according to the current flight plan route to the appropriate designated navigation aid or fix
serving the destination aerodrome and, when required to ensure compliance with 5), hold over this aid or
fix until commencement of descent;
5. Commence descent from the navigation aid or fix specified in 4) at, or as close as possible to, the
expected approach time last received and acknowledged; or, if no expected approach time has been
received and acknowledged, at, or as close as possible to, the estimated time of arrival resulting from the
current flight plan;
6. Complete a normal instrument approach procedure as specified for the designated navigation aid or fix;
and
7. Land, if possible, within 30 minutes after the estimated time of arrival specified in 5) or the last
acknowledged expected approach time, whichever is later.

A note concerning departing aircraft experiencing RCF: If the aircraft has been vectored away from the
route specified in the flight plan then the flight crew is expected to comply with the procedures published
in the appropriate regional air navigation agreement and included in the SID description or published in
the AIP.
Reduced Vertical Separation Minima (RVSM)
A program was initiated by ICAO in 1982 involving worldwide studies to assess the feasibility of a
reduction of the Vertical Separation Minima (VSM) above FL290 from 2,000 feet to 1,000 feet.
The principal benefits which the implementation of the reduced VSM were expected to provide were:

 A theoretical doubling of the airspace capacity, between FL290 and FL410; and
 The opportunity for aircraft to operate at closer to the optimum flight levels with the resulting fuel
economies.
The program relies on the carriage and serviceability of specified aircraft equipment and the existence of
appropriate operating procedures to ensure that the risk of loss of separation is no greater than it would
be outside RVSM airspace

 An operator shall ensure that aeroplanes operated in RVSM airspace are equipped with:

1. Two independent altitude measurement systems;


2. An altitude alerting system;
3. An automatic altitude control system; and
4. A secondary surveillance radar (SSR) transponder with altitude reporting system that can be
connected to the altitude measurement system in use for altitude keeping. (IR-OPS
SPA.RVSM.110, EU-OPS 1.872)

Separation standards within RVSM Airspace

Within RVSMairspace (between FL290 and FL410 inclusive) the vertical separation minimum is:
 1000ft (300m) between RVSM-approved aircraft, and
 2000ft (600m) between non-RVSM approved state aircraft and any other aircraft operating within RVSM
airspace.
 2000ft (600m) between non-RVSM aircraft operating as general air traffic (GAT) and any other aircraft
within RVSM airspace.

Contingency procedures when unable to maintain RVSM

 The pilots shall notify ATC of any equipment failure, weather hazards such as severe turbulence etc.,
which may affect the ability to maintain the cleared level or the RVSM requirements. When an aircraft
operating in RVSM Airspace encounters severe turbulence due to weather or wake vortex which the pilot
believes will impact the aircraft’s capability to maintain its cleared flight level, the pilot shall inform ATC.
ATC is required to establish either an appropriate horizontal separation minimum, or an increased vertical
separation minimum of 2000ft;
 Where a meteorological forecast is predicting severe turbulence within the RVSM Airspace, ATC shall
determine whether RVSM should be suspended, and, if so, the period of time, and specific flight level(s)
and/or area.
 When notified by ATC of an assigned altitude deviation of more than 300ft (90 m), the pilot shall take
action to return to the cleared level as quickly as possible.
 In the event of a pilot advising that the aircraft is no longer capable of RVSM operations, it is particularly
important that the first ATS unit made aware of the failure performs the necessary co-ordination with
subsequent ATS units.
Circling Approach
A circling approach is an extension of an instrument approach procedure which provides for visual circling
of the aerodrome prior to landing. (ICAO Doc 8168: Procedures for Air Navigation Services - Aircraft
Operations (PANS-OPS) Vol I - Flight Procedures)
A circling approach is the visual phase of an instrument approach to bring an aircraft into position for
landing on a runway which is not suitably located for a straight-in approach. (JAR-OPS 1.435 (a) (1))
Go-around from a Circling Approach

Because the runway on which the aircraft makes the instrument approach is not the runway to which it is
circling, confusion may exist in a pilot's mind if a go-around should become necessary. This would create
a dangerous situation if, for example, the pilot flew the missed approach for the landing runway instead of
the instrument runway. Therefore a standard procedure has been established by ICAO to address this
issue:
If visual reference is lost while circling to land from an instrument approach, the missed
approach specified for that particular procedure shall be followed. The transition from the visual (circling)
manoeuvre to the missed approach should be initiated by a climbing turn, within the circling area, towards
the landing runway, to return to the circling altitude or higher, immediately followed by interception and
execution of the missed approach procedure. The indicated airspeed during these manoeuvres shall not
exceed the maximum indicated airspeed associated with visual manoeuvring. (ICAO Doc 8168: PANS-
OPS, Volume 1, Chapter 7, Section 7.4)

What is Barrette

A simple barrette approach lighting system.

Three or more aeronautical ground lights closely spaced in a transverse line, so that from a distance they appear as a sho
rt bar of light (ICAO).
MSA (Minimum Sector Altitude)
MSA is the minimum safe altitude within a 25 nm (Nautical Mile) radius of a navigational aid. MSA will give you
1000 ft separation from the highest terrain within that sector. We normally see the MSA on our Approach
plates for runways.
Can we descend below MSA? Yes, provided we are under radar control and we still remain above the radar
altitudes. (Remember it is still your responsibility to make sure you don’t fly into terrain!)

MEA (Minimum En-Rout Altitude)

MEA is the minimum altitude that will guarantee signal of navigation aids along the route, give you two way
communication with ATC and provide obstacle clearance. MEA will give you 1000 ft separation from terrain
in non mountainous areas and 2000 ft separation in mountainous areas. (Mountainous area is when a terrain
change of more than 3000 ft is experienced within 10 nm)

MOCA (Minimum Obstacle Clearance Altitude)

MOCA will give you a minimum altitude above terrain and guarantee VOR reception within 22 nm from the
beacon. MOCA will give you 1000 ft separation from terrain when terrain is less than 5001 ft and 2000 ft
separation when terrain is more than 5000 ft.

MORA (Minimum Off-Route Altitude)

MORA will give you separation from terrain up to 10 nm off the route center line and 10 nm radius around the
ends of the route. MORA will give you 1000 ft separation from terrain in non mountainous areas and 2000 ft
separation in mountainous areas.

GRID MORA (Grid Minimum Off-Route Altitude)

GRID MORA gives you terrain separation within latitude and longitude lines. GRID MORA will give you 1000 ft
separation from terrain when terrain is less than 5001 ft and 2000 ft separation when terrain is more than
5000 ft.

MVA (Minimum Vector Altitudes)

MVA gives you terrain separation when Vectored by ATC, remember you have to be under radar control to be
able to go down to these altitudes. (often lower than the MSA) MVA will give you 1000 ft separation from
terrain in non mountainous areas and 2000 ft separation in mountainous areas.

I hope these brief explanations help for interviews and make you a more knowledgeable pilot. Please comment
if you have a different understanding of the above or if you would like to see any other topics.
Right-of-way
Description

Rights of way have been established and agreed internationally to ensure that aircraft in proximity
with each other in VMC know which aircraft has right of way over the other and what action must
be taken to avoid collision.
Right-of-way rules are described in ICAO Annex 2: Rules of the Air, as follows:
3.2.1 Right-of-way. The aircraft that has the right-of-way shall maintain its heading and speed.
3.2.2.1 An aircraft that is obliged by the following rules to keep out of the way of another shall
avoid passing over, under or in front of the other, unless it passes well clear and takes into
account the effect of aircraft wake turbulence.
3.2.2.2 Approaching head-on. When two aircraft are approaching head-on or approximately so
and there is danger of collision, each shall alter its heading to the right.
3.2.2.3 Converging. When two aircraft are converging at approximately the same level, the
aircraft that has the other on its right shall give way, except as follows:

a) power-driven heavier-than-air aircraft shall give way to airships, gliders and balloons;
b) airships shall give way to gliders and balloons;
c) gliders shall give way to balloons;
d) power-driven aircraft shall give way to aircraft which are seen to be towing other aircraft or
objects.
3.2.2.4 Overtaking. An overtaking aircraft is an aircraft that approaches another from the rear on
a line forming an angle of less than 70 degrees with the plane of symmetry of the latter, i.e. is in
such a position with reference to the other aircraft that at night it should be unable to see either of
the aircraft’s left (port) or right (starboard) navigation lights. An aircraft that is being overtaken has
the right-of-way and the overtaking aircraft, whether climbing, descending or in horizontal flight,
shall keep out of the way of the other aircraft by altering its heading to the right, and no
subsequent change in the relative positions of the two aircraft shall absolve the overtaking aircraft
from this obligation until it is entirely past and clear.

Separate rules apply to aircraft Landing (3.2.2.5), Taking off (3.2.2.6), and Taxiing (3.2.2.7).
In the EU, the right of way is covered by Commission Regulation (EU) No 923/2012 laying down
the common rules of the air and operational provisions regarding services and procedures in air
navigation (SERA). The specific texts can be found in Annex, Section 3, Chapter 2, SERA.3210.
INSTRUMENT APPROCAH SEGMENTS
Instrument approach procedure (IAP). A series of predetermined manoeuvres by reference to flight
instruments with specified protection from obstacles from the initial approach fix, or where applicable, from
the beginning of a defined arrival route to a point from which a landing can be completed and thereafter, if
a landing is not completed, to a position at which holding or en-route obstacle clearance criteria apply.
Instrument approach procedures are classified as follows:
 Non-precision approach (NPA) procedure
 Approach procedure with vertical guidance (APV)
 Precision approach (PA) procedure
Source: ICAO Doc 8168 PANS-OPS
Description

An instrument approach procedure may have five separate segments:


 Arrival segment - this segment is a transition from the en-route phase to the approach phase of the flight.
 Initial approach segment - this segment begins at the initial approach fix (IAF) and ends at the
intermediate fix.
 Intermediate approach segment - this segment usually begins at the intermediate fix (IF) and ends at
the final approach fix (FAF) or final approach point (FAP). Here the aircraft speed and configuration
should be adjusted to prepare for the final approach. For this reason, the descent gradient is kept as
shallow as possible.
 Final approach segment - this segment usually begins at the FAF/FAP and ends at the missed approach
point (MAPt). Here the alignment and descent for landing are made. Final approach may be made to a
runway for a straight-in landing, or to an aerodrome for a visual manoeuvre. Some aerodromes are not
equipped with a facility to serve as FAF. If this is the case, descent to MDA/H is made once the aircraft is
established inbound on the final approach track.

Approach Segments
An instrument approach may be divided into as many as four approach segments: initial, intermediate, final,
and missed approach.

Initial
The initial approach segment begins at an initial approach fix (IAF) and usually ends where it joins the
intermediate approach segment.

Purpose: To provide a method for aligning your aircraft with the approach course by using an arc procedure, a
course reversal, or by following a route that intersects the final approach course.

Intermediate
This segment begins at the intermediate fix (IF) which is usually aligned within 30° of the final approach
course. If no fix is shown for this segment then it begins at a point where you are proceeding inbound to the
final approach fix and are properly aligned with the final approach course.
Purpose: This is designed primarily to position your aircraft for the final descent to the airport.

Final
For a nonprecision approach, the final approach segment begins either at a designated final approach fix (FAF)
or at a point where you are established on the final approach course. When an FAF is not designate (on-airport
VOR or NDB) this point is typically where the procedure turn intersects the final approach course inbound and
is referred to as the final approach point (FAP).

For a precision approach the final approach segment begins where the glide slope is intercepted at the
minimum glide slope intercept altitude.

Purpose: Allows you to navigate safely to a point at which, if the required visual references are available, you
can continue the approach to a landing. If you cannot see the required cues at the missed approach point, you
must execute the missed approach procedure.

Missed Approach
The missed approach segment begins at the missed approach point (MAP) and ends at a designated point,
such as an initial approach or enroute fix. The MAP depends on the type of approach you are flying. If it’s a
precision approach then the MAP occurs when you reach a designated altitude on the glide slope called the
decision height (DH). For a nonprecision approach it’s when you hit either a fix defined by a navaid or after a
specified period of time has elapsed since crossing the FAF.

Purpose: To allow you to safely navigate from the missed approach point to a point where you can attempt
another approach or continue to another airport.

CAVOK
CAVOK stands for "Ceiling and Visibility OK." It is an abbreviation used in some non-U.S. aviation
surface weather observation reports (METARs). Information on visibility, runway visual range, present
weather and cloud amount, cloud type, and height of cloud base may be replaced in all
meteorological reports by the term "CAVOK" when the following conditions apply:

1. Visibility of 10 km or more, and the lowest visibility is not reported


(Note 1: In local routine and special reports, visibility refers to the value(s) to be reported in
accordance with International Civil Aviation Organization (ICAO) Annexes 3 4.2.4.2 and
4.2.4.3; in METARs and SPECIs, visibility refers to the value(s) to be reported in accordance
with ICAO Annex 3 4.2.4.4.)
(Note 2: The lowest visibility is reported in accordance with ICAO Annex 3 4.2.4.4.);
2. No cloud of operational significance; and
3. No weather of significance to aviation as given in ICAO Annexes 3 4.4.2.3, 4.4.2.5, and
4.4.2.6;

Note that this term is not approved, nor will be used, in the United States.

What Does 9999 Mean in a METAR?


In countries that report visibility in meters, 9999 means that the vis is 9,000 meters or
better.

NOSIG METAR
What is NOSIG? No significant weather. It’s yet another way of denoting no significant
restrictions to ceilings or visibility.

VV/// Meaning in METAR


The code group in a standard METAR report that begins with VV indicates “Vertical
Visibility.” This is the term used when the sky is obscured, in other words, a ground
observer cannot make out a clearly defined ceiling.

If the code appears as VV///, without number, it means that the observer cannot make out
how many feet the visibility is equal to. In most cases, the visibility will be reported in
hundreds of feet. So, for example, VV005 would be 500 feet of vertical visibility.

“Better than 5,000 and 5”


In the US, controllers often abbreviate weather reports over the radio. This reduces
frequency congestion and makes the information a little easier for the pilots to understand.
One common method is to say simply, “The weather is better than five-thousand and five.”
In other words, ceilings are higher than 5,000 feet, and visibility is more than 5 SM—so the
weather is good VFR in most cases.

TAKEOFF SEGMENTS
The Net Takeoff Flight Path for the engine failure case is divided into four segments. Three of these
are climbing segments with specified minimum gradients which are dependent upon the number of
engines installed on the aircraft and one is a level acceleration segment. A brief description of the
four segments is as follows:
1. First Segment - depending upon the regulations under which the aircraft is certified, the first segment
begins either at lift-off or at the end of the takeoff distance at a screen height of 35' and a speed of V2.
On a wet runway, the screen height is reduced to 15'. Operating engines are at takeoff thrust, the
flaps/slats are in takeoff configuration and landing gear retraction is initiated once safely airborne with
positive climb. The first segment ends when the landing gear is fully retracted.
2. Second Segment - begins when the landing gear is fully retracted. Engines are at takeoff thrust and
the flaps/slats are in the takeoff configuration. This segment ends at the higher of 400' or specified
acceleration altitude. In most cases, the second segment is the performance limiting segment of the
climb.
3. Third or Acceleration Segment- begins at the higher of 400' or specified acceleration altitude.
Engines are at takeoff thrust and the aircraft is accelerated in level flight. Slats/flaps are retracted on
speed. The segment ends when aircraft is in clean configuration and a speed of VFS has been
achieved. Note that the third segment must be completed prior to exceeding the maximum time
allowed for engines at takeoff thrust.
4. Fourth or Final Segment - begins when the aircraft is in clean configuration and at a speed of V FS.
Climb is re-established and thrust is reduced to maximum continuous (MCT). The segment ends at a
minimum of 1500' above airport elevation or when the criteria for enroute obstacle clearance have
been met.
Each segment of the one engine inoperative takeoff flight path has a mandated climb gradient
requirement. For example, a gross second segment climb gradient capability of 2.4%, 2.7% or 3.0%
is required for two, three and four engine aircraft respectively. Similarly, the required gross gradients
for the fourth segment are 1.2%, 1.5% and 1.7% respectively.
To ensure obstacle clearance while allowing for aircraft performance degradation and less than
optimum pilot technique, the gross gradients are reduced by 0.8%, 0.9% and 1.0% respectively to
calculate a net gradient. The obstacle identification surface (OIS), or obstruction envelope, starts at
runway elevation at a point directly beneath the end of the takeoff distance (TOD) and parallels the
net gradient profile of the climb segments. If an obstacle in the departure path penetrates the OIS, the
slope of the OIS must be increased and both the net and the gross gradient slopes of the
corresponding segment must also be increased to ensure that the minimum obstacle clearance
criteria is met.
The aircraft net gradient capability, correctable for temperature, altitude and pressure, is published in
the AFM performance data and, in actual operations, must ensure that the limiting obstacle in the
departure path can be cleared by a minimum of 35'. If there is an obstacle within the departure path
that cannot be avoided and would not be cleared by 35', the planned takeoff weight must be reduced
until minimum obstacle clearance can be achieved. Note that, by regulation, turns immediately after
takeoff cannot be initiated below the greater of 50'AGL or one half of the aircraft wingspan and, that
during the initial climb, turns are limited to 15° of bank. Turning will result in a reduction in aircraft
climb capability.
Take off Flight Path
To maximise the payload capability from any given runway, most operators develop and
utilize emergency turn procedures. These procedures follow a specified ground track which minimises
the affects of local obstacles and a specified vertical profile which complies with the more restrictive
of certification or actual obstacle climb requirements.
SCREEN HEIGHT: The take off part of the flight is the distance from where the brakes are released
to the point at which the aircraft reaches a defined height.This defined height is known as
screen height.It is usually 35 ft (for class A aircraft) on a dry runway and if the runway is wet it can
reduce down to 15 ft.

Atmosphere
The atmosphere is the space around the Earth which is filled by a mixture of gasses held against the
Earth by the force of gravity. This mixture of gasses we call air.
Because the Earth spins on its axis, and because the surface temperature is greater at the equator than at
the poles, the atmosphere extends further out into space at the equator than at the poles. Air Density
decreases with increasing altitude.
The atmosphere is divided vertically into four regions: the Troposphere, the Stratosphere,
the Mesosphere, and the Thermosphere.

Troposphere and Stratosphere

The boundary between the troposphere and the stratosphere is known as the Tropopause and is covered
by a separate article.
Most light aircraft and turboprop aircraft fly within the troposphere and this is where most of the water
vapour and therefore cloud formation exists. Many types of jet aircraft are able to cruise in the
Stratosphere, especially in high latitudes where the Tropopause is lower, thereby avoiding almost
all weather - although some particular vigorous thunderstorms may penetrate the stratosphere.
 Temperature falls with height in the troposphere but is generally constant at about -57°C in the
stratosphere.
 Considerable vertical movement of air occurs in the troposphere - warm air rising and cool air descending.
This vertical movement is a direct consequence of Solar Heating.

The Structure of the Atmosphere


There are five layers in the atmosphere starting from the surface of Earth upwards and they are
named as follows:

 Troposphere
 Stratosphere
 Mesosphere
 ionosphere
 Exosphere
The troposphere extends from the surface up to average height of 11 km. within this layer,
temperature decreases 2 degrees Celsius per 1,000 feet and pressure drops by 1 inch of mercury per
1,000 feet. This layer is of an extreme importance to everyone involved in the aviation world due to it
holding all the water vapor in the atmosphere and being the layer where most flying occurs.

The Troposphere
In the troposphere, air masses are constantly moving around from high pressure to low pressure areas
which causes major weather phenomena from different types of clouds and storms to drastic changes
in temperature which makes it an essential part of the meteorological studies a student pilot has to
complete in his/her respective flight academy.

The Tropopause is the upper boundary of the troposphere which separates the troposphere from the
stratosphere. At the tropopause, the temperature ceases to decrease with decreasing height and it is
also the height which the ozone layer exists. The height of the tropopause varies between 26,000 feet
and 52,000 feet depending on the location on the planet and the time of the year.

The stratosphere
The stratosphere extends from the tropopause to approximately 50 km above the surface of the Earth.
Some flying occurs in the lower parts of the stratosphere, so the combination of the troposphere and
lower parts of the stratosphere is therefore often referred to as the aviation atmosphere.

This layer is relatively stable. Initially, the temperature remains constant and then starts to increase so
that is around 0 degrees Celsius at the top of the layer. This is due to the absorption of ultra violet
radiation by ozone in the lower layers of the stratosphere and the retransmission of this radiation as
infra-red heat. The concentration of ozone varies with latitude, being greater over the poles than the
equator. This layer has very little vertical movement concerning cloud formation but very strong
horizontal winds than can very beneficial to pilots.

The remaining layers carry very little to no weather, and are not usable at all by any commercial
aviation aircrafts except for the traveling of GPS transmission and that will be covered at a later stage.

Please refer to figure 01 below which serves as a summary to the structure of atmosphere.

Types of Clouds and How to Recognize Them

The World Meteorological Organization lists over 100 different types of clouds in its
International Cloud Atlas. But, there are three or four main types of genera of clouds,
which combine and form 10 basic cloud types. Here is an overview of the types of clouds,
how to recognize them, and what kind of weather they produce.

 Cumulus clouds look like fluffy cotton balls. Low to the ground, they indicate fair
weather, but when they tower into the sky they produce storms.
 Stratus clouds are sheets or layers of clouds. They can produce an overcast day or light
rain or snow.
 Cirrus clouds are white wisps high in the sky.
 Nimbus clouds are either cumulus or stratus clouds that produce precipitation. Rain
and snow clouds tend to be gray rather than white.

How Are Clouds Classified?


Clouds get classified according to the altitude at which they form, their upper height, and
their shape. The names come from Latin words. Four common types are cumulus, stratus,
nimbus, and cirrus. Alternatively, clouds get names from their height: high (cirrus or ciro-),
middle (alto-), low (stratus, strato-), and multilevel (cumulus, cumulo-, nimbo-). In terms
of shape, stratiform describes flattened sheets, cirriform describes wisps or patches, and
cumuliform clouds are masses or heaps. Nimbus clouds are ones that produce
precipitations.

Types of Clouds
Cumulus Clouds
Cumulus clouds are the puffy popcorn shapes you produce when asked to draw clouds.
They are puffy with rounded tops. Cumulus clouds appear white when well-lit, but have
flattened, darker bases.
You’ll see cumulus clouds form on clear, sunny days as the ground heats up. They often
appear mid-morning, develop, and vanish as the sun sets. They indicate the presence of
moisture in the air, but don’t necessarily produce rain or snow.
Stratus Clouds
Stratus clouds are flat, uniform, gray clouds. Usually, they are gray and resemble fog near
the horizon, rather than actual fog which is on the ground.
Stratus clouds produce overcast days. Sometimes they produce mist or a light drizzle.
Altostratus (and altocumulus) clouds sometimes produce virga, which is precipitation that
evaporates before it reaches the ground.
Nimbus Clouds
Nimbus clouds produce precipitation in the form of rain, hail, or snow. Both cumulus and
stratus clouds can be nimbus clouds, which go by the names cumulonimbus and
stratonimbus.
International Standard Atmosphere (ISA)
Also known as the ICAO Standard Atmosphere, ISA is a standard against which to compare the
actual atmosphere at any point and time.
The ISA is based on the following values of pressure, density, and temperature at mean sea level
each of which decreases with increase in height:
 Pressure of 1013.2 millibar - Pressure is taken to fall at about 1 millibar per 30 feet in the lower
atmosphere (up to about 5,000 feet).
 Temperature of +15 °C - Temperature falls at a rate of 2 °C per 1,000 feet until the tropopause is
reached at 36,000 feet above which the temperature is assumed to be constant at -57 °C. (The
precise numbers are 1.98 °C, -56.5 °C and 36,090 feet)
 Density of 1,225 gm/m3.
The real atmosphere differs from ISA in many ways. Sea level pressure varies from day to day, and
there are wide extremes of temperature at all levels.
Variation in pressure, vertically and horizontally, affects the operation of the pressure altimeter.
The ISA standard sets a temperature of 15°C or 59°F at sea level pressure of 1013.25 hPa or 29.92 inches of mercury. It then
defines how temperature decreases as altitude increases at a standard rate of 1.98°C (3.5°F) per 1,000 feet up to 11,000 meters
(36,000 feet). Above this point, temperatures stabilize at -56.5°C or -69.7°F.

QNH, QFE, QNE


QNH = The pressure measured at station then reduced down to mean sea level pressure. When set on your
altimeter it will read your ALTITUDE. Sat on the tarmac at your airfield the altimeter will display the airfields
elevation above mean sea level.

This is the most commonly used pressure setting in the commercial world. Its probably the most useful setting
to have, as nearly all aviation references to elevation are in relation to mean sea level. The mountain peaks on a
map, airfield elevation, target elevation, minimum safe altitudes enroute etc. Incidently, QNH is given as a
regional pressure setting and should be updated with new ones if you leave its area of reference into a new
QNH pressure region. The QNH is the LOWEST FORECAST pressure at mean sea level for a given day to ensure
that safe terrain seperation is maintained regardless of the days variation in pressure.

QFE = Is mean sea level pressure corrected for temperature, adjusted for a specific site or datum like an airfield,
being the most obvious example. When this is set on your altimeter, it will read your HEIGHT not altitude. It
will read zero at airfield elevation and after take off will read your HEIGHT above that specific airfield. If you fly
to another airfield of different elevation and/or different QFE pressure, you will have to ensure you reset that
particular airfields QFE if you want your altimeter to read zero on touchdown.

QFE is very good for new pilots who are remaining in the circuit around an airfield and keeps things simple for
that task.

QFE Example: Airfield A with elevation 250ft above mean sea level. Airfield B elevation 300ft AMSL. A to B =
10miles. Assuming a uniform atmospheric pressure in the region.

Take off from A, altimeter reads 0ft on runway and after take-off reads HEIGHT above airfield A. Go and land at
B and your altimeter will read 50ft on the runway. This is because B's HEIGHT is 50ft higher then A.

In this example, if we set the regional QNH, then the altimeter will read ALTITUDE and therefore the airfields
altitude AMSL. Airfield A, altimeter will read 250ft. Airfield B will read 300ft. This is why QNH is the primary
pressure setting used in aviation at lower levels. It is far simpler working in a setting that gives ALTITUDE, so
you can reference your vertical position from everything on a map or chart. (All airfield plates (charts) have
their altitudes AMSL on the plate.)

This is all good and well knowing that QNH is the best pressure setting to use in a region for vertical situational
awareness. But it is not always possible to get the regional pressure setting QNH from accurate means and a
reliable network of meteo stations. Remote airfields and isolated combat zones are just 2 examples where it'd
be difficult to get an accurate QNH when you dont have access to good forecasts and numerous pressure
sensing stations.

If pressure info isn't available then you can get QFE easily by selecting an altimeter setting that reads zero on
the airfield. The number in the altimeter pressure window is your QFE.

To get QNH, you just need to know your elevation AMSL and set that in your altimeter. Airfield elevation =
250ft. Set altimeter to read 250ft. Pressure in the altimeter pressure window shows your QNH. (You have to
remember that this won’t be the lowest forecast QNH pressure for the day and just be cautious at low level. But
thats why a radio altimeter is handy!)

There are 2 other Q codes used for aviation pressure settings

QNE = the Internation Standard Atmosphere (ISA). It is the average mean sea level pressure around the globe.
It is planet earths mean atmospheric pressure at sea level basically. This pressure setting is refered to as
STANDARD in aviation. STANDARD is set from QNH when climbing up through the "Transition Level". Your
altimeter will then read your FLIGHT LEVEL. A reading of 25,000ft is FL250. 5,000ft = FL050. 13,500ft = FL135.

Summary. QNH = Altitude (AMSL) QNE = Flight Level. QFE = Height (AGL)

------------------------------------------------------------------------------------------------------------
Divide the airfield altitude in feet by 30 to get the number of millibars above
MSL. Add this to the QFE to get QNH or subtract it from QNH to get QFE

For example, the airfield elevation is 200 feet. Dividing by 30 gives us 6.66r. The QFE is 1023. Add 6.66 to
get 1029.66 and round up to 1030 millibars, which is the QNH.

-------------------------------------------------------------------------------------------------------------------

Now that we have accurate QFEs to use, it is useful to know that you can work out the QNH yourself
from that, as long as you know the altitude of the airport.

Useful Rule of thumb: 1mb is 30 feet.

Divide airport altitude (in feet) by 30 feet. Add resultant number of millibars onto the QFE that you are
given. Then you have QNH.

If you are doing Instrument approaches you would want to add a safety margin onto the DA because it
is a rule of thumb, not 100% accurate.

For Example

airport 600ft amsl. QFE reported as 990mb.

600 divided by 30 equals 20. Therefore air pressure is 20mb lower than at sea level. QFE provided is
990mb, therefore QNH is 1010mb.

How to get the QFE

To get to the figure mathematically, we need to remember that 1 mb = 30 feet.


Take the airfield elevation which in this example is 550 feet (for Popham airfield).
You then find divide that elevation, by 30.
550 ÷ 30 ≈ 18
Then, you take the 18 and take it away from the current QNH.
1024 – 18 = 1006 mb
That will give you your QFE.
FRONTS
Understanding the dynamics of weather fronts is one of the most critical things you need to
know as a pilot. A high percentages of accidents are caused by encounters with poor
weather, but with a good understanding of frontal systems, your weather briefing and flight
planning can be a whole lot more effective.

Cold Fronts

Cold fronts are cold, dense air masses that encounter warm, light air masses, pushing the
warm air up into the atmosphere. You can think of a cold front as a snowplow on a truck,
pushing the snow, or in this case warm air, up and out of the way.
Along the steep edge of a cold front, you'll often find cumulus and cumulonimbus clouds,
because of the rapidly rising warm air. This is why cold fronts are associated with squall
lines, thunderstorms, frontal turbulence, and overall bad weather like what's pictured below.

t8

Warm Fronts

When a faster-moving warm front encounters a slower cold front, the warm air pushes up and
over the colder air, because it's less dense. Lifting action with warm fronts is much more
gradual than what's found with cold fronts. But there's also some bad weather associated with
warm fronts.
Rain or other precipitation from a warm front falls into the colder air below, causing
widespread precipitation, fog, low ceilings/visibilities, and heavy snow (during colder months
of the year). And if the warm front moves slowly across the ground, you'll often find several
days of poor weather and IFR conditions.

Stationary Fronts

They're shown on a surface analysis chart with alternating cold and warm front symbols,
because the frontal boundary of a stationary front can be thought of as a tug of war game
between the cold and warm air masses, with characteristics from both.
Weather found along the front usually reflects the more dominant air mass. And while
thunderstorms are possible, you're more likely to find stratus clouds and steady, light rain or
drizzle. Stationary fronts tend to cover large areas in IMC, with calm surface winds that
parallel the frontal boundary.

Occluded Fronts

Let's just say occluded fronts aren't great for flying... They happen when cold fronts catch up
and overtake warm fronts ahead of them. The end result is two fronts in one area, one at the
surface, and one aloft/above the other. Since this type of front is so unstable, you often see
widespread cloudiness, precipitation, and thunderstorms (some of which may be embedded,
or hidden within clouds).

There are two types of occluded fronts: 1) In a cold front occlusion, cold air pushes
underneath a warm air mass, forcing it skyward. Just like in the diagram above. 2) On the
contrary, in a warm front occlusion, warmer air overruns colder air. But in both cases, warmer
air is lifted between the two air masses.

Fronts can be complex, but with a good understand of the weather that's associated with
each of them, you'll have a lot easier time planning your next flight, and understanding what
your weather briefer is telling you
Thunderstorm
Cumulonimbus (Cb) (Cb) clouds, in which Thunderstorms are found, form when three conditions are met:
 There must be a deep layer of unstable air.
 The air must be warm and moist.
 A trigger mechanism must cause the warm moist air to rise:
o Heating of the layer of air close to the surface.
o Rising ground forcing the air upwards (orographic uplift).
o A front forcing the air upwards.

A Cumulonimbus Cloud develops in three distinct phases:


1. Building Phase. A pocket of warm air begins to ascend as a result of one of the triggers mentioned
above. As the moist air rises, it becomes saturated, cloud forms, and the latent heat released as the
moisture condenses further warms the air and it continues to rise. The air within the cloud is warmer than
the air outside it and more air is drawn into the cloud from the base as well as the sides. The cloud grows
in height rapidly, faster than many aircraft can climb, extending from the surface to a great height,
sometimes as far as the Tropopause. as the temperature of the rising air drops below freezing point, the
water droplets become super cooled and join together to become larger and larger.
2. Mature Phase. As the top of the cloud reaches great heights, precipitation begins to fall. The
falling rain, snow and/or ice (Hail) cools the surrounding air creating downdrafts. The friction between ice
particles descending through the cloud, and ice particles being carried aloft by the updrafts, creates a
static charge in the cloud with the top of the cloud having a positive charge and the bottom of the cloud
having a negative charge. Eventually the difference in potential is so great that powerful electrical
discharges (Lightning) occur, accompanied by Thunder. The top of the cloud begins to flatten out
and Cirrus like cloud, consisting of ice crystals, spreads out creating a distinctive anvil shape.
3. Dissipation Phase. The cooling effect of the downdrafts on the air beneath the cloud reduces the
strength of the updrafts until the updrafts eventually stop and the lower cloud begins to dissipate. The
upper cloud will then linger for some time afterwards.
The active cycle of the Cb cell lasts little more than an hour but many Thunderstorms contain several
active Cb cells in various stages of development meaning that a storm can last several hours and extend
over a large area. The active cells are often embedded in a larger cloud mass consisting of the remains of
decayed cells as well as other cloud types at various levels. This can make the active cells very difficult to
detect visually and appropriate use of Radar is required to safely avoid active weather.

Frequently Asked Questions


1 - What are the 3 stages of a thunderstorm in aviation?
The three stages of a thunderstorm in aviation are the cumulus stage, the mature stage, and
the dissipating stage. The cumulus stage is characterized by the formation of cumulus clouds,
which grow vertically due to updrafts. The mature stage is characterized by the presence of
heavy rain, lightning, and strong winds. The dissipating stage is characterized by the
weakening of the storm and the dissipation of the clouds.

2 - How do pilots avoid thunderstorms during flights?


Pilots can avoid thunderstorms during flights by using weather radar and satellite imagery to
detect and avoid storm cells. They can also use air traffic control to receive updates on
weather conditions and to receive guidance on how to navigate around storms.

3 - What is the meaning of embedded thunderstorms in aviation?


Embedded thunderstorms are thunderstorms that are hidden within a larger cloud mass. They
are difficult to detect and can be dangerous for pilots because they can cause sudden
turbulence and icing conditions.

4 - How do pilots deal with scattered thunderstorms during a flight?


Pilots can deal with scattered thunderstorms during a flight by flying around them or through
gaps in the storm cells. They can also climb or descend to avoid the most severe parts of the
storm.

5 - Can flights be canceled due to scattered thunderstorms?


Flights can be canceled due to scattered thunderstorms if the conditions are too severe for
safe flying. Pilots and airlines prioritize passenger safety above all else, and if there is a risk
of severe turbulence, lightning strikes, or other dangerous conditions, the flight may be
canceled or delayed.

Dew Point
The Dew Point is the air temperature at which a sample of air would reach 100% humidity based
upon its current degree of saturation.
Description

Once the relative humidity of a mass of air becomes 100%, then if the temperature falls it cannot hold
all of the water vapour within it. The excess water vapour will then condense into cloud or fog or, if in
contact with objects on or near the ground, will form dew or hoar frost.

When observing successive weather reports (METARs), a reducing gap between the actual
temperature and the Dew Point temperature gives an indication of impending low visibility conditions
and the possibility of fog.
Severity of Turbulence
For the purpose of reporting and forecasting of air turbulence, it is graded on a relative scale, according to
its perceived or potential effect on a 'typical' aircraft, as Light, Moderate, Severe and Extreme.
 Light turbulence is the least severe, with slight, erratic changes in attitude and/or altitude.
 Moderate turbulence is similar to light turbulence, but of greater intensity - variations in speed as well as
altitude and attitude may occur but the aircraft remains in control all the time.
 Severe turbulence is characterised by large, abrupt changes in attitude and altitude with large variations in
airspeed. There may be brief periods where effective control of the aircraft is impossible. Loose objects
may move around the cabin and damage to aircraft structures may occur.
 Extreme turbulence is capable of causing structural damage and resulting directly in prolonged, possibly
terminal, loss of control of the aircraft.
In-flight turbulence assessment is essentially subjective. Routine encounters involve light or moderate
turbulence, although to inexperienced passengers (or pilots), especially in small aircraft, these conditions
may seem to be severe.
The perception of turbulence severity experienced by an aircraft depends not only on the strength of the
air disturbance but also on the size of the aircraft - moderate turbulence in a large aircraft may appear
severe in a small aircraft. Therefore pilot reports of turbulence should mention the aircraft type to aid
assessment of the relevance to other pilots in, or approaching, the same area.

Reporting of Turbulence

1) Light Chop

Light chop is defined as a slight, rapid, and somewhat rhythmic bumpiness. In your plane, not a lot
changes, you don't experience any large deviations in altitude or attitude.

2) Light Turbulence

Light turbulence is a series of momentary, slightly erratic changes in your altitude or attitude. When
you're in light turbulence, you might feel a slight strain against your seat belt or shoulder straps. Small
unsecured objects might get dislodged in your plane. If you could, it would be easy to walk around the
cabin with little or no difficulty (except for the fact if you're in a GA airplane, you probably can't walk
around anyway).

3) Moderate Chop

Moderate chop is similar to light chop, but it's more intense. Moderate chop has consistent bumps or
jolts, with little to no change in altitude or attitude.
4) Moderate Turbulence

Moderate turbulence consists of changes in your altitude or attitude, but your aircraft remains in
positive control at all times. You'll feel a definite strain against your seat belt or shoulder straps. And if
you have unsecured objects in the cabin, they'll become dislodged. If you could, walking will be
difficult in your cabin.

5) Severe Turbulence

Severe turbulence consists of large, abrupt changes in altitude or attitude. Your aircraft may be
temporarily out of control, and you'll be forced violently against your seat belt.

6) Extreme Turbulence

In extreme turbulence, your aircraft is violently tossed about and practically impossible to control.
Extreme turbulence can cause structural damage or structural failure.

How to Report Turbulence to ATC

When you report turbulence to ATC, it's important to include the turbulence frequency as well. Here's
what you should tell them.

 Occasional: Less than 1/3 of the time.

 Intermittent: 1/3 to 2/3 of the time.

 Continuous: More than 2/3 of the time.

So there you have it! The next time you're flying through turbulence, you'll know how to give a pilot
report to ATC like a pro.

According to the Aeronautical Information Manual (AIM) there are 4 major categories of
turbulence: Light, Moderate, Severe and Extreme.
TAF
In meteorology and aviation, terminal aerodrome forecast (TAF) is a format for reporting weather
forecast information,[1] particularly as it relates to aviation. TAFs are issued at least four times a day,
every six hours, for major civil airfields: 0000, 0600, 1200 and 1800 UTC, and generally apply to a
24- or 30-hour period, and an area within approximately five statute miles (8.0 km) (or 5 nautical miles
(9.3 km) in Canada) from the center of an airport runway complex. TAFs are issued every three hours
for military airfields and some civil airfields and cover a period ranging from 3 hours to 30 hours.
TAFs complement and use similar encoding to METAR reports. They are produced by a human
forecaster based on the ground. For this reason, there are considerably fewer TAF locations than
there are airports for which METARs are available. TAFs can be more accurate than numerical
weather forecasts, since they take into account local, small-scale, geographic effects.
METAR
is a format for reporting weather information. A METAR weather report is predominantly used
by aircraft pilots, and by meteorologists, who use aggregated METAR information to assist in weather
forecasting. Today, according to the advancement of technology in civil aviation, the METAR is sent
as IWXXM model.
Raw METAR is the most common format in the world for the transmission of observational weather
data. It is highly standardized through the International Civil Aviation Organization (ICAO), which
allows it to be understood throughout most of the world.

SIGMET
SIGMET, or Significant Meteorological Information (AIM 7-1-6), is a severe weather advisory that
contains meteorological information concerning the safety of all aircraft. Compared to AIRMETs,
SIGMETs cover more severe weather. Today, according to the advancement of technology in civil
aviation, the SIGMET is sent as IWXXM model.
Types

There are three main types of internationally recognized SIGMETs per ICAO:

 Volcanic ash (VA or WV SIGMET)


 Tropical Cyclone (TC SIGMET)
 Other En-route weather phenomenon (WS SIGMET), which may consist of
o Thunderstorm types
o Turbulences types
o Mountain waves
o Icing/Sleet/Hail
o Dust or sand storms
o Radioactive Cloud
This information is usually broadcast on the ATIS at ATC facilities, as well as over VOLMET stations.
They are assigned an alphabetic designator from N through Y (excluding S and T). SIGMETs are
issued as needed, and are valid up to four hours. SIGMETS for hurricanes and volcanic ash outside
the CONUS are valid up to six hours.

PIREP
A pilot report or PIREP is a report of actual flight or ground conditions encountered by an aircraft.
Reports commonly include information about atmospheric conditions
(like temperature, icing, turbulence) or airport conditions (like runway condition codes or ground
equipment failures). This information is usually relayed by radio to the nearest ground station, but
other options (e.g. electronic submission) also exist in some regions. [1] The message would then be
encoded and relayed to other weather offices and air traffic service units.
Although the actual form used to record the PIREP may differ from one country to another, the
standards and criteria will remain almost the same. At a minimum the PIREP must contain a header,
aircraft location, time, flight level, aircraft type and one other field.
In recent years, a PIREP will also include UA or UUA used to identify the PIREP as routine or urgent.

AIRMET
An AIRMET, or Airmen's Meteorological Information, is a concise description of weather
phenomena that are occurring or may occur (forecast) along an air route that may
affect aircraft safety. Compared to SIGMETs, AIRMETs cover less severe weather:
moderate turbulence and icing, sustained surface winds of 30 knots or more, or widespread restricted
visibility. Today, according to the advancement of technology in civil aviation, the AIRMET is sent
as IWXXM model.

NOTAM
A NOTAM (Notice to Airmen, Notice to Air Men, Notice to Airman]or Notice to Air Missions is a
notice filed with an aviation authority to alert aircraft pilots of potential hazards along a flight route or
at a location that could affect the flight.[6] NOTAMs are notices or advisories that contain information
concerning the establishment, conditions or change in any aeronautical facility, service, procedure or
hazard, the timely knowledge of which may be essential to personnel and systems concerned with
flight operations.[7]
NOTAMs are created and transmitted by government agencies and airport operators under guidelines
specified by Annex 15: Aeronautical Information Services of the Convention on International Civil
Aviation (CICA). A NOTAM is filed with an aviation authority to alert aircraft pilots of any hazards en
route or at a specific location, or Flight Information Region. The authority, in turn, provides a means of
disseminating relevant NOTAMs to pilots.
TREND

TREND forecasts are indicated by BECMG (Becoming) or TEMPO (temporary) which may
be followed by a time group (hours and minutes UTC) preceded by one of the letter
indicated FM (from), TL (until), AT (at)
e.g. BECMG FM1030 TL1130
NOSIG replaces the TREND group when no significant changes are forecast to occur
during the 2 hour forecast period.
To indicate the end of significant weather the abbreviation NSW (No Significant Weather)
is used.
Only those elements for which a significant change is expected should be included in a
TREND
What are the differences between METARs and TAFs?
A METAR is Meteorological Terminal Air Report (or “aviation routine weather report”) is a weather report of a actual
conditions at an airport at a specific time.

A TAF (Terminal Area Forecast) is a weather forecast for a given area around the airport specified in the forecast.

Therefore the main difference between the two is that METARs are a record of actual conditions where TAFs are a prediction
of future weather.

Another difference is that METARs may include a “remarks”(RMK) section. Remarks may include more accurate
temperature, sea level pressure (SLP) or runway state group (for example if the runway is contaminated by snow).

TAFs, METARs and SPECIs use a common code to describe the weather such as TS for thunderstorms or SHRA for rain
showers.

Frequently Asked Questions about METARs.


Are winds in METARs in degrees true or magnetic?
Wind direction in METARs and TAFs are in degrees true.

What does the term METAR mean?


A METAR (Meteorological Terminal Air Report) is a routine aviation weather report of actual
observed conditions at an airport.

Where can you get METARs?


We can get METARS in many places, for example: if airborne we can receive them
via ACARS/Datalink, they can be accessed online (for example at the NOAA Aviation Weather
Center) or through apps such as ForeFlight or AeroWeather.
When are METARs issued?
In the United States METARs are usually issued once per hour, at 55 minutes past the hour. In
other States they are often issued twice per hour, at 20 minutes past, and 50 minutes past the hour.

What does NSC in a METAR mean?


NSC stands for “No Significant Cloud” and is used if no clouds have been observed.

What does NCD in a METAR mean?


NCD stands for “no cloud discernible” and is issued if no cloud is detected by an automatic
reporting station.

What is the meaning of NDV in a METAR?


NDV means No Directional Variation and is used when visibility has been recorded automatically
by only one sensor. As only one sensor has been used it cannot detect any variation in the visibility

What does 9999 mean in a METAR?


9999 refers to a visibility of 10,000m or more.
What does a V mean in a METAR?
V means “variable” and refers to wind direction. For example, a METAR that reads 030/15
010V060 means the predominant wind is from 030º at 15 knots, variable in direction from 010º to
060º.

What is the difference between SKC and CLR?


Both SKC and CLR mean clear skies with SKC used by manual stations and CLR issued by
automatic stations indicating no cloud layers detected at or below 12,000 feet.

What does VV or VV/// mean in a METAR?


VV refers to “vertical visibility” and means the sky is obscured by surface-based phenomenon (e.g.
fog); the three digits that follow VV indicate the height in hundreds of feet. VV/// means the
vertical visibility is of an indefinite height.

What does a P mean in a METAR?


A P indicates greater than the highest reportable value. For example, P6SM means a visibility of
greater than 6 statute miles.

What is CAVOK in aviation weather?


CAVOK stands for “ceiling and visibility OK” and means no cloud below 5000 feet, no
cumulonimbus (CB) or towering cumulus (TCU) at any level, and visibility 10 KM/6 SM or
greater.

What is the meaning of VC in METARs?


VC stands for “vicinity” and refers to weather that is near the airport, but not at it. For example,
VCSH means rain showers in the vicinity of the airport.
“Weather phenomena occurring beyond the point of observation (between 5 and 10 sm) are coded
as in the vicinity (VC)” [FAA Advisory Circular -Aviation Weather Services 3.1.5.8.2]

Are cloud heights in METARs AGL or MSL?


METARs refer to heights above the airport, i.e. AGL.

What does SPECI mean in a METAR?


SPECI refers to a special update of a METAR when there has been a significant change since the
METAR was issued. Find out more about SPECIs here.
How often are TAFs issued?
TAFs are issued 4 times a day, at 6-hourly intervals. 0500z, 1100z, 1700z and 2300z, and are valid
for 30 hours.

What is a “ceiling” in aviation?


A ceiling is when the sky is obscured by five eighths (i.e. 5 octas) of the sky or more i.e. Broken
(BKN) or Overcast (OVC) cloud.

When are TAFs issued?


TAFs are normally issued 4 times a day beginning at 0500Z and reissued every 6 hours (0500z,
1100z, 1700z, 2300z).
When are METARs issued?
METARs in the U.S. are generally issued once per hour on the hour, and internationally often twice
per hour – at 20 minutes last the hour, and 50 minutes past.

The Runway Condition Report (RCR)


An eight digit telegraphic code on runway conditions for some European airports may be
included at the end of hourly METAR messages:
Eight Figure Group
1st two digits Runway designator
3 digit
rd
Runway deposits
4 digit
th
Extent of runway contamination
5th and 6th digits Depth of deposit
7 and 8 digits
th th
Friction coefficient or braking action

The first two digits correspond to the runway designator. For parallel runways LEFT is
indicated by the designator only (18L would be displayed as 18) and RIGHT has 50 added
(18R would be displayed as 68). When all runways are affected the figure group 88 will
be used. If 99 appears as the first two digits the information is a repetition of the last
message because no new message has been received in time for transmission.

Runway Deposits Extent of Runway Contamination Depth of Deposit


0 Clear & Dry 1 <10% contaminated (covered) 00 Less than 1mm
1 Damp 2 11% to 25% contaminated 01-90 Measurement in mm
2 Wet or water particles (covered) 92 10cm
3 Rime or frost covered (normally > 5 26%-50% contaminated (covered) 93 15cm
1mm)
4 Dry Snow 9 51%-100% contaminated 94 20cm
5 Wet Snow (covered) 95 25cm
6 Slush / Not reported (runway clearance in 96 30cm
7 Ice progress) 97 35cm
8 Compacted or rolled snow 98 40cm or more
9 Frozen ruts or ridges 99 Runway not operational due to
snow, slush, ice, large drifts or
runway clearance, depth not
reported
/ Not reported (runway clearance in // Not operationally significant or
progress) not measurable

Note: the quoted depth is the mean of a number of reading or if operationally significant
the greatest depth measured.

Friction Coefficient or Braking Action (7th and 8th digits)


28 Friction coefficient 0.28
35 Friction coefficient 0.35
91 Braking action poor
92 Braking action medium to poor
93 Braking action medium
94 Braking action medium to good
95 Braking action good
99 Figures unreliable
// Braking action not reported or runway not operations or airport
closed.

Note: Where braking action is assessed at a number of points along the runway the mean
value will be transmitted or if operationally significant the lowest value.
If measuring equipment does not allow measurement of friction with satisfactory reliability
(such as contaminated by wet snow, slush or loose snow) the figure 99 will be used.

Relative Humidity (RH)


Relative humidity is the ratio, in percentage, of the actual water vapor present in the air
to the maximum it can contain at the same temperature and pressure.

How does humidity affect aircraft performance?


 High humidity negatively affects aircraft performance and flight planning
by decreasing air pressure and reducing lift, requiring longer runway distances.
Wet air also impacts engine performance by reducing the amount of oxygen
available for combustion, resulting in less thrust generated and affecting
acceleration and climbing ability.

Contrails: What are they and how do they form?


Contrails, or condensation trails, are essentially human-made clouds; they are trails of
condensed water vapor created by jet engines, according to the National Weather Service.
We most commonly see them behind planes at cruising altitude, but they can also be emitted
by rockets.

Contrails are created when the hot water vapor emitted by a jet engine after combustion cools and
condenses in Earth's atmosphere, according to the Environmental Protection Agency (EPA). The atmosphere's
temperature and humidity must be in just the right place for condensation to occur — the air must be
cold with some humidity.

Different Phases of Matter


 Freezing: Change of a substance from liquid phase to solid
 Melting: Change from solid phase to liquid
 Vaporization: Change from liquid to gaseous form
 Condensation: Change from gas to liquid form
 Sublimation: Change from solid to a gas without becoming a liquid
 Deposition: Change from gas to solid without becoming a liquid
 Ionization: Change from a gas phase to plasma (ionized particles)
 Recombination: Change from plasma to gas
CONVERSION
Example
We are going to convert a speed of 50 km/h in knots.
Speed = 50 km/h
1 km/h = 0.539957 knots
Speed = 50 km/h = 50 x 0.539957 = 27 knots

How to Convert Mile to Kilometer


1 mi, mi(Int) = 1.609344 km
1 km = 0.6213711922 mi, mi(Int)
Example: convert 15 mi, mi(Int) to km:
15 mi, mi(Int) = 15 × 1.609344 km = 24.14016 km

 1 knot = 1 nautical mile per hour = 6076 feet per hour 1 mph = 1 mile per
hour = 5280 feet per hour To convert a kilometer per hour measurement to a
knot measurement, multiply the speed by 0.539957 (approximately).
 1 hectopascals = 25.879322442072
What is a microburst?
A microburst is a downdraft (sinking air) in a thunderstorm that is less than 2.5 miles in scale.
Some microbursts can pose a threat to life and property, but all microbursts pose a significant
threat to aviation. Although microbursts are not as widely recognized as tornadoes, they can cause
comparable, and in some cases, worse damage than some tornadoes produce. In fact, wind
speeds as high as 150 mph are possible in extreme microburst cases.

VIRGA
In meteorology, a virga, also called a dry storm,
is an observable streak or shaft of precipitation that evaporates or sublimates before reaching the
ground. A shaft of precipitation that does not evaporate before reaching the ground is a precipitation
shaft. At high altitudes, precipitation falls mainly as ice crystals before melting and finally evaporating.
That is often due to compressional heating, because air pressure increases closer to the ground.
Virga is very common in deserts and temperate climates. In North America, it is commonly seen in
the Western United States and the Canadian Prairies. It is also very common in the Middle
East, Australia, and North Africa.
Virgae can cause varying weather effects because, as rain is changed from liquid to vapor form, it
removes significant amounts of heat from the air due to water's high heat of vaporization.
Precipitation falling into these cooling downdrafts may eventually reach the ground. In some
instances these pockets of colder air can descend rapidly, creating a wet or dry microburst which can
be extremely hazardous to aviation. Conversely, precipitation evaporating at high altitude can
compressionally heat as it falls, and result in a gusty downburst which may substantially and rapidly
warm the surface temperature. This fairly rare phenomenon, a heat burst, also tends to be of
exceedingly dry air.
Virgae also have a role in seeding storm cells. That is because small particles from one cloud are
blown into neighboring supersaturated air and act as nucleation particles for the
next thunderhead cloud to begin forming.
ATMOSPHERE
What are the 5 components of air?
The main components of air are:
 Nitrogen (78%)
 Oxygen (21%)
 Carbon dioxide (0.04%)
 Argon (0.93%)
 Trace amounts of helium, neon, methane, hydrogen and water vapour.

 By volume, the dry air in Earth’s atmosphere is about 78.08 percent nitrogen, 20.95 percent
oxygen, and 0.93 percent argon.

A brew of trace gases accounts for the other approximately 0.04 percent, including the greenhouse
gases carbon dioxide, methane, nitrous oxide and ozone.
RVR
Runway Visual Range (RVR) is the range over which the pilot of an aircraft on the centre line of a runway
can see the runway surface markings or the lights delineating the runway or identifying its centre line.
Source: ICAO Annex 3: Meteorological Service for International Air Navigation

Description

RVR is not an observation or a measurement of a meteorological parameter (such as e.g. surface


wind direction and speed, temperature and pressure). It is an assessment, based on calculations that take
into account various elements, including atmospheric factors such as extinction coefficient of
the atmosphere, physical/biological factors such as visual threshold of illumination, and operational factors
such as runway light intensity.
The main purpose of RVR is to provide pilots, controllers and other aeronautical users with information on
runway visibility conditions during periods of low visibility, due to e.g. fog, precipitation or a sandstorm. It is
required to assess whether the conditions are above or below the specified operating minima for take-off
and landing. It is to be noted that for this purpose RVR values supersede the reported visibility and that in
the case of precision approaches it is normally not permissible to start an approach if the applicable RVR
value(s) is below the required minimum.
RVR information is provided when the visibility or RVR for a particular runway is below 1500 m. It is
representative for:
 The touchdown zone, when the runway is intended for non-precision approaches or Category
I operations.
 The touchdown zone and the mid-point, if the runway is intended for Category II operations.
 The touchdown zone, mid-point and stop-end, if the runway is intended for Category III operations.
RVR information is included in local routine reports, local special reports, METAR and SPECI.

Wind Shear
Runway Visual Range (RVR) is the range over which the pilot of an aircraft on the centre line of a runway
can see the runway surface markings or the lights delineating the runway or identifying its centre line.
Source: ICAO Annex 3: Meteorological Service for International Air Navigation
Description

RVR is not an observation or a measurement of a meteorological parameter (such as e.g. surface


wind direction and speed, temperature and pressure). It is an assessment, based on calculations that take
into account various elements, including atmospheric factors such as extinction coefficient of
the atmosphere, physical/biological factors such as visual threshold of illumination, and operational factors
such as runway light intensity.
The main purpose of RVR is to provide pilots, controllers and other aeronautical users with information on
runway visibility conditions during periods of low visibility, due to e.g. fog, precipitation or a sandstorm. It is
required to assess whether the conditions are above or below the specified operating minima for take-off
and landing. It is to be noted that for this purpose RVR values supersede the reported visibility and that in
the case of precision approaches it is normally not permissible to start an approach if the applicable RVR
value(s) is below the required minimum.
RVR information is provided when the visibility or RVR for a particular runway is below 1500 m. It is
representative for:
 The touchdown zone, when the runway is intended for non-precision approaches or Category
I operations.
 The touchdown zone and the mid-point, if the runway is intended for Category II operations.
 The touchdown zone, mid-point and stop-end, if the runway is intended for Category III operations.
RVR information is included in local routine reports, local special reports, METAR and SPECI.

What is Windshear and How to Avoid It ?


Flight crew awareness and alertness are key factors in the successful application of windshear avoidance and
escape / recovery techniques
This safety article is an overview of operational recommendations and training guidelines for aircraft
operation in forecast or suspected windshear or downburst conditions.
Statistical Introduction
Adverse weather (other than low visibility and runway condition) is a circumstantial factor in nearly 40 percent
of approach-and-landing accidents.
Adverse wind conditions (i.e., strong cross winds, tailwind and windshear) are involved in more than 30
percent of approach-and-landing accidents and in 15 percent of events involving CFIT.
Windshear is the primary causal factor in 4 percent of approach-and-landing accidents and is the ninth cause
of fatalities.

What is the Influence of Windshear on Aircraft Performance ?


The flight performance is affected as:
– Headwind gust instantaneously increases the aircraft speed and thus tends to make the aircraft fly above
intended path and/or accelerate,
– A downdraft affects both the aircraft Angle-Of-Attack (AOA), that increases, and the aircraft path since it
makes the aircraft sink
– Tailwind gust instantaneously decreases the aircraft speed and thus tends to make the aircraft fly below
intended path and/or decelerate.
Windshears associated to jet streams, mountain waves and frontal surfaces usually occur at altitudes that do
not present the same risk than microbursts, which occur closer to the ground.
Aquaplaning/Hydroplaning
Aquaplaning, also known as hydroplaning, is a condition in which standing water, slush or snow, causes
the moving wheel of an aircraft to lose contact with the load bearing surface on which it is rolling with the
result that braking action on the wheel is not effective in reducing the ground speed of the aircraft.
Interestingly, Water Skiing was originally called Aquaplaning and was carried out on a simple board which
underlines that the physics of aquaplaning works at very low speeds by aviation standards. Aquaplaning is
also highly relevant to cars at speeds as low as 40 mph.
The continued incidence of aquaplaning reduces the braking co-efficient to that of an icy or "slippery"
runway - less than 20% of that on an equivalent dry runway.

Causal Factors
A layer of water builds up beneath the tyre in increasing resistance to displacement by the pressure of the
wheel. Eventually, this results in the formation of a wedge between the runway and the tyre. This
resistance to water displacement has a vertical component which progressively lifts the tyre and reduces
the area in contact with the runway until the aircraft is completely water-borne. In this condition, the tyre is
no longer capable of providing directional control or effective braking because the drag forces are so low.
If such a runway surface state prevails, then flight crew are required to make their aircraft runway
performance calculations using "slippery runway" data; this specifically allows for poor deceleration. They
must also take account of crosswind component limits in the AFM which make allowance for less assured
directional control.
Aquaplaning can occur when a wheel is running in the presence of water; it may also occur in certain
circumstances when running in a combination of water and wet snow. Aquaplaning on runway surfaces
with normal friction characteristics is unlikely to begin in water depths of 3mm or less. For this reason, a
depth of 3mm has been adopted in Europe as the means to determine whether a runway surface is
contaminated with water to the extent that aircraft performance assumptions are liable to be significantly
affected. Once aquaplaning has commenced, it can be sustained over surfaces and in water depths which
would not have led to its initiation.
In the case of the most common type of aquaplaning, called dynamic aquaplaning (see below), a simple
formula (Horne's formula) exists for calculating the minimum groundspeed for initiation of this type of
aquaplaning on a sufficiently wet runway based upon tyre pressure where V = groundspeed in knots and
P = tyre inflation pressure in psi:
V = 9 x √P
This formula is based upon the validation of hydrodynamic lift theory by experimental evidence. For many
modern tires the constant maybe closer to 6 or 7 rather than 9. With a typical tyre pressure of about 150
psi, √P will be 12.25 so aquaplaning is possible down to about 70 knots. The effect of the relationship
demonstrated is that most jet aircraft, even relatively small ones, have a significant ‘window’ for the
initiation of dynamic aquaplaning during a landing near to the maximum approved weight and an even
larger one in the case of a high speed rejected take off. It assumes that tyre pressure and tread depth are
both within allowable AMM limitations.

How To Prevent Viscous Hydroplaning


 Land on a grooved runway, if possible.
 Don't land fast.
 Keep your tires inflated. Under-inflated tires hydroplane easier than properly inflated ones.
 Use back pressure and aerodynamic braking to slow down, and use light brake pressure.

Time of Useful Consciousness (TUC)


The Time of Useful Conciousness (TUC) or Effective Performance Time is the period of elapsed time from
the interruption of normal air supply or exposure to an oxygen-poor environment until the time when the
ability to function usefully is likely to be lost at which point an affected individual would no longer be
capable of taking normal corrective or protective action.

The table below shows average TUCs as documented by the Federal Aviation Administration; a rapid
ascent results in a lower TUC.[5] The TUCs for any given individual may differ significantly from
this. Aerobic exercise during the TUC period will reduce the TUCs considerably; so will exercise
immediately prior to the TUC as this induces an oxygen debt prior to exposure.

Altitude
TUC (normal ascent) TUC (rapid decompression)
(measured barometrically)

FL180 (18,000 ft; 5,500 m) 20 to 30 minutes 10 to 15 minutes

FL220 (22,000 ft; 6,700 m) 10 minutes 5 minutes

FL250 (25,000 ft; 7,600 m) 3 to 5 minutes 1.5 to 3.5 minutes

FL280 (28,000 ft; 8,550 m) 2.5 to 3 minutes 1.25 to 1.5 minutes

FL300 (30,000 ft; 9,150 m) 1 to 2 minutes 30 to 60 seconds

FL350 (35,000 ft; 10,650 m) 30 secs to 1 minute 15 to 30 seconds

FL400 (40,000 ft; 12,200 m) 15 to 20 seconds 7 to 10 seconds

FL430 (43,000 ft; 13,100 m) 9 to 12 seconds 5 to 6 seconds

FL500 (50,000 ft; 15,250 m) 8 to 10 seconds 5 seconds


Balanced field takeof
For a given aircraft weight, engine thrust, aircraft configuration, and runway condition, the shortest
runway length that complies with safety regulations is the balanced field length.
The takeoff decision speed V1 is the fastest speed at which the pilot must take the first actions to
reject the takeoff (e.g. reduce thrust, apply brakes, deploy speed brakes). At speeds below V 1 the
aircraft can be brought to a halt before the end of the runway. At V 1 and above, the pilot should
continue the takeoff even if an emergency is recognized. The speed will ensure the aircraft achieves
the required height above the takeoff surface within the takeoff distance.
To achieve a balanced field takeoff, V 1 is selected so the take-off distance with one engine
inoperative, and the accelerate-stop distance, are equal. When the runway length is equal to the
balanced field length only one value for V 1 will exist. Aviation regulations (for transport
category aircraft) require the takeoff distance with one engine inoperative to be no greater than the
take-off distance available (TODA); and the accelerate-stop distance to be no greater than the
accelerate-stop distance available (ASDA).
On runways longer than the balanced field length for the aircraft weight the operator may be able to
choose V1 from a range of speeds if adequate information is supplied by the aircraft manufacturer.
The slowest speed in this range will be determined by the Take Off Distance Available (TODA). For a
low V1, if an engine fails just above V 1, the acceleration to VR on one engine will take more distance.
Whereas, if an engine fails before a low V 1, it will take less distance to stop, so the Accelerate Stop
Distance Required (ASDR) is lower. By contrast, the fastest speed in this range will be determined by
the Accelerate Stop Distance Available (ASDA). If an engine fails above a high V 1, it will take less
distance to reach VR, so Take Off Distance Required (TODR) is lower. Whereas, if an engine fails just
below a high V1, it will take more distance to stop, so the Accelerate Stop Distance Required is
greater.
Alternatively, on runways longer than the balanced field length the pilot can use reduced thrust,
resulting in the balanced field length again being equal to the runway length available.
Factors affecting the balanced field length include:

 the mass of the aircraft – higher mass results in slower acceleration and higher takeoff speed
 engine thrust – affected by temperature and air pressure, but reduced thrust can also be
deliberately selected by the pilot
 density altitude – reduced air pressure or increased temperature increases minimum take off
speed
 aircraft configuration such as wing flap position
 runway slope and runway wind component
 runway conditions – a rough or soft field slows acceleration, a wet or icy field reduces braking
Technology

Calculation of the balanced field length traditionally involves relying on an expansion program model,
where the various forces are evaluated as a function of speed, and step-wise integrated, using an
estimate for V1. The process is iterated with different values for the engine failure speed until the
accelerate-stop and accelerate-go distances are equal. This process suffers from the inherently slow
and repetitive approach, which is also subject to round-off errors if the speed increment between the
steps is not carefully selected, which could cause some issues in first principle aircraft performance
models provided to airlines for day-to-day operations. Alternate approaches using a more
mathematically complex but inherently more accurate and faster algebraic integration method have
however been developed.
Landing and Takeoff Performance Monitoring Systems are devices aimed at providing the pilot with
information on the validity of the performance computation, and averting runway overruns that occur
in situations not adequately addressed by the takeoff V-speeds concept.
------------------------------------------------------------------------------------------------------------------
The balanced field length is the length of the runway required for an aircraft to accelerate to takeoff speed,
experience an engine failure at a critical point, and either continue the takeoff or stop safely on the remaining
runway. In simpler terms, it is the distance needed for an aircraft to either take off or abort the takeoff in the
event of an engine failure during the critical phase of the flight.

The Importance of Balanced Field Length


Understanding the concept and significance of balanced field length is crucial for pilots, flight planners, and
airport authorities. It directly impacts the safe operation of aircraft and helps determine the maximum allowable
takeoff weight, taking into account various factors such as runway length, aircraft performance, and
environmental conditions.

One of the primary purposes of calculating the balanced field length is to ensure that an aircraft has adequate
runway distance to either commit to takeoff or abort it in the event of an engine failure. This allows pilots to
make informed decisions during critical moments, prioritizing the safety of passengers and crew on board.

Additionally, understanding the balanced field length assists in determining the maximum payload an aircraft
can carry for a given runway length and environmental conditions. By knowing the precise length required for a
successful takeoff or safe abort, flight planners can optimize load capacities and fuel quantities, maximizing
operational efficiency.

Moreover, considering balanced field length is crucial for airport authorities when designing, constructing, and
maintaining runways. It ensures that runways are of sufficient length to accommodate a wide range of aircraft
types and sizes, allowing for safe and efficient operations.

Calculation of Balanced Field Length

The balanced field length is derived through a complex calculation that takes into account several variables,
including but not limited to:

 Aircraft weight
 Aircraft performance data
 Runway slope
 Wind speed and direction
 Temperature and atmospheric conditions
 Flap settings
 Anti-icing requirements

By considering these variables, pilots and flight planners can accurately determine the balanced field length for
a given flight scenario. The calculation ensures that an aircraft has sufficient runway distance to safely take off
or abort the takeoff if an engine failure occurs during the most critical phases, such as the takeoff roll and initial
climb.

It is worth noting that the calculation of balanced field length is a regulatory requirement and varies depending
on the jurisdiction’s regulations and the type of aircraft being operated. These regulations aim to ensure
consistent safety standards are met across the aviation industry.

Implementation of Balanced Field Length

Once the balanced field length has been calculated, it is essential to incorporate this information into the
operational planning and decision-making process. Pilots must have accurate and up-to-date data regarding the
required runway length for takeoff and the distance available for aborting the takeoff if necessary.

Flight planners play a crucial role in ensuring that the aircraft’s performance capabilities align with the available
runway length and environmental conditions. They must optimize the payload, fuel, and other variables to
ensure a safe takeoff or abort capability during the most critical phases of flight.

To support pilots and flight planners, aviation authorities and manufacturers provide comprehensive
performance charts and software tools that assist in calculating the balanced field length for various scenarios.
These tools factor in the specific aircraft’s performance capabilities, environmental conditions, and regulatory
requirements to provide accurate and reliable data.

It is crucial for pilots and flight planners to regularly update the balanced field length calculations based on the
latest performance data and any changes in aircraft configuration, such as modifications or repairs. This ensures
that the most accurate and
relevant information is
available for decision-making
during critical flight phases.

Aircraft Oxygen Systems


National regulations for the provision and use of supplemental or emergency oxygen systems are based on the
guidance provided in Annex 6 of the International Civil Aviation Organisation (ICAO) Standards and
Recommended Practices (SARPS). In general terms, this guidance first differentiates between pressurised and
non-pressurised aircraft and then provides specific requirements based on the altitude at which flight is to be
conducted. Some of the more salient items found in the ICAO guidance on oxygen are as follows:
 All Aircraft
o An operator shall ensure that passengers are made familiar with the location and use of: ... d) oxygen dispensing
equipment, if the provision of oxygen for the use of passengers is prescribed...
 Non-pressurised Aircraft
o An aeroplane intended to be operated at flight altitudes at which the atmospheric pressure is less than 700 hPa
(see Note 1) in personnel compartments shall be equipped with oxygen storage and dispensing apparatus
o A flight to be operated at flight altitudes at which the atmospheric pressure in personnel compartments will be
less than 700 hPa shall not be commenced unless sufficient stored breathing oxygen is carried to supply: a) all
crew members and 10 per cent of the passengers for any period in excess of 30 minutes that the pressure in
compartments occupied by them will be between 700 hPa and 620 hPa ; and b) the crew and passengers for any
period that the atmospheric pressure in compartments occupied by them will be less than 620 hPa
 Pressurised Aircraft
o An aeroplane intended to be operated at flight altitudes at which the atmospheric pressure is less than 376 hPa
or which, if operated at flight altitudes at which the atmospheric pressure is more than 376 hPa , cannot descend
safely within four minutes to a flight altitude at which the atmospheric pressure is equal to 620 hPa ... shall be
provided with automatically deployable oxygen equipment. The total number of oxygen dispensing units shall
exceed the number of passenger and cabin crew seats by at least 10 per cent.
o All flight crew members of pressurised aeroplanes operating above an altitude where the atmospheric pressure
is less than 376 hPa shall have available at the flight duty station a quick-donning type of oxygen mask which
will readily supply oxygen upon demand.
Note 1: Approximate hPa-altitude equivalents: 700 hPa = 10,000', 620 hPa = 13,000', 376 hPa = 25,000'
Note 2: National or Regional Authorities use the ICAO guidance as the basis for their regulations. However,
these regulations may be more or less restrictive than the SARPS. Consult the appropriate documentation
provided by the aircraft State of Registry for specific criteria.
Equipment
Flight Deck

 Oxygen for the use of the flight deck occupants is normally stored as pressurised gas in one or more tanks or
cylinders. In certain aircraft types, oxygen is stored as a liquid (LOX).
o The total oxygen capacity must be sufficient to supply all flight deck occupants with adequate oxygen for a
defined period of time at an altitude profile specified in the applicable National Aviation Authority regulations.
Commonly, the altitude profile will incorporate an emergency descent segment and followed by a period in
level flight at a defined altitude.
o A quantity gauge or other means of determining the amount of available oxygen will be incorporated.
o If a LOX system is installed, a LOX converter, which facilitates the transformation of the oxygen from a liquid
to a gaseous state, will also be installed.
 A regulator is installed to reduce storage cylinder pressure to a usable level. Depending upon the aircraft type,
regulators can be constant flow or diluter-demand.
o Constant flow. The constant flow regulator provides the same output pressure or flow regardless of altitude. The
regulator is therefore optimized for a specific altitude. At altitudes lower than the designed optimum altitude, it
will provide more oxygen than is actually required. This type of regulator is most often found in non-
pressurized aircraft and on portable oxygen systems. A single constant flow regulator is able to control the
oxygen flow to all users.
o Diluter-demand. When installed, diluter-demand regulators will be located at each crew position. Depending
upon user selection, the diluter-demand regulator can provide 100% oxygen, 100% oxygen under positive
pressure or a mixture of oxygen "diluted" with cabin air on a specific, altitude based schedule. As an example,
at 8000', the regulator might send 100% ambient air to the mask whereas at 41000', it would provide 100%
oxygen. The regulator also works on "demand". That is, the oxygen or air-oxygen mixture only flows into the
mask during inhalation. Note that the regulator might be a stand alone unit or it could be incorporated into the
mask itself.
 An oxygen mask is provided at each flight deck station.
o The mask could be of the "full face" variety incorporating smoke goggles or a "mouth and nose" type mask with
smoke goggles available separately.
o The masks at the pilot stations will incorporate microphones to allow internal and external communications.
o Masks are fitted to the face utilizing various suspension harnesses. For aircraft which routinely fly above
25,000', masks are generally of the "quick-donning" variety. These are designed to allow them to be put on in 5
seconds or less using only one hand and often utilize oxygen system pressure to activate an inflatable harness
for quick donning.
o For diluter-demand systems, selectors for normal, 100% and positive pressure maybe incorporated into the
mask itself. If not, they will be found on the associated regulator. Diluter-demand oxygen masks are stowed
with the selector in the 100% oxygen position and should be reselected to the normal (or diluting) position when
mask utilization is required for other than a smoke or fume event.
Passenger Compartment

 In non-pressurised aircraft which routinely fly above 10,000', passenger oxygen is typically provided by either a
fixed or a portable system.
o Fixed systems draw their oxygen supply from a pressurised cylinder of gaseous oxygen. This can be a dedicated
cylinder or it might be the same cylinder that is used to supply the flight deck occupants. An oxygen manifold
runs from the cylinder into the passenger compartment via a single regulator. Attachment ports allow passenger
oxygen masks to be connected to the manifold. A shutoff valve capable of isolating the passenger compartment
is normally incorporated.
o Portable systems consist of a storage tank, a regulator and one or more passenger masks. These will be
distributed to the passengers when required.
 Pressurised aircraft which have a certified maximum altitude of 25,000' or less do not require passenger oxygen
systems subject to the aircraft being able to descend to 13,000' or below within 4 minutes of loss of
pressurisation. If the aircraft is not capable of achieving the descent profile or the route structure does not allow
the descent due to terrain, an oxygen system must be fitted in the aircraft as per the provisions which apply to
aircraft which are certified to fly at higher altitudes (above 25,000').
 For pressurised aircraft which are certified to operate above 25,000', emergency oxygen equipment must be
available. Some aircraft utilize cylinders of pressurised oxygen to meet this requirement but most types are
fitted with chemical oxygen generators.
o The emergency oxygen supply must last a minimum of 10 minutes.
o Provisions must be provided in the system to automatically deploy the emergency oxygen masks when the cabin
altitude exceeds a pre-determined level, normally 14,000'.
o Sufficient masks must be provided for at least 10% more passengers than there are seats in the passenger
compartment. This excess requirement provides masks for small children who may not be assigned a seat and
for anyone (such as Flight Attendants) who might not be in their assigned seat at the moment emergency
oxygen is required.
 The most typical passenger oxygen masks consist of a soft, yellow silicone cup fitted with elastic bands for
securing the mask to the face. The bands are adjustable to accomodate passengers of different sizes. The mask
may also have a clear concentrator or re-breather bag. Depending upon the cabin altitude, the concentrator bag
may or may not inflate. Airlines make a point during their safety presentation of pointing out that the bag may
not inflate as, in the past, lack of bag inflation has lead some passengers to believe that their mask was not
working and to remove it resulting in hypoxia. Due to a potentially limited time of useful consciousness, it is
critical that masks be put on immediately and kept on until advised by the crew that it is safe to remove them.
Passengers should always don their own mask prior to assisting others (such as children) with their mask.

Noise abatement departure climb guidance


The first procedure NADP 1 is intended to provide noise reduction for noise-sensitive areas in close
proximity to the departure end of the runway. The second procedure NADP 2 is intended to provide
noise reduction for noise-sensitive areas more distant from the runway end.
The two procedures differ in that the acceleration segment for flap/slat retraction is either initiated
prior to reaching the maximum prescribed height or at the maximum prescribed height. To ensure
optimum acceleration performance, thrust reduction may be initiated at an intermediate flap setting.
These methods are present ICAO methods but they do not constitute the only methods to be
followed. These procedures are provided as examples because the noise reduction obtained greatly
depends on the type of aircraft, engine type, thrust required and the height at which thrust is reduced.
Commercial airlines are encouraged to develop own procedures that maximize noise reduction from
their aircraft.
Noise abatement climb NADP1
This procedure involves:
 Power or thrust reduction at or above the prescribed minimum altitude 240m or 800ft above
aerodrome elevation.
 Initial climbing speed is not less than V2 + 10kt or V2 + 20 km/h and below V2 + 20kt or V2 + 40 km/h
 Delay of flats/slats retraction until the prescribed maximum altitude 900m or 3000ft is attained
 At the prescribed maximum altitude 900m or 3000ft, the aircraft is accelerated and the flaps/slats are
retracted on schedule while maintaining a positive rate of climb, to complete the transition to normal
en-route climb speed.

Noise abatement climb NADP2


This procedure involves:
 Power or thrust reduction at or above the prescribed minimum altitude 240m or 800ft above
aerodrome elevation.
 Flats/slats retraction at or above the prescribed minimum altitude 240m or 800ft above aerodrome
elevation but before the prescribed maximum altitude 900m or 3000ft
 Flaps/slats are retracted on schedule while maintaining a positive rate of climb
 Intermediate flap retraction, if required for performance may be accomplished
 Aircraft body angle of pitch is decreased, aircraft is accelerated towards Vzf (Minimum Safe
manoeuvring Velocity with Zero Flaps)
 Initial climbing speed is not less than V2 + 10kt or V2 + 20 km/h
 Power or thrust reduction is initiated at a point along the acceleration segment that ensure
satisfactory acceleration performance.
 At the prescribed maximum altitude 900m or 3000ft, the aircraft is accelerated to complete the
transition to normal en-route climb speed.
Aircraft deicing fluid
In ground deicing of aircraft, aircraft deicing fluid (ADF), aircraft deicer and anti-icer fluid (ADAF)
or aircraft anti-icing fluid (AAF) are commonly used for both commercial and general aviation.
Environmental concerns include increased salinity of groundwater where de-icing fluids are
discharged into soil, and toxicity to humans and other mammals.
Fluids used

Deicing fluids come in a variety of types, and are typically composed of ethylene glycol (EG)
or propylene glycol (PG), along with other ingredients such as thickening agents, surfactants (wetting
agents), corrosion inhibitors, colors, and UV-sensitive dye. Propylene glycol-based fluid is more
common because it is less toxic than ethylene glycol.
SAE International (formerly known as the Society of Automotive Engineers) publishes standards
(SAE AMS 1428 and AMS 1424) for four different types of aviation deicing fluids:

1. Type I fluids have a low viscosity, and are considered "unthickened". They provide only short
term protection because they quickly flow off surfaces after use. They are typically sprayed hot
(130–180 °F, 55–80 °C) at high pressure to remove snow, ice, and frost. Usually they are
dyed orange to aid in identification and application.
2. Type II fluids are pseudoplastic, which means they contain a polymeric thickening agent to
prevent their immediate flow off aircraft surfaces. Type II prevents snow, ice or frost
contamination from adhering to the aircraft from the apron to takeoff. Typically the fluid film will
remain in place until the aircraft attains 100 knots (190 km/h) or so, at which point the viscosity
breaks down due to shear stress. The high speeds required for viscosity breakdown means
that this type of fluid is useful only for larger aircraft. The use of Type II fluids is diminishing in
favour of Type IV. Type II fluids are generally clear in color.
3. Type III fluids can be thought of as a compromise between Type I and Type II fluids. They are
intended for use on slower aircraft, with a rotation speed of less than 100 knots. Type III fluids
are generally bright yellow in color.
4. Type IV has the same purpose and meets the same AMS standards as Type II fluids, but they
provide a longer holdover time. They are typically dyed green to aid in the application of a
consistent layer of fluid.
The International Organization for Standardization publishes equivalent standards (ISO 11075 and
ISO 11078), defining the same four types.
Deicing fluids containing thickeners (Types II, III, and IV) are also known as anti-icing fluids, because
they are used primarily to prevent icing from re-occurring after an initial deicing with a Type I fluid.
TKS fluid is similar to Type I fluid and is used by in-flight TKS ice protection systems. It can also be
used for ground-based deicing. It conforms to different standards than Type I fluid, namely DTD
406B, AL-5, and NATO S-745.

Anti-icing systems

Anti-icing systems typically use heat to stop ice from forming. These heat-related systems work by
causing the moisture to evaporate into the atmosphere as soon as it touches the heated surface of
your plane.

Bleed air systems – If your aircraft is turbine-powered, then bleed air systems are most likely to be
the anti-icing method used. The term “bleed air” comes from the fact that air is bled off from your
aircraft’s hot engines, and then this hot air is fed through to all the critical surfaces of the plane.

The bleed air system works well for larger aircraft, but because it can affect engine temperatures and
reduce climbing ability, the system is not usually used in most small aircraft. If your plane is not
turbine-powered but piston-powered, then a bleed air system is not used, but electrical power is most
likely used to supply the heat instead.

Thermal anti-icing systems – Electro-thermal or electrical heating systems operate in a similar way
to a stove element. Heating coils embedded in the plane’s structure generate heat using a controlled
electrical current.

In these systems, electricity is used to heat various components on the aircraft to prevent the
formation of ice. These components can include:
 pitot tubes
 static air ports
 TAT and AOA probes
 ice detectors
 engine P2/T2 sensors
 water lines
 wastewater drains
 turboprop inlet cowls

Fuel - Flight Planning Definitions


As in many facets of aviation, Fuel Planning has a list of specific terms and definitions of its own. The following list
identifies the most critical of these terms. Different terms or names for the same concept are often used
interchangeably by different regulatory authorities or flight planning organizations. The most common of these
variants preface the definitions that follow.

Additional Fuel
Additional fuel is fuel which is added to comply with a specific regulatory or company requirement. Examples
include ETOPS fuel, fuel required for a remote or island destination where no alternate is available and fuel required
to satisfy an Minimum Equipment List (MEL) or Configuration Deviation List performance penalty.

Alternate Fuel
Alternate fuel is the amount of fuel required from the missed approach point at the destination aerodrome until
landing at the alternate aerodrome. It takes into account the required fuel for:
 Missed approach at the destination airport
 Climb to enroute altitude, cruise and descent at alternate aerodrome
 Approach at alternate
 Landing at the alternate aerodrome
When two alternates are required by the Authority, alternate fuel must be sufficient to proceed to the alternate which
requires the greater amount of fuel.

Ballast Fuel
Ballast fuel is sometimes carried to maintain the aircraft centre of gravity within limits. In certain aeroplanes, a zero
fuel weight above a defined threshold requires that a minimum amount of fuel be carried in the wings through all
phases of flight to prevent excessive wing bending. In both cases, this fuel is considered ballast and, under anything
other than emergency circumstances, is not to be burned during the flight.

Block Fuel / Ramp Fuel / Total Fuel On Board


Block fuel is the total fuel required for the flight and is the sum of the Taxi fuel, the Trip fuel, the Contingency fuel,
the Alternate fuel, the Final Reserve fuel, the Additional fuel and any Extra fuel carried.

Contingency Fuel / Route Reserve


Contingency fuel is carried to account for additional enroute fuel consumption caused by wind, routing changes
or ATM: ATM/CNS restrictions. According to ICAO Annex 6, the recommended minimum contingency fuel is the
greater of 5% of the trip fuel or 5 minutes holding consumption at 1500' above destination airfield elevation
computed based on calculated arrival weight. However, some regulators have eliminated the minimum time
requirement and some have increased the recommended time interval in their National Regulations. As well, some
regulators allow contingency fuel reduction to 3% of trip fuel, or to specific time increments, with use of enroute
alternates and conditional upon demonstrated performance criteria from the Operator. At least one authority allows,
under very specific circumstances, for contingency fuel to be reduced to 0. In all cases, an Operator can direct that its
crews carry contingency fuel in excess of that required by their National Aviation Authority (NAA).

Extra Fuel
Fuel added at the discretion of the Captain and/or the dispatcher

Final Reserve Fuel / Fixed Reserve Fuel / Holding Fuel


Final reserve fuel is the minimum fuel required to fly for 30 minutes at 1,500 feet above the alternate aerodrome or,
if an alternate is not required, at the destination aerodrome at holding speed in ISA conditions. Some Regulating
Authorities require sufficient fuel to hold for 45 minutes.

Minimum Brake Release Fuel


Minimum brake release fuel is that quantity of fuel which, at the commencement of the takeoff roll, complies with
all regulatory requirements for the flight in question. This is the minimum legal fuel required for departure.

Reserve Fuel / Minimum Diversion Fuel


Reserve fuel is the sum of Alternate fuel plus Final Reserve fuel.

Taxi Fuel
Taxi fuel is the fuel used prior to takeoff and will normally include pre-start APU consumption, engine start and taxi
fuel. Taxi fuel is usually a fixed quantity for an average taxi duration. However, local conditions at the departure
aerodrome such as average taxi time, normal ground delays and any anticipated deicing delays should be taken into
consideration and the taxi fuel adjusted accordingly.

Trip Fuel / Burn / Fuel to Destination


The Trip fuel is the required fuel quantity from brake release on takeoff at the departure aerodrome to the landing
touchdown at the destination aerodrome. This quantity includes the fuel required for:
 Takeoff
 Climb to cruise level
 Flight in level cruise including any planned step climb or step descent
 Flight from the beginning of descent to the beginning of approach,
 Approach
 Landing at the destination
Trip fuel must be adjusted to account for any additional fuel that would be required for known ATS restrictions that
would result in delayed climb to or early descent from planned cruising altitude.
ICAO Policy for FUEL
As per ICAO Annex 6, Part I, section 4.3.6 "Fuel Requirements," airplanes should calculate their required
fuel quantity as follows (summary; see below for actual ICAO text):

 Taxi fuel
 Trip fuel (to reach intended destination)
 Contingency fuel (higher of 5% of "trip fuel" or 5 minutes of holding flight)
 Destination alternate fuel (to fly a missed and reach an alternate)
 Final reserve fuel (45 minutes of holding flight for reciprocating engines, 30 minutes for jets)
 Additional fuel (if needed to guarantee ability to reach an alternate with an engine failure or at lower altitude
due to a pressurization loss)
 Discretionary fuel (if the pilot in command wants it)

General Aviation

For general aviation, ICAO Annex 6 Part II, section 2.2.3.6 "Fuel and oil supply" requires:

 For IFR, enough fuel to reach destination, then alternate (if required), plus 45 minutes
 For day VFR, enough fuel to reach destination plus 30 minutes
 For night VFR, enough fuel to reach destination plus 45 minutes

ICAO Annex 6 Part I

From the ninth edition:


4.3.6.3 The pre-flight calculation of usable fuel required shall include:
a) taxi fuel, which shall be the amount of fuel expected to be consumed before take-off;
b) trip fuel, which shall be the amount of fuel required to enable the aeroplane to fly from take-off, or the point of in-
flight re-planning, until landing at the destination aerodrome taking into account the operating conditions of 4.3.6.2
b);
c) contingency fuel, which shall be the amount of fuel required to compensate for unforeseen factors. It shall be five
per cent of the planned trip fuel or of the fuel required from the point of in-flight re-planning based on the
consumption rate used to plan the trip fuel but, in any case, shall not be lower than the amount required to fly for five
minutes at holding speed at 450 m (1 500 ft) above the destination aerodrome in standard conditions;
Note.— Unforeseen factors are those which could have an influence on the fuel consumption to the destination
aerodrome, such as deviations of an individual aeroplane from the expected fuel consumption data, deviations from
forecast meteorological conditions, extended taxi times before take-off, and deviations from planned routings and/or
cruising levels.
d) destination alternate fuel, which shall be:
1. where a destination alternate aerodrome is required, the amount of fuel required to enable the aeroplane to:
i) perform a missed approach at the destination aerodrome;
ii) climb to the expected cruising altitude; iii) fly the expected routing;
iv) descend to the point where the expected approach is initiated; and
v) conduct the approach and landing at the destination alternate aerodrome; or
2. where two destination alternate aerodromes are required, the amount of fuel, as calculated in 4.3.6.3 d) 1),
required to enable the aeroplane to proceed to the destination alternate aerodrome which requires the greater
amount of alternate fuel; or
3. where a flight is operated without a destination alternate aerodrome, the amount of fuel required to enable the
aeroplane to fly for 15 minutes at holding speed at 450 m (1 500 ft) above destination aerodrome elevation in
standard conditions; or
4. where the aerodrome of intended landing is an isolated aerodrome:
i) for a reciprocating engine aeroplane, the amount of fuel required to fly for 45 minutes plus 15 per cent of the
flight time planned to be spent at cruising level, including final reserve fuel, or two hours, whichever is less; or
ii) for a turbine-engined aeroplane, the amount of fuel required to fly for two hours at normal cruise
consumption above the destination aerodrome, including final reserve fuel;
e) final reserve fuel, which shall be the amount of fuel calculated using the estimated mass on arrival at the
destination alternate aerodrome, or the destination aerodrome when no destination alternate aerodrome is required:
1. for a reciprocating engine aeroplane, the amount of fuel required to fly for 45 minutes, under speed and altitude
conditions specified by the State of the Operator; or
2. for a turbine-engined aeroplane, the amount of fuel required to fly for 30 minutes at holding speed at 450 m (1
500 ft) above aerodrome elevation in standard conditions;
f) additional fuel, which shall be the supplementary amount of fuel required if the minimum fuel calculated in
accordance with 4.3.6.3 b), c), d) and e) is not sufficient to:
1. allow the aeroplane to descend as necessary and proceed to an alternate aerodrome in the event of engine failure
or loss of pressurization, whichever requires the greater amount of fuel based on the assumption that such a
failure occurs at the most critical point along the route;
i) fly for 15 minutes at holding speed at 450 m (1 500 ft) above aerodrome elevation in standard conditions; and
ii) make an approach and landing;
2. allow an aeroplane engaged in EDTO to comply with the EDTO critical fuel scenario as established by the State
of the Operator;
3. meet additional requirements not covered above;
Note 1.— Fuel planning for a failure that occurs at the most critical point along a route (4.3.6.3 f) 1)) may place the
aeroplane in a fuel emergency situation based on 4.3.7.2.
Note 2.— Guidance on EDTO critical fuel scenarios is contained in Attachment D;
g) discretionary fuel, which shall be the extra amount of fuel to be carried at the discretion of the pilot-in-command.

ICAO Annex 6 Part II


2.2.3.6 Fuel and oil supply
A flight shall not be commenced unless, taking into account both the meteorological conditions and any delays that
are expected in flight, the aeroplane carries sufficient fuel and oil to ensure that it can safely complete the flight. The
amount of fuel to be carried must permit:

a) when the flight is conducted in accordance with the instrument flight rules and a destination alternate aerodrome is
not required in accordance with 2.2.3.5, flight to the aerodrome of intended landing, and after that, for at least 45
minutes at normal cruising altitude; or

b) when the flight is conducted in accordance with the instrument flight rules and a destination alternate aerodrome
is required, flight from the aerodrome of intended landing to an alternate aerodrome, and after that, for at least 45
minutes at normal cruising altitude; or

c) when the flight is conducted in accordance with the visual flight rules by day, flight to the aerodrome of intended
landing, and after that, for at least 30 minutes at normal cruising altitude; or

d) when the flight is conducted in accordance with the visual flight rules by night, flight to the aerodrome of
intended landing and thereafter for at least 45 minutes at normal cruising altitude.
Runway Condition Definitions
a. Contaminated runway. Contaminated runway. A runway is contaminated when more
than 25 per cent of the runway surface area (whether in isolated areas or not) within
the required length and width being used is covered by:
o water, or slush more than 3 mm (0.125 in) deep;
o loose snow more than 20 mm (0.75 in) deep; or
o compacted snow or ice, including wet ice.
b. Dry runway. A dry runway is one which is clear of contaminants and visible moisture
within the required length and the width being used.
c. Wet runway. A runway that is neither dry nor contaminated.

Note 1.— In certain situations, it may be appropriate to consider the runway


contaminated even when it does not meet the above definition. For example, if less
than 25 per cent of the runway surface area is covered with water, slush, snow or ice,
but it is located where rotation or lift-off will occur, or during the high speed part of
the take-off roll, the effect will be far more significant than if it were encountered early
in take-off while at low speed. In this situation, the runway should be considered to be
contaminated.

Note 2.— Similarly, a runway that is dry in the area where braking would occur during
a high speed rejected take-off, but damp or wet (without measurable water depth) in
the area where acceleration would occur, may be considered to be dry for computing
take-off performance. For example, if the first 25 per cent of the runway was damp, but
the remaining runway length was dry, the runway would be wet using the definitions
above. However, since a wet runway does not affect acceleration, and the braking
portion of a rejected take-off would take place on a dry surface, it would be
appropriate to use dry runway take-off performance.

Source: ICAO Annex 6, Part 1, Att B, Page B-2

The ICAO definition in Annex 6, Part I, applies only to commercial aviation. The same
definition does not appear in Part II for general aviation.

European Union — Commercial Operators

 "Contaminated runway". A runway is considered to be contaminated when more than


25 % of the runway surface area (whether in isolated areas or not) within the required
length and width being used is covered by the following:
i. surface water more than 3 mm (0,125 in) deep, or by slush, or loose snow,
equivalent to more than 3 mm (0,125 in) of water;
ii. snow which has been compressed into a solid mass which resists further
compression and will hold together or break into lumps if picked up (compacted
snow); or
iii. ice, including wet ice.
 "Damp runway". A runway is considered damp when the surface is not dry, but when
the moisture on it does not give it a shiny appearance.

--------------------------------------------------------------------------------------------------------------------------------------------------

There is a requirement to report the presence of water within the central half of the
width of a runway and to make an assessment of water depth, where possible. To be
able to report with some accuracy on the conditions of the runway, the following terms
and associated descriptions should be used:

 Damp — the surface shows a change of colour due to moisture.


 Wet — the surface is soaked but there is no standing water.
 Water patches — significant patches of standing water are visible.
 Flooded — extensive standing water is visible

Source: ICAO Doc 9137, Part 2, ¶3.3


Aircraft Fire Extinguishing Systems
Onboard systems designed to extinguish fires which occur either in the air or on the ground.
For information on detecting and fighting fires in the cabin, see also the separate article: "Passenger Cabin Fire"

Description
Four types of fire extinguishing installations are found on commercial transport aircraft.
 Portable extinguishers installed at specified locations in both the main cabin and the flight deck
 Hold fire extinguishing systems (with automatic detection)
 Engine fire bottle extinguishing systems (with automatic detection)
 Toilet waste bin bottle extinguishing systems

Portable Extinguishers

General
Fires on board aircraft which occur within the aircraft cabin or flight deck - or are potentially directly accessible
from them - arise in one of three ways:
 Fires that involve energized electrical equipment - in aircraft cabins typically IFE (in flight entertainment) systems in
the passenger cabin, electrical equipment in the galley or avionics equipment in the flight deck or under floor
avionics bay, or personal electronic devices (PEDs) carried by passengers.
 Fires in ordinary combustibles such as cloth, paper, rubber, and many plastics - in aircraft cabins typically in
furnishings
 Fires in flammable liquids, oils, greases, tars, oil-base paints, lacquers, and flammable gases - in aircraft cabins
typically galley oven fires
Portable extinguishers present a special challenge since they must be capable of extinguishing a range of fire types -
solid materials such as cabin fixtures and furnishings, flammable liquids and electrical fires.
Halon 1211 extinguishers have entirely replaced the previous combination of two different types of portable
extinguisher - carbon dioxide and water gycol - on new-build aircraft and no other single extinguisher-type has yet
been identified as a satisfactory alternative to it.

NOTE: Crews must be aware that the toxicity of the Halon gases, especially the combination which makes up
Halon 1211, is such that use in confined spaces requires care to minimize any inhalation of the discharged
gases. Where a portable Halon extinguisher is used by cabin crew, it is usually recommended to consider
donning a smoke hood before discharge to eliminate this risk, but for flight crew use on the flight deck, this
will not be an option and risk awareness is the only defence.

Where the dual fit of extinguisher is encountered on older aircraft, it is essential that Water Glycol extinguishers are
used on solid material fires and Carbon Dioxide extinguishers on liquid or electrical equipment fires.
The minimum dispatch requirement for aircraft portable fire extinguishers is determined by the capacity of the
aircraft cabin and is specified in the Aircraft MEL.
Regulatory Requirement
Relevant authorities (e.g. the European Union Aviation Safety Agency (EASA) in the EU, FAA in the US) specify
the requirements for hand held fire extinguishers in terms of:
 Minimum number of hand held fire extinguishers to be carried on board. This depends on the size of the aircraft, the
number of passenger seats, the number and type of cargo compartments, etc.
 Hand held fire extinguisher distribution within the aircraft (e.g. number of extinguishers in the cockpit, cabin, cargo
compartments, etc.).
 Mounting and marking (e.g. there should be a special sign in case the fire extinguisher is not clearly visible).
 The need for regulatory approval and acceptable means of compliance.
 Restrictions regarding the extinguishing agents used (e.g. cut-off and end dates for Halon-based extinguishers).
 Extinguishing agent quantity requirements (e.g. minimum amount of agent per extinguisher).
 Health considerations (e.g. toxic gas hazard minimization).

Hold Fire Extinguishing Systems


Hold fire extinguishing systems are usually activated as a flight crew response to abnormal heat detection in an
aircraft hold, and usually operate in a dual function. Part of the available fire suppression capability is deployed in an
‘instant’, or ‘knock-down’, discharge of extinguishing agent and the remainder is deployed more gradually over a
longer period of up to an hour, to assist in preventing re-ignition or at least providing partial fire suppression, to
provide more time to get an aircraft with a continuing hold fire warning back on the ground. Various alternatives to
Halon 1301 have been examined including water misting, inert gas and dry powder, either alone or in combination.
The FAA has developed minimum performance standards for these systems and it has been demonstrated that
although water misting alone is unable to pass the exploding aerosol can fire test, a combination of water misting and
inert gas (nitrogen) discharge may be more effective. However, for such a solution to be viable, a means of on-board
nitrogen generation will be needed.

Engine Fire Bottles


Fire bottles in engine compartments are usually electrically operated after manual selection by the flight crew based
upon automatic fire detection. In the airborne case, APU fire bottles are similarly activated but it is usual for
automatic APU fire detection during ground operation to trigger automatic shutdown and fire extinguisher
activation. Until recently, the most common extinguishing agent was Halon 1301 for all Engines/APUs fitted to civil
transport aircraft. However, Halon 1301 is no longer manufactured and has been banned (for new systems) since
1994; often they are now replaced by HFCs (Hydrofluorocompounds).

Toilet Waste Bins


Toilet waste bin fire extinguishers are activated automatically if heat detectors in the vicinity are activated. Toilet
smoke detector activation does not trigger waste bin fire extinguishers. Alternative extinguishing agents to Halon
1301 have been approved for use in fixed toilet waste bin systems and have also been, uniquely in terms of the
search for Halon alternatives, shown to be more effective than Halon 1301 units whilst being the same size. Since
only a documentation change is required to fit these alternative extinguishers, they have been used for retrofit as well
as in new-build aircraft.
Personal Electronic Devices
Equipment has been introduced designed to deal specifically with lithium battery fires in PEDs; lithium ion batteries
(Li-ion) are used to power PEDs such as cellular phones, portable tablets, EFBs and digital cameras; Li-ion batteries
are rechargeable. Non-rechargeable lithium batteries (Li-metal) are similar to Li-ion, but use a different electrode
material – metallic lithium.
All lithium batteries present a potential fire hazard. These batteries are carried on aeroplanes as cargo, within
passenger baggage, and by passengers directly. Like some other batteries lithium batteries are capable of delivering
sufficient energy to start an in-flight fire. Lithium batteries present a greater risk of an in-flight fire than some other
battery types because they are also unable to contain their own energy in the event of a catastrophic failure.
Once extinguished, a lithium battery fire – or a fire in a PED powered by lithium batteries – requires containment
and continued cooling. Halon 1211 or water fire extinguishers are effective at extinguishing the fire and preventing
its spread to additional flammable materials. After extinguishing the fire, dousing the electronic device with water or
other non-alcoholic liquids cools the device and prevents additional battery cells from reaching thermal runaway.
Containment devises are now available and where these are equipped, crews should receive specific training in how
to use them to greatest effect.
These issues are discussed in some detail in the RAeS document "Smoke, fire and fumes in transport aircraft, past
history, current risks and recommended mitigations - Part 1:References". See also the separate articles: "Aircraft Fire
Risk from Battery-powered Items Carried on Aircraft" and 'Personal Electronic Device Fire - Cabin Crew
Checklist".

The solution is to use aviation-approved fire extinguishers. The two most common types of
aviation extinguishers are carbon dioxide and halon (specifically Halon 1211) extinguishers.
Carbon dioxide extinguishers are suitable for airport and hangar use, but never in small
spaces or in any cockpit. Carbon dioxide is poisonous at a concentration of only 4 percent
and lethal at just 8 percent.
Reduced Vertical Separation Minima (RVSM)
A program was initiated by ICAO in 1982 involving worldwide studies to assess the feasibility of a reduction of the
Vertical Separation Minima (VSM) above FL290 from 2,000 feet to 1,000 feet.
The principal benefits which the implementation of the reduced VSM were expected to provide were:
 A theoretical doubling of the airspace capacity, between FL290 and FL410; and
 The opportunity for aircraft to operate at closer to the optimum flight levels with the resulting fuel economies.
The program relies on the carriage and serviceability of specified aircraft equipment and the existence of appropriate
operating procedures to ensure that the risk of loss of separation is no greater than it would be outside RVSM
airspace.

Implementation
Between 1997 and 2005 RVSM was implemented in all of Europe, North Africa, Southeast Asia, North America,
South America, and over the North Atlantic, South Atlantic, and Pacific Oceans.

Approval for RVSM Operations


State airworthiness authorities are responsible for verifying that an aircraft is technically capable of meeting and
maintaining the stringent altimetry system performance requirements. Crews must be trained in appropriate
procedures in RVSM airspace. Providing all these requirements are met, an authority will issue an RVSM
Operational Approval. Operators indicate RVSM approval by filing a W in field 10 of the ICAO model flight plan. It
is a violation of ICAO European regional supplementary procedures for a non-approved aircraft to file a W.
The Regional Monitoring Agency (RMA) is responsible for verifying the approval status of aircraft operating in
RVSM airspace and reporting violations to the appropriate state authority.
An important element of the certification process is the confirmation of the aircraft height keeping performance
across the entire operational flight envelope. The flight envelope covers all combinations of speed, altitude and
weight/atmospheric pressure ratio that the aircraft would expect to operate across in RVSM airspace. The assessment
of the aircraft performance across the flight envelope, together with the service bulletin, continuing airworthiness
instructions and the amendment to the aircraft flight manual are collectively known as the RVSM approval data
package. Confirmation of the RVSM approval data package is a fundamental requirement before any RVSM
operational approval is issued.

Regulatory Requirements
 An operator shall not operate an aeroplane in defined portions of airspace where, based on regional air navigation
agreement, a vertical separation minimum 300 m (1000ft) applies unless approved to do so by the Authority (RVSM
Approval). EASA IR-OPS SPA.RVSM.100 and SPA.RVSM.110, EU-OPS 1.241 See also EU-OPS 1.872.
 Prior to granting the RVSM approval... the State shall be satisfied that:
a) the vertical navigation performance capability of the aeroplane satisfies the (laid down requirements);
b) the operator has instituted appropriate procedures in respect of continued airworthiness (maintenance and repair)
practices and programmes; and
c) the operator has instituted appropriate flight crew procedures for operation in RVSM airspace.
Note: An RVSM approval is valid globally on the understanding that any operating procedures specific to a given
region will be stated in the Operations Manual or appropriate crew guidance. (ICAO Annex 6 Part I Chapter 7, Para
7.2.5.)
 An operator shall ensure that aeroplanes operated in RVSM airspace are equipped with:
1. Two independent altitude measurement systems;
2. An altitude alerting system;
3. An automatic altitude control system; and
4. A secondary surveillance radar (SSR) transponder with altitude reporting system that can be connected to the altitude
measurement system in use for altitude keeping. (IR-OPS SPA.RVSM.110, EU-OPS 1.872)

Separation standards within RVSM Airspace


Within RVSMairspace (between FL290 and FL410 inclusive) the vertical separation minimum is:
 1000ft (300m) between RVSM-approved aircraft, and
 2000ft (600m) between non-RVSM approved state aircraft and any other aircraft operating within RVSM airspace.
 2000ft (600m) between non-RVSM aircraft operating as general air traffic (GAT) and any other aircraft within
RVSM airspace.
There is no exemption for state aircraft to operate as GAT within RVSM airspace with a 1000 ft vertical separation
minimum without an RVSM approval. The absence of such approval does require a separation of 2000 ft to be
observed. State aircraft which are exempted from having to meet the RVSM Minimum Aircraft System Performance
Specification (MASPS) in Field 18 of the ICAO FPL, shall request special handling by filling “STS/NONRVSM”.
Formation flights are to be considered non-RVSM compliant irrespective of the RVSM status of the individual
aircraft within the formation and are not permitted within RVSM airspace with a 1000 ft vertical separation
minimum.

Contingency procedures when unable to maintain RVSM


 The pilots shall notify ATC of any equipment failure, weather hazards such as severe turbulence etc., which may
affect the ability to maintain the cleared level or the RVSM requirements. When an aircraft operating in RVSM
Airspace encounters severe turbulence due to weather or wake vortex which the pilot believes will impact the
aircraft’s capability to maintain its cleared flight level, the pilot shall inform ATC. ATC is required to establish
either an appropriate horizontal separation minimum, or an increased vertical separation minimum of 2000ft;
 Where a meteorological forecast is predicting severe turbulence within the RVSM Airspace, ATC shall determine
whether RVSM should be suspended, and, if so, the period of time, and specific flight level(s) and/or area.
 When notified by ATC of an assigned altitude deviation of more than 300ft (90 m), the pilot shall take action to
return to the cleared level as quickly as possible.
 In the event of a pilot advising that the aircraft is no longer capable of RVSM operations, it is particularly important
that the first ATS unit made aware of the failure performs the necessary co-ordination with subsequent ATS units.
Barometric Altimeter Errors and Setting Procedures

1. General
1. Aircraft altimeters are subject to the following errors and weather factors:
1. Instrument error.
2. Position error from aircraft static pressure systems.
3. Nonstandard atmospheric pressure.
4. Nonstandard temperatures.
2. The standard altimeter 29.92 inches Mercury (“Hg.) setting at the higher altitudes eliminates station
barometer errors, some altimeter instrument errors, and errors caused by altimeter settings derived
from different geographical sources.
2. Barometric Pressure Altimeter Errors
1. High Barometric Pressure: Cold, dry air masses may produce barometric pressures in excess of 31.00
“Hg. Many aircraft altimeters cannot be adjusted above 31.00 “Hg. When an aircraft's altimeter cannot
be set to a pressure setting above 31.00 “Hg, the aircraft's true altitude will be higher than the indicated
altitude on the barometric altimeter.
2. Low Barometric Pressure: An abnormal low-pressure condition exists when the barometric pressure is
less than 28.00 “Hg. Flight operations are not recommended when an aircraft's altimeter is unable to be
set below 28.00 “Hg. In this situation, the aircraft's true altitude is lower than the indicated altitude. This
situation may be exacerbated when operating in extremely cold temperatures, which may result in the
aircraft's true altitude being significantly lower than the indicated altitude.

NOTE-

EXTREME CAUTION SHOULD BE EXERCISED WHEN FLYING IN PROXIMITY TO OBSTRUCTIONS OR


TERRAIN IN LOW PRESSURES AND/OR LOW TEMPERATURES.

3. Altimeter Setting Procedures


1. Manufacturing and installation specifications, along with 14 CFR Part 43, Appendix E requirement for
periodic tests and inspections, helps reduce mechanical, elastic, temperature, and installation errors.
(See Instrument Flying Handbook.) Scale error may be observed while performing a ground altimeter
check using the following procedure:
1. Set the current reported airfield altimeter setting on the altimeter setting scale.
2. Read the altitude on the altimeter. The altitude should read the known field elevation if you are located
on the same reference level used to establish the altimeter setting.
3. If the difference from the known field elevation and the altitude read from the altimeter is plus or minus
75 feet or greater, the accuracy of the altimeter is questionable and the problem should be referred to
an appropriately rated repair station for evaluation and possible correction.
2. It is important to set the current altimeter settings for the area of operation when flying at an enroute
altitude that does not require a standard altimeter setting of 29.92 “Hg. If the altimeter is not set to the
current altimeter setting when flying from an area of high pressure into an area of low pressure, the
aircraft will be closer to the surface than the altimeter indicates. An inch Hg error in the altimeter
setting equals 1,000 feet of altitude. For example, setting 29.90 “Hg. instead of 30.90 “Hg. To quote an
old saying: “GOING FROM A HIGH TO A LOW, LOOK OUT BELOW.”
3. The aircraft cruising altitude or flight level is maintained by referencing the barometric altimeter.
Procedures for setting altimeters during high and low barometric pressure events must be set using the
following procedures:
1. Below 18,000 feet mean sea level (MSL).
1. Barometric pressure is 31.00 “Hg. or less.
1. Set the altimeter to a current reported altimeter setting from a station along the route and within 100
NM of the aircraft, or;
2. If there is no station within this area, use the current reported altimeter setting of the nearest
available station, or;

NOTE-

Air traffic controllers will furnish this information at least once when en route or on an instrument
flight plan within their controlled airspace:

3. If the aircraft is not equipped with a radio, set the altimeter to the elevation of the departure airport
or use an available appropriate altimeter setting prior to departure.
2. When the barometric pressure exceeds 31.00 “Hg., a NOTAM will be published to define the affected
geographic area. The NOTAM will also institute the following procedures:
1. All aircraft: All aircraft will set 31.00 “Hg. for en route operations below 18,000 feet MSL. Maintain this
setting until out of the affected area or until reaching the beginning of the final approach segment on
an instrument approach. Set the current altimeter setting (above 31.00 “Hg.) approaching the final
segment, if possible. If no current altimeter setting is available, or if a setting above 31.00 “Hg. cannot
be made on the aircraft's altimeter, leave 31.00 “Hg. set in the altimeter and continue the approach.
2. Set 31.00 “Hg. in the altimeter prior to reaching the lowest of any mandatory/ crossing altitudes or
1,500 feet above ground level (AGL) when on a departure or missed approach.

NOTE-

Air traffic control will issue actual altimeter settings and advise pilots to set 31.00 “Hg. in their
altimeters for en route operations below 18,000 feet MSL in affected areas.

3. No additional restrictions apply for aircraft operating into an airport that are able to set and measure
altimeter settings above 31.00 “Hg.
4. Flight operations are restricted to VFR weather conditions to and from an airport that is unable to
accurately measure barometric pressures above 31.00 “Hg. These airports will report the barometric
pressure as “missing” or “in excess of 31.00 “Hg.”.
5. VFR aircraft: VFR operating aircraft have no additional restrictions. Pilots must use caution when flight
planning and operating in these conditions.
6. IFR aircraft: IFR aircraft unable to set an altimeter setting above 31.00 “Hg. should apply the following:
1. The suitability of departure alternate airports, destination airports, and destination alternate
airports will be determined by increasing the published ceiling and visibility requirements when
unable to set the aircraft altimeter above 31.00 “Hg. Any reported or forecast altimeter setting over
31.00 “Hg. will be rounded up to the next tenth to calculate the required increases. The ceiling will
be increased by 100 feet and the visibility by 1/4 statute mile for each 1/10 “Hg. over 31.00 “Hg. Use
these adjusted values in accordance with operating regulations and operations specifications.

EXAMPLE-

Destination airport altimeter is 31.21 “Hg. The planned approach is an instrument landing
system (ILS) with a decision altitude (DA) 200 feet and visibility 1/2 mile (200-1/2). Subtract 31.00
“Hg. from 31.21 “Hg. to get .21 “Hg. .21 “Hg. rounds up to .30 “Hg. Calculate the increased
requirement: 100 feet per 1/10 equates to a 300 feet increase for .30 “Hg. 1/4 statute mile per
1/10 equates to a 3/4 statute mile increase for .30 “Hg. The destination weather requirement is
determined by adding the 300-3/4 increase to 200-1/2. The destination weather requirement is
now 500-1 1/4.

2. 31.00 “Hg. will remain set during the complete instrument approach. The aircraft has arrived at the
DA or minimum descent altitude (MDA) when the published DA or MDA is displayed on the
barometric altimeter.

NOTE-

The aircraft will be approximately 300 feet higher than the indicated barometric altitude using
this method.

3. These restrictions do not apply to authorized Category II/III ILS operations and certificate holders
using approved atmospheric pressure at aerodrome elevation (QFE) altimetry systems.
7. The FAA Flight Procedures & Airspace Group, Flight Technologies and Procedures Division may
authorize temporary waivers to permit emergency resupply or emergency medical service operation.
Coffin Corner: Where Stall And Overspeed Meet

Most of us have never had to worry about exceeding V NE - especially in level flight. And in a
piston airplane, VNE is about as far away from stall speed as you can get.

But, the same isn't true in a jet. Especially a subsonic one. At a jet's operating ceiling, its
Maximum Mach Number (MMO) is often extremely close to its stall speed. And that region of
flight is called the "Coffin Corner"

The coffin corner's real name is the "Q Corner", because "Q" is the abbreviation for dynamic
pressure. Coffin corner occurs from the interaction between stall speed and critical mach
speed, which are both caused by pressure over your wing. So, "Q Corner" is the techie name,
but coffin corner sounds more dramatic.

The region is deadly. Get too slow, and you'll stall the jet at high altitude (not something you
want to do). Get too fast, and you'll exceed your critical mach number. The air over your
wings will go supersonic, you'll pitch down, the aircraft will accelerate, and your wings will fall
off. Also bad.
So why does this happen at a jet's maximum ceiling? As you increase altitude, true stall
speed increases, and the true airspeed to reach MMO decreases. Coffin corner is the region
just below their intersection.

True Stall Speed: Increasing With Altitude

As you climb, the air becomes less dense, and your wings need more airflow to generate the
same amount of lift. So, as you climb, your true stall speed increases. This is true in a prop,
turboprop, or jet.
Maximum Mach Number (MMO): Preventing Your Wing From Going Supersonic

As you climb, the true airspeed to reach MMO decreases. In sub-sonic jets, MMO prevents
you from reaching your critical mach number. That's the speed where some air flowing over
your wings begins traveling at the speed of sound.

As the air flows over your wing, it accelerates. At some point, the air in front of your wing may
be subsonic, but it will accelerate past the speed of sound as it flows over the wing's upper
surface. Once this happens, a shock wave forms. Turbulent flow develops behind the wing,
causing a buffet called mach buffet.

You can also experience mach tuck after you pass the critical mach number. The supersonic
air flowing past the leading edge of the wing generates far more lift than the subsonic (and
often stalled) air aft of the shock wave. At this point, your center of lift is located at the leading
edge of your wing. As you speed up, that shockwave moves aft. The high-lift region follows it
aft, and the the center of lift moves aft, as well.
As your center of lift moves aft, your nose begins to pitch down. You accelerate more, and the
center of lift moves farther aft. If your tail uses a conventional elevator, you may find that
supersonic flow also limits its effectiveness, and you can't raise the nose. Now you're a
missile.

The Speed Of Sound Decreases As It Gets Colder

Supersonic flow is the main limitation of a sub-sonic jet's MMO. And while MMO is a fixed
number (e.g. 0.85 Mach), the true airspeed where you reach MMO decreases as the air gets
colder.

As you climb in altitude, air temperature decreases. That's why jet aircraft have a
moveable "barber pole" needle to show MMO, that automatically decreases with temperature.
Glass panel aircraft use a similar digital marking.
Coffin Corner: Where They Meet

As you approach the aircraft's maximum ceiling, you'll find that MMO and stall speed meet, or
at least get close.

Most of today's jets have a fairly wide margin between stall and MMO, but a great example of
a coffin corner aircraft is the U-2. At high altitudes, the U-2 can have as little as a 5
knots between stall and mach buffet. That leaves no room for error.

In the coffin corner, aircraft become extremely difficult to fly, and an autopilot is a necessity.
Your margin for error is small, hence the name "coffin corner." But, despite its name, it still
would be fun to try.

How to calculate the rate of descent


During an approach, we need a way to calculate our descent rates quickly and easily. The
most common IFR glideslope and VFR VASI or PAPI-guided approaches are set up for a
3-degree descent.
Although the more precise mathematical rate of descent calculations can be done,
approximating using one of the two rule-of-thumb methods below is sufficient. These
methods can be used to calculate the approximate rate of descent needed to achieve a
typical 3-degree descent angle.

Rate of descent calculation method 1

It is important to learn about the rate of descent formula, the first method of calculating the
necessary rate of descent is to use the following equation:
Groundspeed x 5
The resulting number is our approximate descent rate in feet per minute. For example, if
our groundspeed is 100 knots (GS), and we multiply it by 5, that would equate to a 500
FPM descent rate to achieve a 3-degree descent angle.
Example:
100 x 5 = 500

Rate of descent calculation method 2

The necessary rate of descent formula for a 3-degree angle can also be calculated with
the following formula:
Groundspeed ÷ 2 + add one decimal place
After dividing groundspeed in half, add one decimal place to the end of the answer to get
your target rate of descent.
For example, if we take our same 100 groundspeed and divide it in half, the answer is 50.
Adding another decimal place to the end provides the same 500 FPM descent rate as
method 1.
Example:
100 ÷ 2 = 50
Add a 0 the end of 50 to get 500.

-------------------------------------------------------------------------------------------------------------------------
On this approach, you need to do a little more math, but it's still pretty straight forward. From
the FAF (RLG VOR) to MAPRN, you need to lose about 1,500 feet (10,600-9120=1480).

You have 4.3NM to accomplish the descent. If you round that down to 4NM to make the
math a little easier, you'll need to descend at 375 feet per nautical mile (1500/4=375).

Going back to the 1 In 60 Rule, that means you'll need to pitch down 3.75 degrees (375
FPNM/100) to accomplish the descent. Round that to a 4 degree descent, which will be easier
to estimate on your PFD, and you're on your way.

If you're backing up the descent with your VSI, at 120 knots, you'll need about an 800 FPM
descent rate (4 degrees X 2 MPM X 100 = 800 FPM).
If you're flying at 90 knots, you'll need about a 600 FPM descent rate (4 degrees X 1.5 MPM
X 100 = 600 FPM).

Pitot/Static Problems

ALIMETRY ERRORS

High-to-low, hot-to-cold, look out below! This aviation axiom is a quick breakdown of the most common
altimetry errors. Without going into too much detail, the basics of altimetry boil down to the altimeter
measuring altitude through changes in atmospheric pressure as the aircraft climbs and descends. If the
pressure and temperature remain constant, and there are no other altimeter failures or a static blockage, this
works great.
Transitioning from an area of high pressure to an area of low pressure feels the same to the altimeter as
climbing. If the altimeter is not set correctly, the aircraft could end up lower than the indicated altitude would
portray. In a similar vein, flying in colder, denser air can cause the altimeter to show higher than the aircraft’s
true altitude. So it is critical to ensure the altimeter is set correctly during all phases of flight. If there is
excessively high or low pressure—so high or low that the altimeter cannot be accurately set—flight operations
should not be conducted in that aircraft.

There are two critical reasons to ensure the aircraft’s altimeter is set. The first is traffic avoidance. If ATC has
an IFR arrival at 3000 feet and a transient VFR aircraft below it at 2500 feet, even a few hundred feet of error
on either aircraft could create hazards.

The second reason an updated altimeter is important is for terrain avoidance. When operating above
mountainous terrain, atmospheric conditions can exist causing the altimeter to indicate up to 1000 feet higher
than the aircraft’s actual altitude. Factoring in this, human error and downdrafts, it is prudent to give high,
mountainous terrain as wide a berth as you can afford.

PITOT TUBE BLOCKAGE

A pitot tube blockage can be caused by dirt, moisture, ice or even bugs. And, of course, aborted takeoffs after
a failure to note “airspeed alive” on the takeoff roll, only to later find the pitot-tube cover still on, almost
mockingly. Assuming a proper preflight inspection, ensuring the pitot tube is free and clear, obstructions can
still unfortunately occur during flight. There are a few possible scenarios.

Blocked pitot tube with unblocked drain hole: This would result in the airspeed indicator reading zero.
Because the pitot tube would not be able to sense any airflow, and the drain hole would let any residual air
out, there would be no pressure differential for the airspeed indicator to measure.

Blocked pitot tube with blocked drain hole: With trapped dynamic pressure, the airspeed will indicate
whatever speed was showing when the blockages occurred. If the aircraft’s static port remains unblocked,
pressure will increase as the aircraft descends and decrease as the aircraft climbs.

If static pressure decreases during a climb, the trapped ambient pressure with allow the diaphragm to expand,
showing an increase in airspeed. Conversely, descending into denser air will force the diaphragm to close,
showing a decrease in airspeed. In short, the airspeed will act in a manner similar to an altimeter.

STATIC PORT BLOCKAGE

Blockages of the static port will affect all pitot/static instruments.

Airspeed Indicator: If the static port is blocked but the pitot tube remains clear, the ASI will function but not
with accuracy. If speed remains constant, and the aircraft climbs or descends, the static pressure will result in
changes to your airspeed indication. It’s the same principle as a blocked pitot tube but reversed.
A climb will result in a decrease in atmospheric pressure entering the pitot tube, which will result in the ASI
indicating decreasing airspeed, causing the ASI to act as a “reverse altimeter.” As long as the pitot tube is not
blocked, changes in airspeed will reflect on the ASI, but not accurately.

Vertical Speed Indicator: With no changes to static air, there can be no differential pressure for the VSI to
work with. With no differential pressure, the VSI will be stuck at zero, and you’ll marvel at how well you’re
maintaining altitude.

Altimeter: The altimeter will freeze at whatever altitude was indicated when the static port was blocked.

----------------------------------------------------------------------------------------------------------------------------------------------

Effects

 If the pitot probe is blocked but the pitot drain and static ports are free, then in straight and level (cruising) flight
the displayed IAS will tend to reduce, eventually indicating zero.
 If the pitot probe and pitot drain are blocked but the static port is free then the IAS will increase during a steady
climb and decrease during a steady descent.
 If the pitot probe, pitot drain, and static ports are all blocked then the IAS will remain constant despite changes
in actual airspeed.
In addition to airspeed indicators, systems which rely on information directly or indirectly (via Air Data
Computers) from the pitot-static system are also unreliable if the pitot static system is blocked in some way.
 If the static vent only is blocked, then the altimeter will freeze on the altitude that the blockage occurred,
the VSI will show zero climb or descent, and the IAS will over-read in the descent or under-read in the climb.

Traffic Collision Avoidance System (TCAS)


An aircraft system based on secondary sumeillance radar (SSR) transponder signals which operates
independently of ground-based equipment to provide advice to the pilot on potential conflicting aircraft that are
equipped with SSR transponders.
Source: ICAO Annex 10, vol. IV

Types of ACAS
 ACAS I gives Traffic Advisories (TAs) but does not recommend any manoeuvres. The only implementation of
the ACAS I concept is TCAS I. These equipments are limited to interoperability and interference issues with
ACAS II.
 ACAS II gives Traffic Advisories (TAs) and Resolution Advisories (RAs) in the vertical sense (direction). The
only implementations of the ACAS II concept are TCAS II Version 7.0 and Version 7.1.
 ACAS III gives TAs and RAs in vertical and/or horizontal directions. ICAO SARPs for ACAS III have not been
developed. Currently, there are no plans to proceed with such a development.

Not all TCAS systems can be considered as accepted ACAS.


TCAS I is mandated in the United States for certain smaller aircraft.
TCAS II Versions 7.0 and 7.1 are mandated in Europe and elsewhere Collision avoidance systems can be
passive, like for example the PCAS or Portable Collision Avoidance System which only monitors the
surrounding aircraft without emitting any signal. This portable system is often used within General Aviation.
Some gliders are now equipted with an avoidance system named FLARM.

The TCAS system can be implemented as:


 a dedicated instrument
 a combined instrument with the vertical speed indicator
 a combined instrument with the Navigation Display or the Electronic Horizontal Situation Indicator.

Different modes
TCAS can be currently operated in the following modes:
 Standby: TCAS does not issue any interrogations and the transponder only replies to discrete interrogations.
This mode is used on the ground, outside of the runway.
 Transponder: the transponder replies to all appropriate ground and TCAS interrogations and TCAS remains in
stand-by. This is a passive mode which is the minimum mandatory mode to be set by any airborne aircraft
(typically in General Aviation)
 Traffic Advisory (TA): TCAS issues the appropriate interrogations and perform all tracking functions. However,
TCAS will only provide traffic advisories (TA) and the resolution advisories (RA) are inhibited
 Automatic (TA/RA): TCAS provides traffic advisories (TA) and resolution advisories (RA) when appropriate.
This mode is mandatory for all Commercial Aviation aircrafts.

Technical description
The TCAS II is a system that:
 Monitors the airspace around the aircraft and communicates with all traffic equipped with a corresponding
active transponder
 Queries all surrounding aircraft on the frequency 1030 MHz and each aircraft transponder replies on the
frequency 1090MHz
 Warns pilots of the presence of other transponder-equipped aircraft which may present a threat of mid-air
collision (MAC)

Regulatory Requirements

Aircraft registered in the U.S. and operating under Part 91 of the FARs are not required to be equipped with
TCAS. However, if an aircraft is equipped, it must be an approved system operating under the regulations
contained in FAR 91.221. For operations conducted under FAR part 135, the aircraft must be equipped with
TCAS if it is turbine powered and has 10 to 30 passenger seats (FAR 135.180). Whether the aircraft is operated
under part 91 or part 135, if it is equipped with TCAS II, it must be version 7 (TSO C-119).
The European Aviation Safety Agency (EASA) requires ACAS II (effectively TCAS II, version 7.1) for all
fixed wing turbine powered aircraft that have a maximum takeoff weight of greater than 5,700 kg (12,566 lbs)
or have more than 19 passenger seats. This requirement applies to all flights conducted in European Union
airspace.

ICAO Standards
Operators should also be familiar with the International Civil Aviation Organization (ICAO) Standards and
Recommended Practices (SARPS) regarding TCAS/ACAS.
For non-commercial aircraft, the SARPS in Annex 6, Part 2 state:
Annex 6 Part II: International General Aviation Aeroplanes

3.6.10 Aeroplanes required to be equipped with an airborne collision avoidance system (ACAS)
3.6.10.1 Recommendation.— All turbine-engined aeroplanes of a maximum certificated take-off mass in
excess of 15 000 kg, or authorized to carry more than 30 passengers, for which the individual airworthiness
certificate is first issued after 24 November 2005, should be equipped with an airborne collision avoidance
system (ACAS II).
3.6.10.2 All turbine-engined aeroplanes of a maximum certificated take-off mass in excess of 15 000 kg or
authorized to carry more than 30 passengers, for which the individual airworthiness certificate is first issued
after 1 January 2007, shall be equipped with an airborne collision avoidance system (ACAS II).
3.6.10.3 Recommendation.— All turbine-engined aeroplanes of a maximum certificated take-off mass in
excess of 5 700 kg but not exceeding 15 000 kg, or authorized to carry more than 19 passengers, for which the
individual airworthiness certificate is first issued after 1 January 2008, should be equipped with an airborne
collision avoidance system (ACAS II).
For commercial aircraft, the SARPS in Annex 6, Part 1 state:
Annex 6 Part I: International Commercial Air Transport- Aeroplanes

6.18 Aeroplanes required to be equipped with an airborne collision avoidance system (ACAS II)
6.18.1 From 1 January 2003, all turbine-engined aeroplanes of a maximum certificated take-off mass in excess
of 15 000 kg or authorized to carry more than 30 passengers shall be equipped with an airborne collision
avoidance system (ACAS II).
6.18.2 From 1 January 2005, all turbine-engined aeroplanes of a maximum certificated take-off mass in excess
of 5 700 kg or authorized to carry more than 19 passengers shall be equipped with an airborne collision
avoidance system (ACAS II).
6.18.3 Recommendation.— All aeroplanes should be equipped with an airborne collision avoidance system
(ACASII).
6.18.4 An airborne collision avoidance system shall operate in accordance with the relevant provisions of
Annex 10,Volume IV.

What Are V-Speeds?


V-speeds are specific airspeeds that are defined for operational reasons, such as limitations (e.g., maximum
flaps extended speed – VFE) or performance requirements (e.g., best rate of climb speed – VY).

In other words, V-speeds serve as critical benchmarks that guide pilots in managing the aircraft’s performance
and ensuring safety.

For example, the rotation speed (V R) is the speed at which the pilot initiates a gentle rotation of the aircraft to
lift off the ground during takeoff.

A V-speed may change depending on factors such as aircraft weight and weather conditions, but its designation
(e.g., VR) remains the same.
You may find several V-speeds on the internet that aren’t listed here. That’s because the V-speeds we’re talking
about today are defined in 14 CFR Part 1, as well as 14 CFR Part 23 and Part 25 (used for aircraft certification).

Any other V-speeds you encounter are likely manufacturer-specific and aren’t regarded as official V-speeds by
the Federal Aviation Administration (FAA).

Mach Numbers and V-Speeds

You may find V-speeds with an “M” instead of the usual “V” (MMO instead of VMO, for example).

This means that the particular speed is defined using a Mach number.

V-speeds can be defined using any type of airspeed, such as knots or miles per hour, but the designation
remains “V” unless a Mach number is used – then it becomes “M”.

V-
Spee Description
d

VA Design maneuvering speed.

VB Design speed for maximum gust intensity.

VC Design cruising speed.

VD Design diving speed.

VDF Demonstrated flight diving speed.

Speed at which the critical engine is assumed to fail during


VEF
takeoff.

VF Design flap speed.

VFC Maximum speed for stability characteristics.

VFE Maximum flap extended speed.

VFTO Final takeoff speed.

VH Maximum speed in level flight with maximum continuous


power.

VLE Maximum landing gear extended speed.

VLO Maximum landing gear operating speed.

VLOF Lift-off speed.

VMC Minimum control airspeed with the critical engine inoperative.

VMO Maximum operating limit speed.

VMU Minimum unstick speed.

VNE Never-exceed speed.

VNO Maximum structural cruising speed.

VR Rotation speed.

VREF Reference landing speed.

Stalling speed or minimum steady flight speed at which the


VS
airplane is controllable.

VS0 Stall speed in the landing configuration.

Stall speed in a specific configuration (e.g., ‘clean’


VS1
configuration).

VSR Reference stall speed.

VSR0 Reference stall speed in the landing configuration.

VSR1 Reference stall speed in a specific configuration.

Speed at which onset of natural or artificial stall warning


VSW
occurs.

VTOSS Takeoff safety speed for Category A aircraft.

VX Speed for best angle of climb.


VY Speed for best rate of climb.

V1 Takeoff decision speed.

V2 Takeoff safety speed.

V2min Minimum takeoff safety speed.

Most Important V-Speeds Explained


Let’s take a look at the V-speeds you’re most likely to encounter – and the ones you should know.
As we go through them, use the Pilot’s Operating Handbook (POH) for the airplane you fly, and make a note of
the speed for each V-speed. If it isn’t defined in the POH or is variable, make sure you know how to calculate it.
You’ll make your life a whole lot easier if you take the time to memorize them.
VR: Rotation Speed
VR is the speed at which the pilot gently pulls back on the control column to lift the nose off of the runway
during takeoff.
For most commercial aircraft, V R varies for each takeoff depending on the weight and configuration of the
aircraft as well as environmental factors like weather or runway conditions.
In most General Aviation (GA) aircraft, VR is usually the same regardless of conditions.
It might seem obvious, but VR cannot be less than the stall speed (VS1 – more on that later).
VX: Best Angle of Climb Speed
VX is the airspeed that provides the best angle of climb. In other words, if you maintain V X, you’ll gain the most
altitude in the shortest horizontal distance.
This speed is your go-to for a short-field takeoff, particularly when there are obstacles that you need to climb
above during takeoff.
You should practice climbing at V X (and short-field takeoffs) regularly, as it is a critical skill during short-field
operations.
VY: Best Rate of Climb Speed
VY is the airspeed for best rate of climb. In other words, if you maintain V Y, you’ll gain the most altitude in the
shortest amount of time.
Compared to VX, you’ll use more horizontal distance.
VY is the speed typically used during climb.
VA: Maneuvering Speed
VA is the aircraft’s design maneuvering speed. It is the speed above which you risk damaging the aircraft’s
structure if you make a full deflection of a flight control (e.g., full-up elevator).
If you make a full deflection of a flight control at or below V A, the aircraft will stall before the structure is
damaged.
You should not use full deflection of any flight control above V A. That being said, repeated full deflection of
any flight controls (such as full right rudder and then full left rudder, for example) is not recommended, even
below VA.
VA isn’t a fixed figure; it varies with weight. If the aircraft’s weight decreases, V A decreases as well, and vice
versa.
VFE: Maximum Flaps Extended Speed
VFE, or maximum flap extended speed, is the highest speed permissible with the flaps extended.
This speed is your boundary marker when flying with flaps down, ensuring you don’t cause potential structural
damage.
Not all aircraft treat VFE as a singular speed regardless of flap setting. Most aircraft, like the Cessna 172, have
different VFE speeds for different flap settings.
In the Cessna 172, you can fly with 10 degrees of flaps below 110 knots. Anything more than 10 degrees of
flaps, and you’re limited to 85 knots instead.
VLE: Maximum Landing Gear Extended Speed
VLE, or maximum landing gear extended speed, is the top speed at which you can safely fly with the landing
gear extended.
A related speed is VLO, or maximum landing gear operating speed, the speed above which you cannot extend or
retract the landing gear.
VLO is typically lower than VLE due to the aerodynamic forces exerted on the landing gear during extension or
retraction.
VNE: Never Exceed Speed
VNE, or “never exceed” speed, is exactly that. The speed above which you should never venture under any
circumstances.
VNO: Maximum Structural Cruising Speed
VNO, the maximum structural cruising speed, is the highest speed that you can safely fly in smooth air.
VNO is marked by the upper limit of the green arc on the airspeed indicator.

If you’re above VNO (in the yellow arc or “caution range”) and you encounter air that is not smooth, you could
cause damage to the aircraft.
For example, if you encounter turbulence, the “bumps” you experience will increase the load factor. If you fly
above VNO in these conditions, the increase in load factor could damage the aircraft’s structure.
VS: Stall Speed
VS represents stall speed, essentially the lowest speed your aircraft can maintain steady flight.
When it comes to VS, there’s an important caveat.
An aircraft can stall at any speed.
A stall occurs when the aircraft exceeds the critical angle of attack. This can happen at any airspeed.
Say a pilot is descending at a high airspeed, far from V S. If they quickly pitch up, the aircraft may exceed the
critical angle of attack and stall, despite being at a high airspeed.
So, why do we define VS?
Well, in a “normal” attitude (think straight-and-level), the aircraft is only at risk of stalling if:
1. The pilot makes a dramatic control input that quickly increases the angle of attack, or
2. The pilot maintains altitude while the airspeed decreases, gradually increasing the angle of attack and
eventually stalling at VS.
So, can the aircraft stall at any airspeed? Yes.
When is it most likely to stall? At VS.
The V-speed for stall speed is divided into two types:
1. VS0 – the stall speed in the landing configuration (e.g., flaps and gear down)
2. VS1 – the stall speed in a specific configuration (e.g., ‘clean’ – flaps and gear up)
The difference between the stall speed with the flaps down versus the flaps up is significant, so it makes sense
to differentiate between the two.
One final note about VS.
Every manufacturer determines the stall speed for their aircraft. The test for stall speed is performed with the
throttle closed at maximum takeoff weight.
This means that you may experience a lower stall speed than published in the POH if you’re flying at a lower
weight or the throttle isn’t closed.
For more information on stall speed testing regulations, see AC 23-8C, § 23.49, page 15.
V1: Takeoff Decision Speed
V1, or the takeoff decision speed, is the speed by which the decision to continue the takeoff or abort must be
made.
The primary purpose of V1 is to serve as a decision point. If a critical system fails (such as an engine) or other
anomalies occur before reaching V1, there will be sufficient runway remaining to abort the takeoff safely.

However, once V1 is surpassed, the takeoff should continue, as there will not be enough runway left to stop
safely.
V1 is not a fixed number and is calculated before each takeoff, taking into account several factors, including
aircraft weight, runway length, environmental conditions, and aircraft performance data.
V1 is where the pilot must take the first action (such as reducing thrust) to stop the aircraft, or risk a runway
overrun.
It’s important to note that V1 also relates to the aircraft’s performance capability in case of an engine failure.
After V1, the aircraft must have the performance capability to continue the takeoff on the remaining engines and
achieve the required climb performance.
That’s where V2, or takeoff safety speed, comes into play.
V2: Takeoff Safety Speed
V2, known as the takeoff safety speed, is the minimum speed at which the aircraft can maintain a specified rate
of climb with one engine inoperative.
The primary goal of V2 is to ensure a safe climb gradient in an engine failure scenario. This speed ensures that
the aircraft can maintain a positive rate of climb to clear obstacles and reach a safer altitude.
The aircraft must be able to achieve V 2 at a minimum of 35 ft above the end of the runway distance after an
engine failure at V1.
VEF: Critical Engine Failure Speed During Takeoff
VEF is the worst possible speed the critical engine can fail while allowing the takeoff to be completed
successfully.
Interestingly, it is not at V1, but actually before.
This may sound strange, because we should abort the takeoff if an engine failure occurs before V1, right?
Well, regulations state that takeoff performance calculations should account for an engine failure that is close
enough to V1 that the pilot does not have enough time to abort at V1.
In other words, if the engine fails right before V 1 without enough time to react, the aircraft must be able to take
off safely and achieve V2 at the specified height and distance.
VMC: Minimum Control Speed
VMC, or minimum control speed, represents the lowest speed at which a multi-engine aircraft can maintain
controlled flight with one engine inoperative and the other at full power.
VEF may not be less than VMC, and V2min may not be less than 1.1 times VMC.
VMC is often divided into two distinct speeds: V MCA and VMCG, each addressing a different aspect of aircraft
control under asymmetric thrust conditions.
VMCA: Minimum Control Speed Air
VMCA is the minimum speed at which the aircraft can maintain controlled flight in the air with one engine failed
and the other at full power.
Below VMCA, the aircraft may become uncontrollable due to the loss of directional control, making it a critical
speed to be aware of during flight operations.
VMCG: Minimum Control Speed Ground
VMCG, on the other hand, is the minimum speed at which the aircraft can maintain directional control on the
ground with one engine inoperative and the other at full power.
It’s a vital speed to know during the takeoff roll, ensuring that control can be maintained if an engine fails
during takeoff.
Air Temperature
The temperature of the air outside an aircraft is measured and indicated within the cockpit or used, together with
outputs from the Pitot Static System, as an input to aircraft equipment, e.g. Air Data Computer (ADC).
Outside Air Temperature (OAT)
The ambient temperature measured outside an aircraft is known as the Outside Air Temperature (OAT) or Static
Air Temperature (SAT). The sensor which detects OAT must be carefully sited to ensure that airflow over it
does not affect the indicated temperature.

Total Air Temperature (TAT)


If temperature is measured by means of a sensor positioned in the airflow, kinetic heating will result, raising the
temperature measured above the OAT. The temperature measured in this way is known as the Total Air
Temperature (TAT) and is used in ADCs to calculate True Airspeed (TAS). Careful design and siting of the
TAT probe is necessary to ensure accurate measurement of TAT.
Terrain awareness and warning system
In aviation, a terrain awareness and warning system (TAWS) is generally an on-board system aimed at
preventing unintentional impacts with the ground, termed "controlled flight into terrain" accidents, or CFIT. The
specific systems currently in use are the ground proximity warning system (GPWS) and the enhanced ground
proximity warning system (EGPWS).[1] The U.S. Federal Aviation Administration (FAA) introduced the generic
term TAWS to encompass all terrain-avoidance systems that meet the relevant FAA standards, which include
GPWS, EGPWS and any future system that might replace them.
As of 2007, 5% of the world's commercial airlines still lacked a TAWS. A study by the International Air Transport
Association examined 51 accidents and incidents and found that pilots did not adequately respond to a TAWS
warning in 47% of cases.
Several factors can still place aircraft at risk for CFIT accidents: older TAWS systems, deactivation of the EGPWS
system, or ignoring TAWS warnings when an airport is not in the TAWS database.

TAWS types

Class A TAWS includes all the requirements of Class B TAWS, below, and adds the following additional
three alerts and display requirements of:

 Excessive closure rate to terrain alert


 Flight into terrain when not in landing configuration alert
 Excessive downward deviation from an ILS glideslope alert
 Required: Class A TAWS installations shall provide a terrain awareness display that shows either the
surrounding terrain or obstacles relative to the airplane, or both.
Class B TAWS is defined by the U.S. FAA as: A class of equipment that is defined in TSO-C151b
and RTCA DO-161A.[As a minimum, it will provide alerts for the following circumstances:

 Reduced required terrain clearance


 Imminent terrain impact
 Premature descent
 Excessive rates of descent
 Negative climb rate or altitude loss after takeoff
 Descent of the airplane to 500 feet above the terrain or nearest runway elevation (voice callout "Five
Hundred") during a non-precision approach.
 Optional: Class B TAWS installation may provide a terrain awareness display that shows either the
surrounding terrain or obstacles relative to the airplane, or both.
Class C defines voluntary equipment intended for small general aviation airplanes that are not required to
install Class B equipment. This includes minimum operational performance standards intended for piston-
powered and turbine-powered airplanes, when configured with fewer than six passenger seats, excluding
any pilot seats. Class C TAWS equipment shall meet all the requirements of a Class B TAWS with the
small aircraft modifications described by the FAA.] The FAA has developed Class C to make voluntary
TAWS usage easier for small aircraft.
Different Aircraft Weights
Standard Empty Weight: The airframe, engines, and all fixed operating equipment.
Basic Empty Weight (BEW): Includes the airplane, optional equipment, unusable fuel, and all
operating fluids.
Licensed Empty Weight: Mirrors Basic Empty Weight but excludes full engine oil, accounting
only for undrainable oil.
Gross Landing Weight: The takeoff weight minus the fuel burned en route.
Ramp Weight: The weight of the airplane loaded for flight before the engine start.
Zero Fuel Weight (ZFW): The weight of the aircraft before the addition of fuel.
Gross Takeoff Weight: The weight of the airplane just before brake release to begin the takeoff
roll.
Useful Load: The weight of crew, cargo, passengers, baggage, and usable fuel.
Maximum Ramp Weight (MRW): The maximum weight for ground operations.
Maximum Takeoff Weight (MTOW): The maximum weight for takeoff.
Maximum Landing Weight (MLW): The maximum weight for landing is based on the stress of
impact on the gear.
Payload: The weight of passengers, cargo, and baggage.
Standard Empty Weight
Standard Empty Weight is the weight of the airplane straight from the manufacturer, with all the standard
equipment installed. It includes the airframe, engines, and all items of operating equipment that have fixed,
permanently installed locations in the airplane.
To get even more specific, it includes the weight of fixed ballast, hydraulic fluid, unusable fuel, and full engine
oil.
By unusable fuel, we mean fuel that can’t be drained. Similarly, full engine oil refers to the maximum oil
capacity of the aircraft’s engines.
When you hear “Standard Empty Weight,” think of it as the baseline or starting point for all other weight
calculations. It’s like the weight of a naked aircraft before passengers, cargo, and usable fuel come into play.
Basic Empty Weight (BEW)
Basic Empty Weight is a step up from the Standard Empty Weight, as it includes the weight of the airplane,
optional equipment, unusable fuel, and full operating fluids, including full engine oil.
What does “optional equipment” fall under? Anything not included in the standard equipment list. It might
include extra avionics, upgraded seats, or additional fuel tanks. Each piece of optional equipment adds weight to
the aircraft, which is why it’s part of the BEW.
Remember, BEW also includes full operating fluids. That means hydraulic fluid, coolant, and full engine oil.
Basic Empty Weight is the base weight used for all weight and balance calculations.
Licensed Empty Weight
Licensed Empty Weight is very similar to the Basic Empty Weight (BEW) but has one key difference: engine
oil.
While the BEW includes full engine oil in its calculation, the Licensed Empty Weight does not. Instead, it only
includes undrainable oil, which refers to the oil remaining in the system after it has been drained to the
maximum extent possible.
Why this distinction?
In some jurisdictions, aviation authorities only consider undrained oil for certification purposes. This is because
it is considered part of the aircraft’s structure, similar to unusable fuel.
The rest of the elements – the weight of the airplane, optional equipment, and unusable fuel – remain the same
as in the BEW.
So, when you think of Licensed Empty Weight, think of it as the Basic Empty Weight, but with a focus only on
the oil that can’t be drained from the aircraft’s engine.
Gross Landing Weight
The Gross Landing Weight refers to the weight of the airplane when it’s ready to land. Simply put, it’s the
takeoff weight minus the fuel burned en route.
Fuel consumption plays a significant role in this weight calculation. As the aircraft flies and burns fuel to keep
its engines running, it becomes lighter. So, the weight of the airplane when it touches down at its destination
(the Gross Landing Weight) is less than when it took off.

Understanding the Gross Landing Weight is critical because it directly impacts the safety of the aircraft. Each
aircraft has a Maximum Landing Weight, which is the heaviest weight it can safely land at without causing
undue stress on its landing gear and structure.
If the Gross Landing Weight exceeds the Maximum Landing Weight, it could harm the aircraft’s structure or
cause a catastrophic failure. Therefore, pilots must manage their fuel consumption and cargo load to ensure
the Gross Landing Weight stays within safe limits.
Ramp Weight
Ramp Weight, also known as Taxi Weight, refers to the weight of the airplane when loaded for flight, just
before the engine start.
Ramp weight includes everything; The weight of the aircraft, crew, passengers, baggage, cargo, and most
importantly, all the fuel needed for the flight.
Unlike other weight terms, Ramp Weight includes the entire fuel load, even the fuel you will burn during taxi
and run-up. That’s why it’s often slightly higher than the Gross Takeoff Weight (which we’ll discuss later).
Knowing the Ramp Weight helps pilots ensure they’re within the aircraft’s Maximum Ramp Weight, another
safety limitation set by the aircraft manufacturer. Going over this maximum could jeopardize the structural
integrity of the aircraft or affect its performance.
So when you think about Ramp Weight, picture the aircraft fully loaded, fueled, and ready for flight, standing
on the ramp (hence the name).
Zero Fuel Weight (ZFW)
Zero Fuel Weight refers to the weight of the airplane without any added fuel.
It includes the weight of the airplane, the crew, passengers, baggage, and cargo, but no fuel. Imagine your
aircraft, ready to go, but sitting on the runway with empty fuel tanks. That’s your Zero Fuel Weight.
Why is Zero Fuel Weight important? Because it limits the total weight of the passengers, crew, and cargo.
If you exceed this limit, you could place undue stress on the aircraft structure, as the wings (where fuel is
stored) provide significant structural support.
In other words, if the weight of the fuselage is too heavy compared to the weight of the wings, additional
stress is placed on the point where the wings connect with the fuselage.
Zero Fuel Weight
Therefore, pilots need to ensure that the weight of the fuselage doesn’t make up the largest part of the
aircraft’s weight compared to fuel weight.
Gross Takeoff Weight
Gross Takeoff Weight is the weight of the airplane just before the brakes are released to begin the takeoff roll.
This weight includes the airplane, crew, passengers, cargo, and all the usable fuel needed for the flight. It’s
essentially your Ramp Weight minus the fuel used for engine start, taxi, and run-up.
This weight is of critical importance as it directly impacts an airplane’s performance during takeoff.
More weight means longer runways are needed for takeoff, slower climb rates, and higher stall speeds.
Therefore, every aircraft has a Maximum Takeoff Weight that must not be exceeded to ensure safe flight.
Useful Load
The Useful Load is the weight of everything the aircraft carries, which can vary from flight to flight. It includes
the crew, passengers, baggage, cargo, and usable fuel.
It’s the difference between the Maximum Takeoff Weight and the Basic Empty Weight of the aircraft.
Components of Useful Load
Why is this weight term “useful”? Because it tells you how much weight you can use for different operational
needs. For example, how many passengers and how much cargo you can carry or how much fuel you can load.
The useful load varies depending on your mission, making it a functional part of flight planning.
As a pilot, managing the useful load involves finding a balance between passengers, cargo, and fuel within the
aircraft’s weight limitations to ensure a safe and efficient flight.
Maximum Ramp Weight (MRW)
Maximum Ramp Weight is the highest weight allowed for an aircraft while it’s still on the ground, just prior to
engine start for takeoff.
Why does this weight matter? Simply put, an aircraft is designed to handle specific stresses, both in flight and
on the ground. Exceeding the Maximum Ramp Weight can place excessive stress on the aircraft’s structure,
particularly the landing gear, during ground operations.
Maximum Takeoff Weight (MTOW)
Maximum Takeoff Weight is the most an aircraft can weigh to safely takeoff. It’s determined by the aircraft
manufacturer based on numerous factors, including the aircraft’s power, wing surface area, environmental
conditions, and the runway length at the typical operating airports.

Exceeding the Maximum Takeoff Weight poses serious safety risks, including reduced climb performance,
longer takeoff runs, and increased stress on the aircraft’s structure. Importantly, it could also affect the
aircraft’s ability to clear obstacles during the climb out after takeoff.
As a pilot, respecting the Maximum Takeoff Weight is an integral part of your pre-flight planning and weight
management. It’s a crucial safeguard, ensuring that your aircraft can successfully and safely defy gravity.
Maximum Landing Weight (MLW)
Maximum Landing Weight is the heaviest weight at which an aircraft can safely land.
It’s typically less than the Maximum Takeoff Weight, reflecting the fact that landing puts more stress on an
aircraft’s structure (especially the landing gear) than taking off does.
Landing an aircraft that’s over its Maximum Landing Weight weight risks structural damage and also affects
the aircraft’s braking performance, potentially leading to a runway overrun.
That’s why, if you have a situation where you need to return to the airport shortly after a full-fuel takeoff, you
may need to burn off or dump fuel to reduce the aircraft’s weight before landing.
Payload
Payload is the weight of everything an aircraft carries that serves the “purpose” of the flight. This includes
passengers, crew, baggage, and cargo. It’s the weight you’re carrying for profit or purpose.
The payload is a subset of the useful load. The useful load includes payload plus usable fuel.
An important part of flight planning is to maximize the payload while staying within the aircraft’s weight limits,
often a juggling act between fuel, passengers, and cargo.
The Gross Landing Weight refers to the weight of the airplane when it’s ready to land. Simply put, it’s the
takeoff weight minus the fuel burned en route.
Fuel consumption plays a significant role in this weight calculation. As the aircraft flies and burns fuel to keep
its engines running, it becomes lighter. So, the weight of the airplane when it touches down at its destination
(the Gross Landing Weight) is less than when it took off.
Understanding the Gross Landing Weight is critical because it directly impacts the safety of the aircraft. Each
aircraft has a Maximum Landing Weight, which is the heaviest weight it can safely land at without causing
undue stress on its landing gear and structure.
If the Gross Landing Weight exceeds the Maximum Landing Weight, it could harm the aircraft’s structure or
cause a catastrophic failure. Therefore, pilots must manage their fuel consumption and cargo load to ensure the
Gross Landing Weight stays within safe limits.

Ramp Weight
Ramp Weight, also known as Taxi Weight, refers to the weight of the airplane when loaded for flight, just
before the engine start.
Ramp weight includes everything; The weight of the aircraft, crew, passengers, baggage, cargo, and most
importantly, all the fuel needed for the flight.
Unlike other weight terms, Ramp Weight includes the entire fuel load, even the fuel you will burn during taxi
and run-up. That’s why it’s often slightly higher than the Gross Takeoff Weight (which we’ll discuss later).
Knowing the Ramp Weight helps pilots ensure they’re within the aircraft’s Maximum Ramp Weight, another
safety limitation set by the aircraft manufacturer. Going over this maximum could jeopardize the structural
integrity of the aircraft or affect its performance.
So when you think about Ramp Weight, picture the aircraft fully loaded, fueled, and ready for flight, standing
on the ramp (hence the name).
Zero Fuel Weight (ZFW)
Zero Fuel Weight refers to the weight of the airplane without any added fuel.
It includes the weight of the airplane, the crew, passengers, baggage, and cargo, but no fuel. Imagine your
aircraft, ready to go, but sitting on the runway with empty fuel tanks. That’s your Zero Fuel Weight.
Why is Zero Fuel Weight important? Because it limits the total weight of the passengers, crew, and cargo.
If you exceed this limit, you could place undue stress on the aircraft structure, as the wings (where fuel is
stored) provide significant structural support.
In other words, if the weight of the fuselage is too heavy compared to the weight of the wings, additional stress
is placed on the point where the wings connect with the fuselage.
Therefore, pilots need to ensure that the weight of the fuselage doesn’t make up the largest part of the aircraft’s
weight compared to fuel weight.
Gross Takeoff Weight
Gross Takeoff Weight is the weight of the airplane just before the brakes are released to begin the takeoff roll.
This weight includes the airplane, crew, passengers, cargo, and all the usable fuel needed for the flight. It’s
essentially your Ramp Weight minus the fuel used for engine start, taxi, and run-up.
This weight is of critical importance as it directly impacts an airplane’s performance during takeoff.
More weight means longer runways are needed for takeoff, slower climb rates, and higher stall speeds.
Therefore, every aircraft has a Maximum Takeoff Weight that must not be exceeded to ensure safe flight.
Useful Load
The Useful Load is the weight of everything the aircraft carries, which can vary from flight to flight. It includes
the crew, passengers, baggage, cargo, and usable fuel.

It’s the difference between the Maximum Takeoff Weight and the Basic Empty Weight of the aircraft.
Why is this weight term “useful”? Because it tells you how much weight you can use for different operational
needs. For example, how many passengers and how much cargo you can carry or how much fuel you can load.
The useful load varies depending on your mission, making it a functional part of flight planning.
As a pilot, managing the useful load involves finding a balance between passengers, cargo, and fuel within the
aircraft’s weight limitations to ensure a safe and efficient flight.
Maximum Ramp Weight (MRW)
Maximum Ramp Weight is the highest weight allowed for an aircraft while it’s still on the ground, just prior to
engine start for takeoff.
Why does this weight matter? Simply put, an aircraft is designed to handle specific stresses, both in flight and
on the ground. Exceeding the Maximum Ramp Weight can place excessive stress on the aircraft’s structure,
particularly the landing gear, during ground operations.
Maximum Takeoff Weight (MTOW)
Maximum Takeoff Weight is the most an aircraft can weigh to safely takeoff. It’s determined by the aircraft
manufacturer based on numerous factors, including the aircraft’s power, wing surface area, environmental
conditions, and the runway length at the typical operating airports.
Exceeding the Maximum Takeoff Weight poses serious safety risks, including reduced climb performance,
longer takeoff runs, and increased stress on the aircraft’s structure. Importantly, it could also affect the aircraft’s
ability to clear obstacles during the climb out after takeoff.
As a pilot, respecting the Maximum Takeoff Weight is an integral part of your pre-flight planning and weight
management. It’s a crucial safeguard, ensuring that your aircraft can successfully and safely defy gravity.
Maximum Landing Weight (MLW)
Maximum Landing Weight is the heaviest weight at which an aircraft can safely land.
It’s typically less than the Maximum Takeoff Weight, reflecting the fact that landing puts more stress on an
aircraft’s structure (especially the landing gear) than taking off does.
Landing an aircraft that’s over its Maximum Landing Weight weight risks structural damage and also affects the
aircraft’s braking performance, potentially leading to a runway overrun.
That’s why, if you have a situation where you need to return to the airport shortly after a full-fuel takeoff, you
may need to burn off or dump fuel to reduce the aircraft’s weight before landing.
Payload
Payload is the weight of everything an aircraft carries that serves the “purpose” of the flight. This includes
passengers, crew, baggage, and cargo. It’s the weight you’re carrying for profit or purpose.
The payload is a subset of the useful load. The useful load includes payload plus usable fuel.
An important part of flight planning is to maximize the payload while staying within the aircraft’s weight limits,
often a juggling act between fuel, passengers, and cargo.
An aircraft’s payload capacity can be a major factor in its efficiency and profitability, particularly in
commercial and cargo operations.
Distress Calls
“Mayday” is the international announcement of immediate danger to life or property, a true emergency. A
mayday call will shut down any other radio traffic on the frequency.

“Pan-pan” is an “urgent” call. Not an emergency but definitely important, and informs everyone not to interfere
with the communication.

Sécurité (/seɪˈkjʊərɪteɪ/; French: sécurité) (often repeated thrice, "Sécurité, sécurité, sécurité") is a procedure
word used in the maritime radio service that warns the crew that the following message is important safety
information. The most common use of this is by coast radio stations before the broadcast of navigational
warnings and meteorological information.

A Mayday radio call should be reserved for life threatening situations. These may include, but are not
limited to:

 Loss, or imminent loss of aircraft control for any number of different reasons
o aircraft upset by turbulence;
o pilot incapacitation;
o spatial disorientation;
o control surface or structural failure;
o engine failure that will lead to a forced landing/ditching/ejection/bailout;
 Or, an onboard fire.
A Pan-Pan call should be used for urgent situations that are not immediately life threatening, but require
assistance from someone on the ground. These include, but are not limited to:

 Becoming lost;
 A serious aircraft system failure, that requires an immediate route or altitude change;
 Other emergencies that require immediate attention and assistance from the ground.

Sécurité: A radio call that usually issues navigational warnings, meteorological warnings, and any
other warning needing to be issued that may concern the safety of life at sea, yet may not be
particularly life-threatening.
Pan-pan: This is the second most important call. This call is made when there is an emergency
aboard a vessel, yet there is not immediate danger to life, or the safety of the vessel itself. This
includes, but is not limited to injuries on deck, imminent collision that has not yet occurred, or being
unsure of vessel's position.
Mayday: This is the most important call that can be made, due to the fact that it directly concerns a
threat to life or the vessel. Some instances when this call would be made are, but not limited to death,
collision, and fire at sea. When the Mayday call is made, the vessel is requiring immediate
assistance.

Forward vs. Aft CG Explained


We’re about to demystify these terms and dive into the fascinating world
of aircraft balance. By the end of this journey, you’ll be an expert at managing CG and
appreciating how it shapes your aircraft’s performance and handling.
Here’s a quick summary:
Forward CG AFT CG
More stable (longitudinal stability) Less stable (longitudinal stability)
Less fuel efficient (more drag) More fuel efficient (less drag)
Higher stall speed (higher Angle of Attack) Lower stall speed
Good stall recovery characteristics Bad stall recovery characteristics

Understanding the Center of Gravity (CG)


First things first. What exactly is the Center of Gravity? Simply put, it’s the point where
the weight of an aircraft is perfectly balanced. Picture an invisible dot somewhere in your
aircraft—around this dot, your aircraft balances.

Now, this isn’t static. Load your aircraft differently, and the CG shifts. Add cargo to the
back, and the CG moves aft (backward). Load up the front, and it shifts forward. It’s a
fluid point, changing with every tweak in the weight (and distribution of weight) of your
aircraft.

Why does this matter? Because your CG plays a starring role in how your aircraft
performs. From stability and maneuverability to fuel efficiency, it’s the CG pulling the
strings behind the scenes.

Understanding the CG (and managing it correctly) is the ticket to safer and more efficient
flights. Let’s dive deeper into the intricacies of forward and aft CG.

Forward CG Explained
Forward CG Summary:
 More stable (longitudinal stability)
 Less fuel efficient (more drag)
 Higher stall speed (higher Angle of Attack)
 Good stall recovery characteristics
So, let’s start with a forward center of gravity, or forward CG, as it’s commonly known.
This occurs when the weight of your aircraft is tilted more toward the front. It might be
due to heavy items stored in the front, or maybe you have little or no payload in the back.

But what does a forward CG mean for you in the cockpit?

Let’s look at all the factors at play.

The lift vector (the blue line) acts through the center of the lift, while the weight vector
(the red line) acts through the center of gravity. Think of the center of lift as the lift
version of the center of gravity.

Notice how the CG is forward of the center of pressure? This creates what we call the
“lift/weight couple.” Can you guess what kind of pitching moment it produces – up or
down?
It produces a downward pitching moment. This is great when it comes to stall
characteristics – an aircraft with a forward CG will naturally resist a stall and pitch down.
A forward CG is not that great if we want to fly straight and level.

Why?

Because we need a continuous force from the horizontal stabilizer to counter-act the
nose-down pitching tendency.

This force usually comes from the deflection of the elevator. In other words, you need to
pull back on the yoke to overcome the nose-down tendency of the lift/weight couple. The
further forward the CG, the more you need to pull back.
Naturally, this creates drag, and drag isn’t our friend. More drag leads to higher fuel
consumption and lower efficiency.

You will, of course, use the trim to keep the elevator in the correct position and relieve
any pressure on the control column.

So, how does this affect stability?

For starters, an aircraft with a forward CG is more stable. There are many types of
stability, but, generally speaking, stability is your aircraft’s ability to resist motion. In
other words, how easily your aircraft pitches, yaws, or rolls.
You will require a larger elevator input to change the pitch of an aircraft with a forward
CG – making it more stable and less maneuverable. In other words, the longitudinal
(pitch) stability is increased.
But, there’s the issue of efficiency.

With a forward CG, the increase in the force from the horizontal stabilizer creates more
drag, leading to higher fuel consumption.

Furthermore, we also need a higher Angle of Attack (AoA) to maintain level flight
(which also creates more drag).

This is because, in order for the aircraft to maintain level flight, lift must equal weight.
The issue is that we’re creating a downward force with the horizontal stabilizer that
effectively adds to the weight. So now weight is effectively more than lift (actual weight
plus downward force).

What do we need to do so that lift is equal to effective weight? We need to increase lift.
This means we need a higher angle of attack.
A higher angle of attack creates more drag and reduces our efficiency.

A high angle of attack isn’t necessarily a bad thing – it’s often needed for certain
maneuvers and during takeoff and landing. However, consistently needing a high angle of
attack during cruise flight (when the aircraft needs to be at its most efficient) leads to
higher drag and lower fuel efficiency.

This also leads to an increase in the stall speed because operating at a higher angle of
attack brings the aircraft closer to the critical angle of attack (where the aircraft will stall).

So while a forward CG can provide increased stability and better stall recovery
characteristics, it has a negative impact on aircraft performance.

Now, let’s turn our attention aft.

Aft CG Explained
Aft CG Summary:

 Less stable (longitudinal stability)


 More fuel efficient (less drag)
 Lower stall speed
 Bad stall recovery characteristics
Now let’s shift our focus to the other end of the spectrum – the aft center of gravity, or aft
CG. This happens when the weight in your aircraft is skewed towards the back. Perhaps
you’ve loaded heavy cargo in the rear, or maybe you’re the only one in the front with
passengers in the back.

So, what’s the impact of an aft CG on your flight?

On the bright side, an aft CG can make your aircraft more efficient. With the weight
towards the rear, your aircraft doesn’t need as large of a force from the horizontal
stabilizer, and it doesn’t need a much higher AoA to balance out lift with weight. This
leads to less drag and better fuel efficiency – a win for the aircraft’s performance and
your wallet.

But it’s not all roses and rainbows. An aircraft with an aft CG is less stable.

While the aircraft is more responsive to control inputs (more maneuverable), this comes
with a trade-off. Recovering from a stall or spin is more challenging due to reduced
stability.

Generally speaking, the trade-off between stability and efficiency makes an aft CG far
more preferable to a forward one. The gain in fuel efficiency is significant, while the
negative stall characteristics and reduced stability are easily managed.

As with a forward CG, understanding the pros and cons of an aft CG is key to managing
your aircraft effectively.

Managing Your CG

Knowing the effects of forward and aft CG is one thing. But how do you manage it?

First and foremost, you need to understand your aircraft’s CG limits. These are outlined
in the aircraft’s Pilot Operating Handbook (POH). Staying within these limits ensures
your aircraft remains controllable and safe.

Next, pay attention to loading. Where you place cargo, how you distribute passenger
weight, and even how fuel is burned during flight can shift the CG. Careful planning
before each flight can help keep the CG within limits.
Remember, a forward CG can result in higher stall speeds and increased fuel
consumption. So if you’re planning a flight with no need for extra stability, it might be
worth managing your load to shift the CG slightly aft for better efficiency. But always
stay within the limits!

As the fuel burns off, your CG can shift. This isn’t a major concern for smaller aircraft
like the C172, but you should calculate your takeoff and landing CG to ensure you stay
within limits for the duration of the flight.
Remember, managing your CG isn’t just about performance – it’s about safety. So
whether you’re gearing up for your first solo or prepping for a cross-country adventure,
make CG management a part of your pre-flight routine. You’ll not only fly better – you’ll
fly safer.

Conclusion
Understanding CG isn’t just about passing your exams, it’s about ensuring each flight
you take is safe and efficient. Get to know your aircraft, be mindful of loading, and
always check your weight and balance before departure!

TAKE OFF-CLIMB SEGMENTS


“CLEARED FOR TAKE OFF”- when a pilot hears these words from the Air Traffic
Controller, he knows it is time for him to bring the highest order of focus he can for one of
the most critical phases of flight , that is ,take off.
Need for Take off Segments

One of the significant emergencies that an aircraft can face is an engine failure on take off ,
it means that the engine is no longer providing the necessary thrust needed.In order for the
aircraft to be safe from any emergencies that may occur , it is necessary that it achieves the
minimum climb gradients and clears its surrounding area and obstacles with sufficient
altitude. Hence the take off segments help in achieving the above requirements.

SCREEN HEIGHT:The take off part of the flight is the distance from where the brakes are
released to the point at which the aircraft reaches a defined height.This defined height is known as
screen height.It is usually 35 ft (for class A aircraft) on a dry runway and if the runway is wet it can
reduce down to 15 ft.
TAKE OFF SAFETY SPEED (V2): V2 is the target speed the aircraft should attain prior to or
before reaching the screen height. The reason it is called take off safety speed is because it should
be attained with one engine inoperative and avoid the aircraft from stalling or the pilots loosing
control of the aircraft.
NOTE:WE WILL DISCUSS THE DIFFERENT AIRCRAFT SPEEDS IN A SEPARATE POST.

CLIMB GRADIENT:The ratio of change in height (altitude gained), during a portion of a climb,
to the horizontal distance traversed in the same time period.It is expressed as a percentage. It also
refers to the angle at which the aircraft climbs.For an aircraft to climb , thrust has to balance drag
and a part of the weight as well . Hence we require excess thrust to give us the climb gradient or
angle of climb .

Climb gradient =Excess Thrust *100/Weight

The take off climb segments start from the screen height that is 35 ft above the take off
surface and end at 1500 ft above the take off surface and are divided into 4 segments.

Segment 1
 The first segment starts when the aircraft reaches the screen height, that is 35 ft .
 The aircraft keeps climbing at the take off safety speed, that is V2 speed, until the gear
is retracted.
 The objective of this segment is to expedite the climb and to make sure there is
reduction in drag .
 There are two ways to reduce drag in this scenario , retracting the flaps or the landing
gear.
 Since retracting the flaps very close to the ground is dangerous, we choose the option
of retracting the gear .
 The first segment ends as soon as the landing gear is retracted .

Segment 2
 The second segment commences from the gear retraction point and the aircraft still
has to maintain the take off safety speed (V2).
 The next important step is to retract flaps so that we can start accelerating the aircraft .
 However , companies have a set altitude from which the flap retraction can start for
example , 400 ft AAL (above aerodrome level).
 As the gear is already retracted in the previous segment, the main source of drag is
removed.
 This makes it easier for the aircraft to climb at a higher climb gradient than segment
1. The climb gradient should not be less than 2.4%.
 The main objective of the second segment is to clear the aircraft from the surrounding
obstacles by maintaining the necessary climb gradient .
 The second segment concludes at 400 ft AGL or the height decided by the company
from where flap retraction can commence .

Segment 3
 The third segment begins from 400 ft or flap retraction altitude set by the company .
 The main objective of this segment is to accelerate the aircraft so that the flaps can be
retracted step by step .
 The reason we accelerate while retracting the flaps is because the stall speed will
increase when we retract flaps .
 Hence as the aircraft accelerates from take off safety speed (V2) to a higher speed that
is minimum drag speed or best angle of climb speed so that the aircraft doesn’t get
close to stall speed .
 Once the flaps are retracted , we can set the thrust levers to maximum continuous
thrust (MCT) from take off thrust .
 In the Airbus 320, take off thrust can only be used continuously for a period of 05 mins
on both engines and in case of an engine failure , it can be used continuously for 10
mins.
 The segment however ends once the thrust lever are set to MCT and flaps are
retracted .

Segment 4
 The fourth segment starts once the trust levers are set to MCT and flaps are retracted.
 The climb gradient for the last stage should not be less that 1.2%.
 In the fourth segment, the airplane is climbed to above 1500 ft AAL (above aerodrome
level) where the take off flight path ends.
 In case of an engine failure , the pilot can make a decision to land at the departure
aerodrome or go on further to an alternate airfield.

Fire Extinguishing Agents


There are different classifications of fires differentiated by the type of material that is burning. In general terms,
there are four primary fire types:

1. Class A - Ordinary combustibles (solid material fires) - wood, paper, plastic etc,
2. Class B - Flammable liquids or gases - fuels, alcohol, aerosols,
3. Class C - Electric fires,
4. Class D - Combustible metal fires - magnesium, potassium etc.

Fires of all classifications cannot be dealt with in the same way. As a consequence, the airport rescue and fire
fighting services (RFFS) must have a variety of fire suppression tools at their disposal. This article identifies the
three primary categories of fire extinguishing agents currently used by the RFFS in dealing with aircraft fires.

These categories are:


1. Primary Agents
2. Supplementary Agents
3. Other Agents

Extended Range Operations


EXTENDED DIVERSION TIME OPERATIONS (EDTO)
The Regulatory Context

ICAO Requirements for Extended Range Twin-engine Operations (ETOPS) have been in place since 1985,
when they were introduced to apply an overall level of operational safety for twin-engined aeroplanes which
was consistent with that of the modern three and four-engined aeroplanes then flying, to which no restrictions
were applied. As aeroplane reliability and range improved, it became clear that all multi turbine-engined aircraft
were pushing the boundaries of flight away from nearby alternates to increasingly distant ones and a review of
the existing arrangements for ETOPS began.
After many years of discussion about how to broaden the facilitation of international flights for all large
transport aeroplanes which necessitated tracks with no close-by diversion aerodromes (or could be more
efficiently routed with the use of these tracks), led in 2012 to changes to ICAO Annex 6 Part 1 under
Amendment 36. This introduced the Extended Diversion Time Operations (EDTO) regime in place of
ETOPS. However since then, although the EDTO regime has been widely accepted, the term EDTO has not
been universally adopted the continued use of ETOPS is explicitly allowed for in Annex 6 provided that EDTO
concepts "are correctly embodied in the concerned regulation or documentation". Given this flexibility, the term
'ETOPS' has been retained by the FAA and others by redefining it as an abbreviation for 'ExTended range
OPerationS' rather than as previously 'Extended range Twin OPerationS'. EASA currently continues to use
ETOPS as originally defined and the abbreviation 'LROPS' (Long Range OPerationS) for extended range
operation by three and four-engined aircraft.

The new ICAO guidance

Annex 6, and particularly Attachment D to that Annex, now contains guidance on extended range operations for
all turbine-engined aeroplanes which are conducted beyond 60 minutes from a point where it is possible to fly
to an en-route alternate aerodrome. The main change is that a distinction is drawn between such operations
which do not exceed an established 'Threshold Time' defined as "the range, expressed in time, established by
the State of the Operator to an en-route alternate aerodrome, whereby any time beyond it requires an EDTO
approval from that State". ICAO uses the flying time at the one engine out speed in ISA and Still Air to convert
Threshold Time to distance for aeroplanes with two engines but the all engines operating speed in ISA and Still
Air for the same conversion for aeroplanes with more than two engines. However, note that the FAA uses the
one engine out speed as the basis for all aeroplane type EDTO approvals.
Non-EDTO flights are expected, without any detailed specification, to be subject to flight planning principles
which are additional to those for 'normal' operations in respect of:
 Operational control and dispatch procedures
 The identification of alternate aerodromes
 The provision of comprehensive and current information on aerodromes to be used as alternates
 The assurance to the extent possible for twin-engined aeroplanes only that an aerodrome will be available if its
use as an alternate becomes necessary
 The inclusion of appropriate content in relevant personnel training programmes.
EDTO flights are subject to a process of explicit approval which, as with the former Extended Range Twin
Operations system, has both aeroplane type design and aeroplane operational requirements. System Safety in
EDTO Extended Range Operations is based, as it was in the case of the original ETOPS concept, on two
fundamental principles:
 Precluding the need for an en-route diversion by designing reliable aeroplane engines and critical systems and
then implementing specific maintenance precautions prior to dispatch.
 Protecting any en-route diversion which might become necessary by implementing, at the design level, system
reliability which is able to facilitate a safe diversion and then having operational systems in place which will
ensure that aeroplane management during a diversion secures the completion of such a diversion.
These principles are satisfied as previously by:

 An Aeroplane Type Design Assessment of the airframe/engine combination to ensure that the likely continued
airworthiness of an aircraft used in such operation is compatible with the extended range operations which it is
permitted to fly. This must be approved/validated by the Type Design Authority.
 An Operational Safety Approval for each airline which wishes to use aeroplanes they operate on EDTO to
ensure that the way in which those operations are conducted is appropriate. This is issued by the State of the
Operator.
Although the first of these principles is founded on the reliability of the aeroplane engines, other considerations
such as cargo compartment fire suppression and/or containment capability are also important as are the
reliability / redundancy associated with systems other than those directly related to propulsion such as the
integrity / reliability of the pressurisation system. The second principle requires the existence of an operational
process which manages the risk involved in EDTO so that it is no greater than the risk of any other flight.

Maximum Diversion Time under EDTO

Each EDTO Type Design or Operational Approval introduces the concept of Maximum Diversion Time so that
an approved authorised area of operations can be defined. Type Design requirements have been amplified so
that when the Maximum Diversion Time exceeds 180 minutes at one engine out speed in ISA and Still Air, the
world fleet In-Flight Shut Down (IFSD) rate for the specific airframe/powerplant combination must be less than
0.01/1000 hours and the capability of time-limited systems, in particular the cargo fire suppression system
relative to a maximum diversion at the All Engines Operating speed.

EDTO Type Design Assessment


This must include:
 demonstration that the airframe/engine combination is designed to fail-safe (CS-25/FAR Part 25) criteria and
documentation of the effects of operation with a failed engine.
 validation of an appropriate fuel management system, independent AC electrical power sources and satisfactory
cargo fire protection and equipment cooling systems.
 an analysis of the consequences of system/component failure.
 the identification of time-limited systems - typically that for cargo fire suppression.
 the in-service experience (world fleet)
 a manufacturer validation flight test
The submission to the Type Design Authority must be in the form of an 'EDTO-Configuration, Maintenance
and Procedures (CMP) Document' which must contain "the particular aeroplane configuration minimum
requirements, including any special inspection, hardware life limits, MEL constraints and maintenance
practices found necessary to establish the suitability of an airframe-engine combination for EDTO". Whilst
each derivative of an aeroplane type variant requires explicit approval, Type Design Approval does not depend
on in-service experience and recently-introduced aeroplane types have achieved EDTO qualification prior to
initial entry into service.
The EDTO Operational Approval Process

An applicant for EDTO Operational Approval must apply in the manner specified by the State of the Operator.
The type of approval process which is then followed will depend on whether the applicant has direct in-service
experience with the candidate aircraft, for example one year of non-EDTO operations for a 120 minute twin
engine approval or one year of 120 minutes EDTO operations for a 180 minute twin engine approval, or not.
With such experience, the 'In-service' (formerly called 'Conventional)' track is followed to develop a simplified
Approval Plan but if an applicant is unable to demonstrate such prior experience then the longer 'Accelerated
EDTO' approval track will apply.
Any applicant will be expected to:
 Define the EDTO routes that will be covered by application.
 For the aeroplane type to be used, establish EDTO maintenance procedures (including but not limited to
appropriate engine condition, IFSD and oil consumption monitoring and validation of APU start reliability).
 For each proposed route, establish a list of adequate en-route alternate airports (requirements include but are not
limited to the availability of at least one approach aid and a minimum [and Fire Fighting Category] of 4.
 Determine the EDTO diversion time required and the EDTO engine inoperative planning speed.
 Establish the required EDTO Area of Operations.
 Establish a system for obtaining EDTO flight plan data i.e. EDTO en-route alternates, equal time points, the
critical fuel scenario and time limited systems.
 Make arrangements to obtain weather data for EDTO en-route alternates.
 Ensure there is a reliable method of communication between the aeroplane and the airline during the flight.
 Review the EDTO provisions in the approved MMEL in order to establish the airline MEL for EDTO.
 Establish a method of checking the in-flight start reliability of the APU.
 Designate an EDTO Check Pilot.
 Establish and fully document airline operating procedures for EDTO in EDTO Procedures Manuals for both
Flight Operations and Maintenance processes.
 Expand normal flight crew and aeroplane maintenance guidance material to include EDTO practices and
procedures.
 Establish appropriate EDTO training procedures for line maintenance and flight operations ground staff and for
flight crew.
Before issuing an EDTO approval, a State must normally ensure that the maximum diversion time for the
operator of a particular aeroplane type does not exceed the most limiting EDTO-significant-system time
limitation which is relevant to that particular operation and identified in the AFM (directly or by reference) is
not exceeded and, in the case of aeroplanes with two turbine engines only, that the aeroplane type involved is
EDTO certified. However, if a specific safety risk assessment conducted by the aeroplane operator has
demonstrated to the satisfaction of the State that an equivalent level of safety can be maintained during
operations beyond the time limits of the most time-limited system, then there is discretion to approve such an
operation. The specific safety risk assessment required to allow this exceptional EDTO approval to be granted
must:
 describe the capabilities of the operator
 provide data which demonstrates the overall reliability of the aeroplane
 assess the reliability of each time-limited system
 provide relevant information from the aeroplane manufacturer
 detail any specific mitigation measures

A representation of the EDTO regulatory regime (Source ICAO)


Any EDTO approval issued by a State must also reference the need for compliance with the EDTO critical fuel
scenario established by the State and any flight conducted under an EDTO approval must not continue beyond
the specified Threshold Time until the identified alternates have been re-evaluated for their continued
availability and current information indicates that, during the period of time when their use might be required,
the prevailing conditions at those aerodromes will not be less than the applicable aerodrome operating minima.
If this is not the case, then an alternative course of action, which does not involve exceeding the applicable
Threshold Time must be determined.
Approvals for aeroplanes with two turbine engines are also required to take into account the following when
determining if such an approval will provide the overall level of safety for any flight intended by the provisions
of ICAO Annex 8 in respect of:
 the reliability of the propulsion system
 the airworthiness certification of the aeroplane type
 the EDTO-specific maintenance programme
Continuing Surveillance of EDTO-relevant operating experience

The continued applicability of EDTO approvals depends on the reporting of any EDTO Relevant Event -
defined as "any system malfunction, degradation or other in-flight event that requires the crew to make a
decision whether to turn back, divert or continue under an increased level of alertness". It is understood that the
majority of these events occur because of non-technical circumstances such as adverse weather, passenger
medical issues, flight or ground personnel error and bird strikes. However, awareness of technical events allows
oversight of continued airworthiness by both aeroplane manufacturers and their primary certifying authorities so
that any necessary corrective actions can be and are promptly identified and implemented.
ETOPS Grandfather Rights

ICAO SARPs include a specific Recommendation that State ETOPS approvals issued prior to 25 March 1986
which permitted operations beyond the Threshold Time since established for such operations "should give
consideration to permitting such an operation to continue on that route after that date"

Affect of Temp, Humidity, Altitude etc on Aircraft Performance


The higher the pressure, the lower the temperature, and the lower the humidity,
the denser the air. The density of the air affects the performance of the aircraft in
several ways. The denser the air, the more lift and drag the wings generate, the
more thrust the engine produces, and the more fuel the engine consumes
Temperature also has a tangible impact on aircraft performance in the air. Cold
air is denser, providing more lift and making the engines more efficient, improving
fuel consumption. Conversely, in hot weather, the air is less dense, reducing lift
and engine efficiency.
Humidity decreases the performance of most aircraft, not only because of it's
effect on the wings, but also the effect on the engines. Humidity has a major
affect on the way planes fly. This is due to the weight of the air when it is humid.
Whether due to high altitude, high temperature, or both, reduced air density
(reported in terms of density altitude) adversely affects aerodynamic performance
and decreases the engine's horsepower output. Takeoff distance, power
available (in normally aspirated engines), and climb rate are all adversely
affected.
Whether due to high altitude, high temperature, or both, reduced air density
(reported in terms of density altitude) adversely affects aerodynamic performance
and decreases the engine's horsepower output. Takeoff distance, power
available (in normally aspirated engines), and climb rate are all adversely
affected.

EFFECT OF WIND ON AIRCRAFT


A steady (continuous) head wind (or tail wind) will not affect your climb rate, only your climb angle. It
means you will reach a specific altitude in the same time interval, but your ground distance will be
affected.

A head wind increase (as in a gust) will momentary increase your indicated airspeed, which you can trade
for a (momentary) increase of climb rate. Similarly, a tail wind gust will temporarily decrease your indicated
airspeed and you might have to pitch the nose down a little, to maintain airspeed and therefore reducing
climb rate. These are only transitory effects, until the plane settles back to its original trim speed.

HEADWIND
When the wind blows against the direction in which the aircraft moves it is called headwind.
Effects
In aerodynamics, flying against the wind at a certain speed is equivallent to flying in calm air at an increased
airspeed and reduced groundspeed.

Impact on Operations

Headwinds impact all phases of the flight:

 During take off and landing, headwind increases the airflow, hence the necessary lift is achieved earlier
and at lower speeds (the wind speed is added to the aircraft speed). As a result, less runway is required
to perform a safe take off or landing. Therefore, headwind is more favourable for these phases of the
flight.
 A climb or descent in headwind conditions results in an increase of the gradient (i.e. the level change
over the distance travelled) even though the rate of climb or descent (i.e. the level change over time)
remains the same. As a result, aircraft reach their cleared levels earlier compared to the calm wind
scenario.
 In the cruise phase, having headwind means increased fuel burn (due to the increased drag) and
increased flight time (due to the reduction in groundspeed). Operators try to avoid flying in headwind
conditions whenever practicable by e.g. choosing an alternative route (which can be a bit longer but
with less or no headwind) or choosing a different (normally lower) cruising level if the predicted wind
speed is more favorable.

1. Why Are Headwinds Good For Takeoff And Landing?


During takeoff headwinds help to increase lift, meaning a lower ground speed and a shorter runway
distance is needed for the plane to get airborne. Landing into the wind has similar advantages; less
runway is needed and ground speed is lower at touchdown.

Landing in crosswinds and tailwinds make takeoff and landing more challenging and at times can
mean it is not possible to attempt either. Every aircraft has set limits for takeoff and landing in a cross
wind and is not permitted to operate if conditions exceed those limits.

2. Are Wind And Weather Conditions Considered Pre-Flight?


Flight planning is an important part of any flight. Every pilot takes into consideration weather
conditions and winds when choosing a suitable flight path or when planning a flight.

Tailwinds are useful to travel faster and save on fuel as less power is needed to drive the aircraft in
the direction it needs to go. Headwinds mean the exact opposite; more fuel is needed and the flight
will take more time. A crosswind pushing to the side can cause you to drift off course, so the plane’s
direction must be adjusted to keep it moving along the right path despite the wind.

3. Do Strong Winds Affect Flights?


Gusts of wind that change direction quickly and abruptly are one of the most dangerous wind
conditions both in flight and on takeoff and landing. A sudden change in headwind or tailwind causing
rapid changes in lift to the aircraft also known as ‘Wind Shear’ is one of the worst wind effects to
experience, however pilots are trained to deal with these kinds of situations.

Difficult wind conditions can be an influencing factor in some aviation accidents, however there are
almost always other risk factors involved. Pilots are well trained in controlling aircraft during windy
conditions and understand the limitations of their aircraft. Each individual aircraft and different runway
has its own pre-defined wind limits for both wet and dry conditions, but the pilot has the final decision
on whether it is safe to takeoff or land, and fly in the existing conditions.
4. How Does The Pilot Keep Track Of The Wind Effect During
Flight?
Pilots use a range of instruments to keep track of winds and atmospheric conditions while in-
flight. Airspeed Indicators and other instruments are used to give the pilot the information they need in
order to compensate for the changes in wind during flight.
Altimeters measure air pressure and altitude, Attitude and Heading Indicators are used to show the
aircraft’s orientation and direction, and the Vertical Speed Indicator shows the rate of climb or
descent.

Density Altitude:
What It Is, And How It Affects Your Performance
Density altitude is a measure of how 'thick' the air is, and it's based on three
factors: atmospheric pressure, temperature, and humidity.

The technical definition of density altitude is "pressure altitude, adjusted for non-standard
temperature." What that really means is on hot days, the air is much 'thinner', or less dense,
than it is on cold days.

Why does that matter? It's a big factor in your airplane's performance, because when the air
surrounding your plane is less dense, it means your wings, propeller, and engine will have a
lot less performance, and it will take you more time to get airborne during takeoff.
1) It starts with pressure.

The first factor in density altitude is pressure, or more specifically, atmospheric pressure. The
lower the pressure, the fewer air molecules surround your airplane.

In fact, decreasing atmospheric pressure by one inch of Mercury (inches Hg) increases your
pressure and density altitudes by 1,000 feet. Your airplane performs like it's 1,000 feet higher
than the field elevation. So if your airport's field elevation is 1,500' MSL, your plane is going to
perform like it's actually at 2,500' MSL.

2) Next, add temperature.

Temperature is the single biggest factor in density altitude. That's because when you
heat air, the air molecules have more energy, and they spread further apart, making the air
less dense.

The effects of temperature are eye-opening. Take Denver, CO (5,434' field elevation) for
example, where the average July temperature is 31 degrees C. That temp increases Denver's
density altitude by 3,012', to a total of 8,446' density altitude.

How well does your plane perform at 8,446' MSL? How much runway do you think you'd need
for takeoff at 8,446'?
Boldmethod

3) Finally, add in humidity.

Humidity has the smallest effect on density altitude, but it can make a difference of several
hundred feet. Water vapor weighs less than the nitrogen and oxygen that make up most of
the atmosphere. When the humidity is high, the air is less dense.

Humidity is complex to compute, but there's a great calculator for it here.

4) The result? A major performance penalty.

What this all comes down to is a major performance penalty for your plane on hot days, and
when the atmospheric pressure is low.

Look at the difference in takeoff distance on a hot day in Denver versus a cold one; takeoff
roll is increased by 30%. And clearing a 50' obstacle? It's an increase of 32% Those
numbers can make a big difference, especially on a short runway.
What Does It Mean For You?

Density altitude is something you always need to consider, especially when your airplane is
heavy, you're at high altitude, and it's warm outside. Use your POH to calculate your takeoff
distance, and make sure you have enough runway for a safe takeoff.

How much extra runway should you have for takeoff? It's often recommended to add 50% to
your takeoff performance calculations. That gives you plenty of extra room for takeoff, no
matter what the weather is doing.

You might also like