You are on page 1of 164

THIS PAGE IS INTENTIONALLY BLANK

ii General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
PRODUCED & PRINTED BY BCFT

Please note that the information contained in these notes is for instructional use only. Every effort
has been made to ensure the content is valid, accurate and complete. No responsibility is accepted
for errors or discrepancies. The text is subject to regular change without notice.

©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011.


All intellectual property rights including copyright in the content of this manual are
owned and controlled for the purposes of BCFT. They may only be used for your
own personal non-commercial uses. You are not permitted to copy, broadcast,
download, store (in any medium), transmit, show or play in public, adapt or change in
any way the content of this manual for any purpose whatsoever without the prior
written permission of BCFT.

These notes are designed for use during BCFT Modular ATPL (A) courses.
The notes are also suitable for distant learning with appropriate
instructor guidance and worksheets.
The layout and order of the notes follows a logical learning sequence and is based upon the structured
JAR/EASA ATPL (A) learning objectives 2008 (NPA25)

General Navigation (NPA25) - Edition 4 – 110630 iii


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

iv General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
SECTION ONE THE EARTH .............................. 1-1
Chapter 1 Form of the Earth ............................................... 1-1
1 Shape of the Earth .............................................................................................................1-1
2 The North and South Poles, East and West ......................................................................1-2
3 Great and Small Circles ....................................................................................................1-2
4 Meridians and Parallels .....................................................................................................1-4

Chapter 2 Defining Direction on the Earth – Part 1 .......... 2-1


1 Angular Definition of Direction ........................................................................................2-1
2 True Direction ...................................................................................................................2-1
3 Magnetic Direction ...........................................................................................................2-2
4 Dip .....................................................................................................................................2-2
5 Variation............................................................................................................................2-3
6 Changes in Variation .........................................................................................................2-3
7 Definitions .........................................................................................................................2-4
8 Converting Between Magnetic and True Datums .............................................................2-4
9 Isogonals ...........................................................................................................................2-5
10 Compass Direction and Deviation ....................................................................................2-5
11 Summary of Changes of Direction Reference ..................................................................2-6

Chapter 3 Defining Position on the Earth ........................... 3-1


1 Position Reference Systems ..............................................................................................3-1
2 Latitude & Longitude ........................................................................................................3-1

Chapter 4 Measuring Distance on the Earth ...................... 4-1


1 Change of Latitude/Difference in Latitude .......................................................................4-1
2 Change of Longitude/Difference in Longitude .................................................................4-2
3 Distance Measurement ......................................................................................................4-3
4 Ch. Lat to Nautical Mile Conversion ................................................................................4-4
5 Departure (Earth Distance) ...............................................................................................4-6

Chapter 5 Meridian Convergence and Conversion Angle . 5-1


1 Meridian Convergence ......................................................................................................5-1
2 Definition of Convergence: ...............................................................................................5-1
3 Importance of Convergency ..............................................................................................5-2
4 Great Circles and Rhumb Lines ........................................................................................5-2
5 Conversion Angle .............................................................................................................5-4

General Navigation (NPA25) - Edition 4 – 110630 v


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
SECTION TWO BASIC NAVIGATION .............. 5-1
Chapter 6 The Navigation Computer (Part 1) .................... 6-1
1 Definitions .........................................................................................................................6-1
2 The Circular Slide Rule.....................................................................................................6-1

Chapter 7 The Triangle of Velocities................................... 7-1


1 Velocities ..........................................................................................................................7-1
2 Vectors ..............................................................................................................................7-1
3 The Vector Triangle ..........................................................................................................7-1
4 Examples ...........................................................................................................................7-3
5 Directional Datums ...........................................................................................................7-5

Chapter 8 The Navigation Computer (Part 2) .................... 8-1


1 The Slide ...........................................................................................................................8-1
2 Use of the Computer .........................................................................................................8-1

Chapter 9 Pilot Navigation Techniques (Part 1) ................ 9-1


1 The ‘One in Sixty Rule’. ...................................................................................................9-1
2 Tracking Errors and Corrections .......................................................................................9-2
3 Other Common Applications of the 1 in 60 Rule .............................................................9-4
4 Navigation in climb and descent .......................................................................................9-7

Pilot Navigation Techniques (Part 2) ..................................... 9-9


5 Topographical Aviation Mapping .....................................................................................9-9
6 Topographical Features ...................................................................................................9-10
7 Reading the Map .............................................................................................................9-11
8 Visual Check Points. .......................................................................................................9-12
9 Lost Procedure. ...............................................................................................................9-12

SECTION THREE CHARTS AND PLOTTING...... 9-1


Chapter 10 A Flat Earth? .................................................... 10-1
1 Chart Projections. ............................................................................................................10-1
2 Scale Factor. ....................................................................................................................10-1
3 Types of Projection. ........................................................................................................10-1
4 Plane Projections. ............................................................................................................10-1
5 Conic Projections ............................................................................................................10-3
6 Cylindrical Projections ....................................................................................................10-4
7 Chart Properties...............................................................................................................10-5

vi General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 11 The Mercator Projection.................................. 11-1
1 Description ......................................................................................................................11-1

Chapter 12 Other Mercator Charts .................................... 12-1


1 The Transverse Mercator. ...............................................................................................12-1
2 The Oblique Mercator .....................................................................................................12-2

Chapter 13 Navigation – Plotting ........................................ 13-1


1 The Manual Track Plot ...................................................................................................13-1
2 The Track Plot – Plotting Exercise. ................................................................................13-2

Chapter 14 Plotting Using Position Lines. .......................... 14-1


1 The Radio Magnetic Indicator (RMI) .............................................................................14-1
2 Plotting the Bearings .......................................................................................................14-2
3 Transferring Position Lines. ............................................................................................14-3
4 Construction of a Fix by Transferring Position Lines.....................................................14-5
5 Application of Variation when Plotting. .........................................................................14-5
6 Conversion Angle on Mercator Charts. ..........................................................................14-6
7 Finding Conversion Angle. .............................................................................................14-7
8 Rules for Application of Variation and Conversion Angle. ............................................14-8
9 Relative Bearings. ...........................................................................................................14-8
10 Current Use of Mercator Charts. .....................................................................................14-9

Chapter 15 The Lambert Conformal Conic Projection. ... 15-1


1 Properties of a Lambert’s Conformal Conic Chart. ........................................................15-2
2 Calculations. ....................................................................................................................15-3
3 Advantages and Disadvantages of the Use of Lambert’s Charts. ...................................15-5
4 Plotting of Tracks ............................................................................................................15-6
5 Plotting of Position Lines ................................................................................................15-7
6 Fix from Position Lines...................................................................................................15-9

Chapter 16 The Polar Stereographic Projection. ............... 16-1


1 The Projection. ...............................................................................................................16-1
2 Properties ........................................................................................................................16-1
3 Uses. ................................................................................................................................16-2
4 Calculations .....................................................................................................................16-2

Chapter 17 Grid Navigation. ............................................... 17-1


1 The Navigation Problem. ................................................................................................17-1
2 The Solution. ...................................................................................................................17-1
3 Grid Calculations on a Polar Stereographic Chart ..........................................................17-3

General Navigation (NPA25) - Edition 4 – 110630 vii


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
SECTION FOUR SPACE AND TIME ................. 17-1
Chapter 18 The Solar System .............................................. 18-1
1 Orbits and Kepler’s Laws ...............................................................................................18-1
2 Rotation of Planets ..........................................................................................................18-3
3 The Earth’s Motion .........................................................................................................18-3
4 The Celestial Sphere .......................................................................................................18-6
5 Relative Co-ordinates ......................................................................................................18-8
6 The Ecliptic .....................................................................................................................18-9

Chapter 19 Time ................................................................... 19-1


1 Definitions of Day ...........................................................................................................19-1
2 Local Mean Time ............................................................................................................19-4
3 Standard Time .................................................................................................................19-7
4 The International Date Line ..........................................................................................19-10
5 The Year ........................................................................................................................19-10

Chapter 20 Sunrise and Sunset ........................................... 20-1


1 Variation of Sunrise and Sunset ......................................................................................20-1
2 Sunset and Sunrise Tables...............................................................................................20-3
3 Civil Twilight ..................................................................................................................20-5

SECTION FIVE SPEED TIME DISTANCE


CALCULATIONS ........................... 20-1
Chapter 21 Relative Velocity and Speed Adjustments ...... 21-1
1 Relative Velocity.............................................................................................................21-1
2 Speed Adjustments ..........................................................................................................21-3
3 Summary .........................................................................................................................21-5

Chapter 22 Critical Point / Point of Equal Time................ 22-1

Chapter 23 Point of Safe Return, Point of No Return and


Radius of Action ..................................................................... 23-1

viii General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
SECTION SIX NAVIGATION INSTRUMENTS .. 23-1
Chapter 24 Magnetic Instruments ...................................... 24-1

Chapter 25 Inertial Navigation Systems ............................. 25-1

Chapter 26 Flight Management System (FMS) ................. 26-1

General Navigation (NPA25) - Edition 4 – 110630 ix


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

x General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section One The Earth

Chapter 1 Form of the Earth

1 Shape of the Earth


For most purposes the earth is assumed to be a perfect sphere, but in fact it is slightly flattened at
the poles. This shape is known as an ‘ellipsoid’ or an ‘oblate spheroid.’ The earth’s diameter is
about 27 statute miles (or 23 nautical miles) less pole to pole than at the equator. This flattening,
which is called ‘compression’, is about 0.3% or 1/298.
The diameter measured at the equator is called the ‘major axis’, with the radius at the equator being
the ‘semi-major axis’. When measured pole-to-pole, these are the ‘minor axis’ and the ‘semi-minor
axis’ respectively.

Fig 1-1 An oblate spheroid (exaggerated)

The diameter of the earth measured at the equator (major axis) is 6883.7nm. Given this
measurement and the compression ratio of 1/298, we can calculate the diameter from pole to pole:

Minor axis = Major axis / 298 * 297 = 6860.6nm (~23nm shorter)

Due to this compression it is also interesting to note that if you wanted to travel from a point on the
equator to another point diametrically opposed, or antipodal, then the shortest route would actually
be the great circle route via the pole.

The maximum radius actually occurs somewhat South of the equator, making the earth slightly
pear-shaped. For general navigation purposes the compression factor is small enough to be ignored.

Some other interesting facts to know about the earth:

Circumference = 40,000km or 21,600nm (360 * 60)


Distance from equator to pole = 10,000km
Diameter = Circumference / (where = 3.1415927)
Radius = Diameter / 2

General Navigation (NPA25) - Edition 4 – 110630 1-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2 The North and South Poles, East and West
The earth rotates on its axis and the extremities of that axis are known as the North and South poles.
East is defined as the direction in which the earth is rotating and West is the direction opposite to
East. The North Pole is on the left of an observer who is facing East. An observer will also note
that the sun rises in the East and, if viewed from a position above the North Pole, the Earth would
appear to be rotating in an anti-clockwise direction.

3 Great and Small Circles

3.1 Great circle - definition

A great circle on the surface of the earth is one in which the centre and radius of the circle are those
of the earth itself. It is the largest radius circle which can be drawn on the earth’s surface.

Figure 1-2 The Great Circle

The shortest distance between two points on the earth is always the shorter arc of the great circle
passing through them both. Only one great circle passes through any two points on the earth’s
surface unless they are diametrically opposite to each other, in which case an infinite number can be
drawn (in theory).

The equator is a great circle which has its plane perpendicular to the earth’s axis of rotation. It
divides the earth into Northern and Southern hemispheres, and every point on the equator is
equidistant from the two poles.

Each meridian of longitude, together with its corresponding anti-meridian, form a great circle.

1-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3.2 Great circle - vertices

Any great circle will cross the equator at two points (except the equator itself) and will reach a most
northerly and most southerly latitude. The most northerly and most southerly points are known as
the north vertex and the south vertex respectively. The vertices are defined by the angle at which the
great circle crosses the equator, measure from the equator itself.
For example, a great circle crossing the equator at an angle of 40 degrees will have its north vertex
at 40 degrees north and its south vertex at 40 degrees south.

Figure 1-3 Vertices of a great circle

3.3 Small circle - definition

Any circle on the earth’s surface which does not have the earth’s centre as its centre is known as a
small circle.

Figure 1-4 The Small Circle

General Navigation (NPA25) - Edition 4 – 110630 1-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4 Meridians and Parallels
Position is defined on the earth by a wide variety of methods. The most common of these is by
using Meridians of Longitude and Parallels of Latitude. This will be covered in more detail later,
but at this stage the following definitions are required:

4.1 Meridians of Longitude.

Meridians are semi-great circles or 180-degree arcs of a great circle and each one joins the poles.
All meridians therefore indicate North-South directions and every great circle passing through the
poles forms a meridian and its anti-meridian.

4.2 Parallels of Latitude

A “parallel” of latitude is a small circle drawn on the earth joining all points that are at the same
latitude. Any such line is parallel to any other such line, hence the name. The equator is the only
parallel of latitude which is not a small circle but is a great circle. It joins all points on the earth that
are at 0 degrees N/S.

Figure 1-5 Meridians and Parallels

Lines of Latitude and Longitude are drawn on what is known as a ‘Reduced Earth’, forming a
‘Graticule’. The earth’s compression factor is ignored.

1-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 2 Defining Direction on the Earth – Part 1
In order to fly in any particular direction over the earth, it is necessary to define that direction by
using a fixed reference of some kind. No less than four types of direction are necessary in air
navigation, but in this chapter we will begin with just two of them, True direction and Magnetic
direction. The Sexagesimal system (degrees) is used for angular measurement.

1 Angular Definition of Direction


By convention in air navigation, directions are always measured clockwise from whatever datum is
in use at the time. Three-figure groups are used to avoid confusion, and a suffix is added to show
which datum is being used. For example, an aircraft steering in a direction 30° to the right of
magnetic North would show its direction as 030°(M). An aircraft flying in a direction 30° left of
true North is said to be flying 330°(T).

The four directions North (000° or 360°), South (180°), East (090°) and West (270°) are known as
the Cardinal Points, and the intermediate directions at 45° to them are known as the Quadrantal
Points.

Figure 2-1 Defining Direction

2 True Direction
The most convenient method of defining direction is to use the local meridian, the one which passes
through the aircraft’s present position. This defines the local North/South line between the poles
which form the earth’s axis of rotation and it has the distinct advantage of remaining constant at all
times, whereas all three of the other directional datums vary in some way. The meridians are
therefore said to define the True North datum and this is extensively used in air navigation.

General Navigation (NPA25) - Edition 4 – 110630 2-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3 Magnetic Direction
The earth behaves like an enormous magnet, having a magnetic field which is strong enough to
influence the position of any magnetic needle which is freely suspended, whatever its position on
the earth. The field varies in the angle it makes with the earth’s surface, having both a horizontal
and vertical component. The horizontal component is strongest at the magnetic equator, and the
vertical component there is non-existent. The vertical component, however, is at its strongest over
the magnetic poles, whereas the horizontal component over both of the magnetic poles is zero.

Figure 2-2 Terrestrial Magnetism

In equatorial latitudes a suspended magnet, such as that in a simple child’s compass, will align itself
along the lines of force of the earth’s field in a roughly North/South direction, with the North-
seeking end of the magnet pointing towards the North magnetic pole. The alignment of the earth’s
magnetic field thus defines Magnetic direction, and the lines of force are known as Magnetic
Meridians. A compass needle, which is designed to be sensitive to the horizontal component of the
Earth’s magnetic field (known as the Directive Force), will therefore align itself along a magnetic
meridian.

4 Dip
As can be seen in the diagram above, the horizontal component of the Earth’s field is parallel to the
Earth’s surface at the Magnetic Equator, but at all other magnetic latitudes there will be an angle
between the field and the local horizontal. This angle is known as ‘Dip’.

Dip varies with magnetic latitude. At low magnetic latitudes, i.e. near the magnetic equator which is
equidistant from the two poles dip is very low. At the magnetic equator itself dip is approximately
zero (local field anomalies may mean it is not quite zero), and this is known as the aclinic line. At
high magnetic latitude, near the magnetic poles dip approaches 90°.

2-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
As dip increases the horizontal component, the H component, of the magnetic field decreases. This
component is known as the directive force because it is directing the compass needle. If this drops
below 6 micro Tessla (6 T) then it will no longer reliably direct the compass. The regions in which
the directive force is too weak for the compass to be used are therefore known as the 6-micro-Tessla
zones. They are areas around each magnetic pole.

Lines of joining points of equal dip are known as isoclines.

5 Variation
The North magnetic pole is situated in Northern Canada near Hudson Bay, and the South magnetic
pole is in Antarctica in South Victoria Land. Since they are some considerable distance away from
the true poles, there will usually be a difference between true direction and magnetic direction. The
angle between magnetic and true direction (i.e., the magnetic and true meridians), at any point is
known as Variation.

Variation at any point is the angular difference between True and Magnetic direction at that
point measured East or West of True North.

Figure 2-3 Variation

An accurate knowledge of the variation at any point enables the navigator to convert from True
direction to Magnetic direction and vice-versa.

6 Changes in Variation
Since the magnetic poles are constantly moving, however, Variation is not fixed but changes very
slowly with time. The variation in the UK, for instance, has changed by about 2° in the past 30
years, and over a period of about 960 years the Magnetic poles will rotate completely around the
True poles.

Short-term variation changes also occur. An annual change takes place, which is due to the position
of the earth in its orbit. Both the rotation of the earth and sun’s effect on the ionisation of the

General Navigation (NPA25) - Edition 4 – 110630 2-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
atmosphere cause daily (diurnal or day/night) changes in variation, and magnetic storms can cause
significant alterations.

7 Definitions
Definitions required in this and subsequent chapters are:

a. Heading. (Hdg). The heading of an aircraft is the direction in which the aircraft is steered,
measured in degrees clockwise from a specified datum. It may be measured from True North,
Magnetic North, or Grid North datums. The grid datum is covered in a later chapter.

b. Track. (Trk). The track of an aircraft is the direction taken by the path of the aircraft over the
ground, measured in degrees clockwise from True North, Magnetic North, or Grid North
datums. It is often known as ‘Course’ (Crs) in parts of Europe and the USA.

8 Converting Between Magnetic and True Datums


Whenever variation is known, it may be used to convert from Magnetic direction to True and vice-
versa.

Figure 2-4 East and West Variation

Note that where the variation is Westerly, the Magnetic heading is greater numerically than the True
heading. This direction of conversion can easily be confused, so it is best to remember a
mnemonic:

Variation West, Magnetic best. Variation East, Magnetic least

Example 1: Aircraft heading 227°(M), variation 7°E, what is the true heading?

Variation is East, so the magnetic heading is numerically less than the true heading. Therefore add
the variation of 7° to the magnetic heading to find the true heading.

227° + 7° = 234°°(T)

Example 2: Aircraft heading 358°(T), variation 19°W, what is the Magnetic heading?

2-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Variation is West, so magnetic is ‘best’, i.e. greater than true. We therefore add the variation to the
true heading to find the magnetic heading.

358° + 19° = 017°°(M)

9 Isogonals
Since aircraft frequently use magnetic direction whilst flying, it is frequently necessary to make the
conversions between true and magnetic. The variation is therefore printed on navigation charts as
‘Isogonals’, which are defined as lines which join places of equal variation. The date at which the
printed variation was correct is usually printed next to the variation figure, together with a statement
of the annual amount and direction of movement so that the chart can be used when the variation
marking is a little out of date.

Where the variation is zero on the chart, the isogonal marking those positions is known as the
Agonic line

Figure 2-5 Isogonals

10 Compass Direction and Deviation


Due to magnetic fields that exist around all aircraft, and are discussed in more detail in the
Instruments syllabus, a magnetic compass will generally not align exactly towards magnetic north.
A bearing as shown on an aircraft’s compass is known as a Compass Bearing, and the difference
between compass bearing and magnetic bearing is known as deviation. Deviation is found from the
compass card for the aircraft, and varies depending on the aircraft’s current heading.

General Navigation (NPA25) - Edition 4 – 110630 2-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Deviation is applied to magnetic bearing to find compass bearing in the same way that variation is
applied to a true bearing to find magnetic, and in the same sense. The mnemonic to be remembered
for this runs:

Deviation West, Compass best. Deviation East, Compass least

Instead of East or West deviation may sometimes be labelled positive or negative, with a ‘+’ or ‘-‘
sign. Positive means deviation is East, negative means West.

Most aircraft have a compass deviation card derived by carrying out a “compass swing”. The
aircraft is placed in an area known as a compass base where the earth’s magnetic field has been
measured and known headings marked. The aircraft compass is then read when the aircraft is
pointing to these known magnetic headings and any deviation noted. This deviation card will then
allow the pilot to allow for any compass deviation when flying required magnetic headings.

11 Summary of Changes of Direction Reference


Changing between the different references for bearings can be a two-stage process, and variation or
deviation can easily be applied incorrectly, or in the wrong sense. To remember how to apply these
remember the mnemonic

Cadbury’s Dairy Milk Very Tasty

For : Compass Deviation Magnetic Variation True

Draw out a chart with these headings, and write in the information which you have available.
Calculate across the table, if you are calculating from left to right then add Easterly variation or
deviation (or positive deviation) subtracting Westerly (or negative). If your calculation runs from
right to left on the table subtract all Easterly and add all Westerly.

For Example

C D M V T
242° 2° W 27°E

To calculate magnetic bearing, subtract the 2° W deviation form the 242° compass bearing.

C D M V T
242° 2° W 240° 27°E

To then calculate true bearing add 26°E, because we are calculating from left to right on the table.

C D M V T
242° 2° W 240° 27°E 267°

2-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 3 Defining Position on the Earth

1 Position Reference Systems


Any system which defines position on the earth must be clear and unambiguous, especially when
used for navigation. The best and most common systems are based on Cartesian co-ordinates like
the typical x and y axes of a graph.

Figure 3-1 Cartesian Co-ordinates

The position of the object on the above scale is simply defined by stating the distance from the
origin ‘0’ along the x axis followed by the distance along the y axis.

Three systems are in common use in the UK;

a) Geographical Reference system (GEOREF)


b) UK National Grid
c) Universal Transverse Mercator (UTM) Grid

None of these systems are covered in the JAR syllabus, but are likely to be encountered
occasionally. All are suitable for use over relatively small areas, but are not suitable for world-wide
use.

2 Latitude & Longitude


Position on the earth’s surface is generally defined by an angular system of reference known as
latitude and longitude. The two axes are along the equator and along the Greenwich meridian
(often known as the Prime Meridian).

2.1 Latitude

Latitude is defined as the angular distance of a point North or South of the equator. It is a
maximum of 90° and is annotated North or South, depending on the position being North or South
of the equator. The units of measurement are degrees, minutes and seconds, (or sometimes
decimals of a minute) and latitude is always written with 2-fugure degrees and minutes.

General Navigation (NPA25) - Edition 4 – 110630 3-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 3-2 Latitude Measurement

The simplest definition of latitude is that it is the angle measured at the centre of the earth, known
as the Geocentric Latitude. Because the earth is not a true sphere, however, use of geocentric
latitudes would result in navigational errors, so a different definition is used. The Geodetic (or
Geographic) Latitude, defined as the angle between the plane of the equator and the normal (line at
90 degrees) to the spheroid surface, is used in navigation.

Figure 3-3 Comparison of Geocentric and Geodetic Latitudes

3-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2.2 Longitude.

The longitude of a point is the shorter angular distance measured at the polar axis between the plane
of the prime meridian and the plane of the local meridian. The Prime Meridian is the meridian
passing through an astronomical telescope at the original Greenwich Royal Observatory and is also
known as the Greenwich Meridian for this reason. It is annotated East or West, and is measured in
degrees, minutes and seconds relative to the Prime Meridian and always written as 3-figure degrees,
and 2-figure minutes if required. It cannot be greater than 180°. The 180° west meridian is the
same as the 180° east meridian, and is known as the Prime (or Greenwich) Anti-Meridian, because
it is the opposite side of the Earth from the Prime (or Greenwich) meridian.

Figure 3-4 Measurement of Longitude

General Navigation (NPA25) - Edition 4 – 110630 3-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

3-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 4 Measuring Distance on the Earth

1 Change of Latitude/Difference in Latitude


The change of latitude between two points is an angular measurement along the shorter arc of a
meridian between their two parallels of latitude. Its direction is annotated N or S measuring from
the first point to the second. If both points are in the same hemisphere, the smaller latitude is
subtracted from the larger. If, however, the points are on opposite sides of the equator the two
latitudes must be added to each other.

Change of latitude is normally abbreviated as ch. lat., and cannot be more than 180°. It is
sometimes referred to as ‘Difference in Latitude’ (d. lat).

Figure 4-1 Measurement of Ch. Lat

Example 1. Find the ch. lat between positions 45° 27′N and 23° 55′N.

Both positions are on the same side of the equator, so simply subtract the smaller latitude from the
larger, remembering that there are 60′ in one degree. Note especially that the ch. lat is in a South
direction here because the second position is South of the first.

Answer: 21°32′S

Example 2. Find the ch. lat between positions 27°49′S and 49°57′N.

These positions are on opposite sides of the equator, so the two angles should be added together.
The ch. lat is from South to North, so the answer is North.

Answer: 77°46′N

General Navigation (NPA25) - Edition 4 – 110630 4-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2 Change of Longitude/Difference in Longitude
The change of longitude is measured in a very similar way to the ch. lat. The change of longitude
(ch. long or d. long) between two points is the smaller arc of the equator intercepted by the
meridians through the two points. It is measured E or W depending on the direction from the first
point to the second.

Figure 4-2 Change of Longitude

4-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
As before, if both positions are East of Greenwich, the ch. long is found simply by subtracting one
longitude from the other (diagram ‘a’ above). If one point is East and the other West, however, the
two angles are added to each other (diagram ‘b’).

The maximum ch. long is 180°. Care must be taken to measure the ch. long in the shorter direction,
crossing the Greenwich anti-meridian if necessary. (Diagram ‘c’).

3 Distance Measurement
The distance between two points on the earth’s surface has so far been expressed only in terms of
angles. A number of other units are used, and it is often necessary to convert from one set of units
to another.

3.1 Units and Conversion Factors

1 yard = 3 feet
1 metre = 3.28 feet
1 kilometre = 3280 feet
1 statute mile = 5280 feet
1 nautical mile = 6080 feet (UK) 1852 metres (ICAO) - (but see following notes)

These conversion factors need to be remembered. Conversions from one unit to another often do
not work out precisely, (for instance, 1 nautical mile at 6080 feet ÷ 3.28 = 1853.6 metres), but they
are accepted as sufficiently accurate by the Civil Aviation Authority unless other conversion factors
are given.

3.2 Choice of Units

ICAO recommendations permit feet and metres for height measurement, and feet, metres,
kilometres and nautical miles for horizontal distance measurement. The units used by a particular
nation state are determined by the national government.

3.3 The Nautical Mile

The most common unit of distance used in air navigation is the Nautical Mile. There are various
definitions of its length, the simplest of which is:

The distance measured at the surface of the earth which subtends an angle of 1 minute
of arc at the earth’s centre.

Since the earth is not a perfect sphere, however, this definition results in variations in length.
Nevertheless, for most navigational purposes this definition is quite acceptable.

Another recognised definition of the length of 1nm. is:

The length of arc on the surface of the earth which subtends an angle of 1 minute at its own centre
of curvature.

General Navigation (NPA25) - Edition 4 – 110630 4-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 4-3 Variations in Length of 1 Nautical Mile

This definition also leads to a variation in the length of 1 nm because the earth is flatter at the poles
than at the equator. Because of the complications caused by the variation in length, ICAO has
standardised the nautical mile at 1852 metres. The very slightly different length of 6080 ft. is also
commonly used in calculations.

Flight at high altitudes means that, in order to cover 1 nm. at the surface, an aircraft needs to fly
further than 1 nm. through the air because of the greater radius of the aircraft’s flight path from the
earth’s centre. The correction required is very small, but is allowed for in some sophisticated types
of navigational equipment.

Note that 1 minute of latitude is 1 nm in length, and 1° of latitude is 60nm. Because the meridians
of longitude all converge at the poles, however, 1 minute of longitude is only equivalent to 1 nm at
the equator. It is therefore very easy to convert ch. lat to nautical miles, but the conversion of ch.
long to nm is not quite so easy.

4 Ch. Lat to Nautical Mile Conversion


Example 1: Find the distance between 47°36′N 020°05′E and 54°28′N 020°05′E in nm.

Both of the given positions are at the same longitude, so the distance between them is purely a
distance equivalent to the latitude difference. Simply subtract 47°36′ from 54°28′ and convert the
answer from degrees and minutes to minutes. Since 1′ = 1nm, the distance between the two points is
easily found.

54°28′ - 47°36′ = 6°52′ Since 1° = 60nm, 6°52′ = 412nm.

Example 2: Find the distance between 37°24′N 017°37′W and 23°47′S 017°37′W in nm.

Again, the two positions are at the same longitude, so the distance between them is simply one of
latitude. In this case, however, the positions are on opposite sides of the equator, so the distance
must be found from each of the two positions to the equator, and the answers added together to
provide a final answer.

4-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
The ch. lat from 37°24′N 17°37′W to 23°47′S 17°37′W is

37°24′ + 23°47′ = 61°11′.

61° X 60 = 3660′. Add the extra 11′ to give the answer 3671′ = 3671nm.

Example 3: Find the distance in nm. between point ‘A’, 57°28′N 026°38′E and point ‘B’,
83°52′N 153°22′W

At first sight this does not look like as straightforward ch. lat to nm. conversion because the
longitudes are different. However, a sharp eye will reveal that 26°38′E and 153°22′W are on a
meridian and its anti-meridian. The shortest distance between the two points is thus along the great
circle through the North Pole.

Figure 4-4 Distance Measured Over the Pole

The calculation of distance between the two points is thus relatively straightforward;

(90° - 57°28′) + (90° - 83°52′) = 32°32′ + 6°08′ = 38°40′

38° X 60 = 2280nm plus 40′ = 2320nm

It is also quite possible to even further complicate this type of exercise by making points A & B in
the question above to be very close to 180° apart in latitude as well as 180° apart in longitude. Care
must then be taken to calculate the shorter of the two distances.

General Navigation (NPA25) - Edition 4 – 110630 4-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5 Departure (Earth Distance)
The distance between two given meridians, when measured along a particular parallel of latitude, is
called Departure or ‘Earth Distance’ (ED). It is usually measured in nm. In the diagram below, the
distance from point A to point B is ‘departure’.

Figure 4-5 Departure

In the diagram, however, note that the ch. long from points A to B is the same as from C to D, but
the departure between those same points is quite different. For any given ch. long, the departure
between the two points C and D on the equator will always be greater than at any other latitude. As
we move towards the pole the meridians converge, and the departure between two points therefore
reduces, eventually reaching zero at the pole. Departure in nm. may be calculated using the
formula:

Departure (in nm) = Ch. long (in minutes) x Cos. lat

Note that departure is always measured along a given parallel of latitude. The distance between two
points in which both latitude and longitude are different cannot be calculated except by spherical
trigonometry - which is fortunately beyond the scope of this syllabus.

Example: Find the departure at latitude 59°N between positions at 017°W and 004°E.

The total ch. long is 17° + 4° = 21°. Converting 21° to minutes, 21 x 60 = 1260 mins of long.

∴ Departure = 1260 x cos. 59° = 1260 x 0.5150 = 648.95 nm.

4-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example:
What is the Earth Distance between positions 48°S 178°23′W and 48°S 161°17′E?

In an example like this it is first necessary to be sure that the shortest distance between the two
points is being calculated. Here it is clear that the shortest distance is across the Greenwich anti-
meridian, but remember that the ch. long between two points can never be more than 180°.

Ch. long from 178°23′W to 161°17′E is:

(180° - 178°23′) + (180° - 161°17′) = 20°20′ = 1220 mins.

Dep. = 1220 x cos 48° = 816.34nm.

From the diagram above, it can be seen that the departure in nm. between two meridians at points
on the equator is the same as the ch. long in minutes, since 1° of longitude at the equator is 60nm.
At any other latitude, however, earth distance must be calculated using the above formula because
of the convergence of the meridians.

General Navigation (NPA25) - Edition 4 – 110630 4-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

4-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 5 Meridian Convergence and Conversion Angle

1 Meridian Convergence
The representation of a spherical earth on a flat paper chart often makes life difficult for a navigator.
In earlier diagrams, it had already been seen that the meridians converge at the poles, and it is
sometimes necessary to calculate the convergence of the meridians in degrees. The diagram below
shows that, at the equator, all meridians are parallel to each other, but the angle between them at the
pole is equal to the ch. long. Convergence may also be known as Convergency, Earth Convergence,
Meridian Convergence or True Convergence.

2 Definition of Convergence:

The angle of inclination between two selected meridians at a specified latitude.

Figure 5-1 Convergence of Meridians

At intermediate latitudes, the convergence between the meridians is found by using the formula:

Convergence = Ch. Long (in degrees) x sin. Mean Lat

Occasionally it is necessary to find the meridian convergence over a route or part of a route in
which the latitude changes during the flight. In such a case the sine of the mean latitude is used in
the formula.

Example:

Find the convergence at a latitude of 59°N between positions 142°55′E and 171°47′W.

General Navigation (NPA25) - Edition 4 – 110630 5-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
The ch. long between these two positions, (measured, note, across the Greenwich anti-meridian) is:
(180° - 142°55′) + (180° - 171°47′) = 45°18′

∴Convergence = 45°18′ x sin. 59° = 38°°49′′

Example:
What is the average convergency between positions 5°23′S 007°35′W and 32°47′N 029°49′E?

This is a somewhat complicated question, even though the procedure is basically simple. The
latitude change crosses the equator, and the longitude change crosses the prime meridian, so care is
needed in calculations. The mean latitude is half way between 5°23′S and 32′47′N, so first we must
find the ch. lat. Using the method taught in earlier chapters, the ch. lat in this case is found by
adding the two latitudes together:

5°23′ + 32′47′ = 38°10′

This is the total ch. lat, so we have to divide the ch. lat by 2 ( = 19°05′ ) and add or subtract this
figure from one of the given latitudes to find the mean lat. In this case it is simplest to subtract it
from 32°47′N.

Mean Latitude is: 32°47′N - 19°05′ = 13°42′N

Now find the ch. long. This is done in this case by simply adding the two longitudes together:

Ch. Long is: 7°35′ + 29°49′ = 37°24′

Convergency = ch. long x sin. mean lat.

= 37°24′ x sin. 13°42′

= 8°°51′′

3 Importance of Convergency
The importance of convergency is its relationship with the bearing of a great circle. Note that a
great circle does not have a constant bearing (Rhumb lines, which do, are discussed next). As an
aircraft moves along a great-circle track its true course must constantly change. However a great
circle is the equivalent of a straight line – the shortest distance between two points, so does not
really change in direction. The only reason the course changes is that it is measured from the
meridian as a north reference. These meridians are at a different angle, so the bearing changes.

Of course the new bearing has changed by the same angle as the change in meridian direction. This
change is the same as the angle between those meridians, which is convergency. The importance of
Earth Convergence is therefore as a measure of the change of great-circle bearing.

4 Great Circles and Rhumb Lines


It has been seen in an earlier chapter that a great circle is a circle with its centre at the centre of the
earth, and that the shortest distance between two points is along a route which follows the great
circle. For long-range navigation, therefore, the great circle route would be chosen whenever
possible. Navigation on a great circle route, however, is somewhat complicated by the fact that
5-2 General Navigation (NPA25) - Edition 4 – 110630
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
most great circle routes require constant changes of direction in order to follow them. A simpler
method would be to follow a rhumb-line track.

Definition of Rhumb line:

A rhumb line is a regularly curved line on the surface of the earth which cuts all meridians at
the same angle.

Look at the diagram below. Note that there is a constant change of direction with respect to True
North as this particular great circle route passes from the equator to the high latitude destination.
The only great circles which do not change direction are the equator or any of the meridians. Note
also that the earth convergence between any two points on the great circle (such as A and B in the
diagram) is the same as the change in great circle direction between those same two points.

Figure 5-2 Great Circles and Rhumb Lines

The rhumb line route corresponding to the same route crosses all the meridians at the same angle. It
is longer, but may be easier to use for navigation depending on the method used. The rhumb line
between two points is always on the equatorial side of the great circle. The equator and all parallels
of latitude are rhumb lines as they cross all meridians at a constant angle of 90 degrees. All
meridians and anti-meridians are also special cases of rhumb lines as they meet the definition of a
rhumb line in that they cross all meridians at a constant angle – 0 degrees.

General Navigation (NPA25) - Edition 4 – 110630 5-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5 Conversion Angle
Any two points on the earth’s surface can be joined by both a rhumb line and a great circle.
Occasionally the two will coincide, for example, if they are along a meridian or the equator, but
usually they will diverge. The angle at which the rhumb line and the great circle between two
points diverge is called the Conversion Angle

The conversion angle (CA) between two points is half of the convergence between those points.

∴Conversion Angle = ½ Ch. Long x Sin. Mean Lat

The diagram below illustrates these various relationships in the Northern Hemisphere. In the
Southern Hemisphere the meridians will of course be converging in the opposite direction, and the
great circle will be on the opposite side of the rhumb line.

Figure 5-3 Relationship between Conversion Angle, Rhumb Line and Great Circle in the
Northern Hemisphere

Note: Over short distances in areas away from polar regions the average great circle track is
approximately equal to the rhumb line true track between two positions.

5-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section Two Basic Navigation

Chapter 6 The Navigation Computer (Part 1)

1 Definitions
Definitions required in this and following chapters are:

Indicated Airspeed (IAS). Indicated Airspeed is the airspeed read directly from the airspeed
indicator, without corrections applied.

Rectified Airspeed (RAS). Rectified Airspeed, also known as Calibrated Airspeed (CAS), is the
indicated airspeed corrected for instrument and pressure (or position) errors.

True Airspeed (TAS). True Airspeed is the speed of the aircraft measured relative to the air mass
through which it is flying. It is found by applying corrections for density and, if necessary,
compressibility to the CAS.

Groundspeed (GS). Groundspeed is the speed of the aircraft relative to the earth’s surface. It is
found by applying a correction to the true airspeed to allow for the effect of the wind on the aircraft.

Equivalent Airspeed (EAS). Equivalent Airspeed is almost never used by the pilot. It is found by
applying a correction only for compressibility to the RAS. Air density corrections are not included.

2 The Circular Slide Rule

Figure 6-1 The Circular Slide Rule

This chapter is based on the Pooley’s CRP-5 Navigation Computer. Other types of navigation
computers which are able to perform all of the same functions as the Pooley’s instrument are quite

General Navigation (NPA25) - Edition 4 – 110630 6-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
acceptable, but in some cases the answers to problems may differ slightly from those given in these
notes. The illustrated side of the CRP-5 computer consists of a circular slide rule which enables
many different types of navigation calculation to be carried out. Two logarithmic scales are marked
on the computer, the inner of which rotates.

2.1 Multiplication

To carry out a simple multiplication exercise such as

4 x 6 = 24

First find the figure 10 on the inner scale. Now rotate the inner
scale until this figure 10 is exactly aligned with 4 (usually
marked as 40) on the outer scale. Without moving the inner
scale, turn the whole computer until you find the figure 6
(usually marked as 60) on the inner scale. The answer to the
problem (24) is to be found on the outer scale aligned with 60 on
the inner scale.

Figure 6-2 Simple Multiplication

Note that no indication of the correct position of the decimal point is given. This must be worked
out by doing a rough mental calculation. Now try a few more of this type of very simple
calculation. Once this has been mastered, practise getting the position of the decimal point in the
right place each time, doing calculations like these:

62 x 50 = ? 17 x 25 = ? 4.5 x 90 = ?

2.2 Division

Division using the Nav Computer is the


reverse of the method of multiplication shown
above. To divide 15 by 2, for instance, place
15 on the outer scale against 20 on the inner
scale, and read off the answer of 7.5 against
the 10 on the inner scale. Once again, great
care is required to ensure that the decimal
point is correctly placed in the answer.

2.3 Proportions

The CRP-5 is very useful for proportion


calculations. Once the CRP-5 has been set in
the example above, the proportion 75 over 10
is the same as 15 over 2, and is the same as all
the other readings with the two scales in this Figure 6-3 Division and Proportion
position. Without moving the scale, for
example, look at 90 on the outer scale. You
will find that 90 over 12 is exactly the same

6-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
ratio as 75 over 10 or 15 over 2. The risk of misplacing the decimal point is very great, however.
Still without moving the scale, look at the figure 30 on the outer scale. Is the proportion 30 over 40
or 30 over 4?

75 15 90 30
= = =
10 2 12 4

Example:

An aircraft descends at 650 ft per minute. How long will it take to descend through 14 000 ft?

This is a straightforward proportion question, and the answer is found on the CRP-5 by setting 650
over 1, and reading the answer against 14 000.

Set 650 on the outer scale against the figure 1 on the inner scale. Now find 14 000 on the outer
scale. The inner scale reads 21.5 against the 14 000. Answer - 21.5 minutes.

2.4 Conversions Between Units

The CRP-5 is able to assist in converting from one set of units to another.
i. Metres, Yards and Feet
Find the ‘metres’ mark (shown as ‘km-m-ltr’) on the outside scale, located at the figure 10
position. To convert 250 metres to yards, set 250 on the inside scale against the metres mark,
and read off the yards equivalent against the yards mark. (Answer 273 yards). To convert to
feet, read off the answer against the ‘feet’ mark. (Answer 820 feet). The accuracy of the
answers is, of course, limited by our ability to set and read the scale precisely.
ii. Litres, Imperial and US Gallons
In converting from litres to gallons and vice-versa, it will be helpful to remember that there
are roughly 4 litres to one gallon. To convert 325 litres to gallons, set 325 on the inner scale
against the ‘km-m-ltr’ mark on the outside scale. Read off the gallons equivalent against the
Imperial or US gallons marks on the outside scale. (Answers: 71.5 Imperial gals, 86 US gals).
iii. Kilometres, Nautical and Statute Miles
To convert 30 statute miles to nautical miles and kilometres, set 30 against the statute mile
marker on the outside scale. Read off the nautical miles and kilometres against their
respective markers. (Answers: 26 nm., 48.3 km.).
iv. Volume to Weight
Both pounds and kilogrammes are commonly used in fuel calculations. Conversion of
volume to weight requires a knowledge of the specific gravity of the fuel. Although this is
normally about 0.8, variations from approximately 0.72 to 0.82 may be encountered. To
convert 708 imperial gallons at a S.G. of 0.74 to kg. and pounds, set 708 on the inner scale
against the imp. gal mark on the outer scale.

For kgs., find the kgs. S.G. marks on the outside scale, (do not confuse these marks with the
lbs. S.G. marks), and set the cursor to the S.G. of 0.74. Read off the answer of 2380 kgs. on
the inner scale. It will help if you remember that there are about 4 litres to a US gallon and
4.5 litres to an Imperial gallon.

General Navigation (NPA25) - Edition 4 – 110630 6-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
For lbs., find the lbs. S.G. marks on the outside scale. Set the cursor to the S.G. of 0.74, and
read off the answer of 5250 lbs on the inside scale.

2.5 Distance, Speed and Time

The circular slide rule is designed to perform the many distance, speed and time calculations which
are required in navigation. Always set time on the inner scale.

Example: How far do you travel in 35 minutes at 190 kts?

Note the prominent black ‘speed’ arrow at the ‘60’ position on the inner scale. Set this to the speed
of 190 kts on the outer scale. Now, remembering the rule that time always appears on the inside
scale, find 35 minutes on the inner scale and read off the distance of 111nm.

Example: You fly 480 nm in 1 hour 18 minutes. What is your speed in kt?

Set 480 nm on the outer scale against 78 minutes on the inner scale. (Note that the time is measured
in minutes. Some computers have an additional scale for hours which may be helpful.) Read off
the speed of 370 kt on the outside scale against the speed arrow. Using an electronic calculator will
give a slightly more accurate answer for this type of calculation, but the greater accuracy is never
required for routine navigation.

Example: How long will it take you to fly 625nm at 485 kt?

Set the speed arrow to 485 kt on the outer scale. Remembering that time always appears on the
inner scale, read off the time in minutes against 625 on the outer scale.
(Answer: 77.3 minutes).

Example: An aircraft is found to use 8 000 kg of fuel in 40 minutes. If it is flying at a groundspeed


of 380 kt, how many US gallons (S.G. 0.75) are required to cover a distance of 290 nm?

This is a common type of fuel consumption problem. Work in kgs throughout, converting to US
gallons at the end. Set the speed arrow to the groundspeed of 380 kt, and against the distance of
290 nm on the outer scale, read off the time to fly on the inner scale. (45.8 mins.)

Now set 8 000kg on the outer scale against 40 minutes on the inner scale – this gives the hourly
consumption if we need it (we don’t in this case,) – but against the flight time of 45.8 mins on the
inner scale read off the fuel required on the outer scale. (9150 kg).

Finally, convert this to US gallons. Set the moveable cursor to 0.75 on the kg/S.G. scale, and set
the required fuel in kg under the cursor. Before you convert to gallons, look at the litres marker,
and you will note that our 9150 kg at this S.G. is about 12 500 litres. At roughly 4 litres to one US
gal., therefore, you should be expecting an answer of about 3 000 US gals. Now rotate the cursor to
the US gals marker on the outer scale and read off the answer.
Your answer should be 3225 US gals.

6-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2.6 Calculation of True Airspeed

The airspeed indicator is only accurate when the air density is that of the ISA at mean sea level,
1225g/m3. Whenever any other air density prevails, a correction must be applied to the indicated
airspeed to find true airspeed. The CRP-5 airspeed window is used to calculate the correction.

Figure 6-4 CRP-5 Airspeed Window

Example: Calculate the true airspeed of an aircraft flying at a RAS of 130 kt at a pressure
altitude of 31 000 ft in an ambient temperature of minus 28°C.

Set up the airspeed window of the CRP-5 as in the picture above, with 31 000 ft against the
temperature of minus 28°C. Now find the RAS of 130 kt on the inner scale, and read off the TAS
on the outer scale. You should find an answer of 221 kt.

Example: Find the TAS of an aircraft which is flying at 25 000 ft and a temperature of minus
32°C at a RAS of 154 kt. (Answer: 228 kt)

Example: What is the TAS of an aircraft flying at 1000 ft press. alt. at a RAS of 245 kt in an
ambient temperature of plus 30°C? (Answer: 256 kt)

Beware of the unusual. Mistakes are often made by setting minus temperatures when the question
gives a temperature above zero. Note also that at low levels and in temperatures much colder than
ISA the TAS will be lower than the RAS:

Example: What is the TAS of an aircraft flying at 2000 ft press alt. at a RAS of 195 kt in an
ambient temperature of minus 35°C? (Answer: 184 kt)

2.7 Mach Number Calculations

Aircraft which fly at high speeds and high levels use Mach numbers for speed control. Mach
number is covered in detail in the training notes for Instrumentation and Principles of Flight, so it is
only necessary here to cover the conversion of TAS to Mach number and vice-versa.

Example: What is the TAS of an aircraft which is flying at a Mach number of 0.84 in an
ambient air temperature of minus 40°C?

First find the Airspeed window on the CRP-5, and rotate the inner scale until the Mach index arrow
appears in the window. (See the next diagram.) Set the temperature of - 40°C against the Mach
arrow. Now look on the inner scale for the Mach number of 0.84, and read off the TAS on the outer
scale.

General Navigation (NPA25) - Edition 4 – 110630 6-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 6-5 The Mach Scale of the CRP-5

Your answer should be close to 502 kt.

2.8 Compressibility Errors and Corrections

As aircraft speed increases, the air in the pitot line becomes compressed. At true airspeeds of 300 kt
and above, this causes significant errors which need to be allowed for in calculating TAS from
RAS. This is done by using the compressibility correction window on the CRP-5.

Figure 6-6 Compressibility Correction Window (1)

Example: Calculate the TAS of an aircraft flying at an RAS of 250 kt at Flight Level 410 with
a temperature of – 56.5°C.

First, calculate the TAS in the normal way by setting the temperature of –56.5°C against FL410 in
the airspeed window. An RAS of 250 kt gives a TAS of 504 kts, so a correction is required because
the calculated TAS is more than 300 kts. Now look at the instructions below the compressibility
correction window in the picture above. Divide the TAS by 100 and subtract 3 from the answer.
504 ÷ 100 = 5.0, and 5.0 - 3 = 2 .0. Now, using the vertical arrow in the compressibility
correction window as a reference, rotate the inner scale anti-clockwise by 2.0 divisions. It is most
likely that this will have to be estimated, as the vertical arrow is unlikely to fall in a convenient
place. Finally, re-enter the inner scale with the given RAS of 250 kt, and re-read the TAS on the
outer scale. The TAS after compressibility correction is 481 kt.

6-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example: Calculate the TAS of an aircraft flying at an RAS of 196 kt at Flight Level 370 and
at an ambient temperature of – 48°C.

The initial uncorrected TAS works out at 366 kt. Dividing this figure by 100 and subtracting 3
gives a correction of 0.66 divisions in the window. After rotating the inner scale anti-clockwise by
this amount and re-entering the inner scale with an RAS of 196 kt, the final corrected TAS is 361
kt.

2.9 Calculation of RAS from TAS or Mach Number

This is a straightforward procedure provided that the TAS is not high enough to have been affected
by compressibility – i.e. up to a maximum of 300 kt.

Example: Calculate the RAS to fly in order to achieve a TAS of 295 kt at a pressure altitude of
28 000 ft in an ambient air temperature of - 22°C.

Set the CRP-5 airspeed window to 28 000 ft at -22°C. Against the TAS on the outside scale read
off the required CAS on the inside scale. Answer: 183 kt.

Example: (This example is not at all easy because compressibility is involved.) Calculate the
CAS of an aircraft which is flying at FL 410 in an outside air temperature of -53°C at Mach 0.85.

First find the TAS. This is easily done by setting the Mach arrow against the temperature of minus
53°C and reading off the TAS on the outer scale against the Mach number on the inner scale.
Answer: 491 kt.

The rest of the procedure is seriously complicated by the fact that the CRP-5 is not intended to
make this calculation. However, it can be done with care. First, set the airspeed window to the
given figures of -53°C and FL 410 (Pressure alt 41 000 ft.) Now without moving the inner scale,
turn the computer round until you can read the compressibility window.

Figure 6-7 Compressibility Correction Window (2)

Entering the formula written below the window with your TAS of 491 kt, work out how much you
need to move the scale. [(491 ÷ 100) – 3 divisions = 1.9 divisions.] Move the scale in the direction
shown (anti-clockwise) by this amount. Now read off the RAS on the inner scale against the TAS

General Navigation (NPA25) - Edition 4 – 110630 6-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
on the outer scale – your answer should be close to 252 kt. Make a note of this answer, although it
is not a correct figure because the compressibility error has not yet been properly applied. Now
reset the airspeed scale to FL 410 & - 53°C, and enter the inner scale with the RAS of 252 kt you
just found. Read off the TAS on the outer scale. (510 kt.) Now go back to the compressibilty
window, and recalculate how much you need to move the scale for compressibility using this new
figure of 510 kt.
The figure now (the correct one this time,) is: (510 ÷ 100) – 3 = 2.1 divisions. Rotate the inner
scale by this amount, then re-enter the outer scale with the correct TAS of 491 kt and read off the
correct RAS on the inner scale. Answer: 254 kt.

2.10 Indicated to True Altitude Corrections.

An altimeter which is set to the correct QNH will not read accurately if the air temperature below
the aircraft does not match the ISA lapse rate. The CRP-5 may be used to find the true altitude as
follows:

Example: What is the true altitude of an aircraft which is flying at 3000 ft indicated, using the
correct local QNH, and with an ambient air temperature at the aircraft of -45°C?

Using the Altitude window, set 3000 ft against the temperature of - 45°C. Now enter the inner scale
with 3000 ft (the figure 30), and read off the true altitude on the outer scale. (Answer: 2430 ft)

Note how large the error can be if you were flying, for instance, in a Siberian or Canadian winter.
In deciding where to place the decimal point, remember that the error is roughly 100 ft per 1°
difference from ISA, and cold weather is the dangerous side of the line. If the temperature is above
ISA, the aircraft will be above the indicated altitude.

2.11 Density Altitude

A correction window is provided on the CRP-5 for calculating density altitude where this is
required. The UK Civil Aviation Authority, however, have declared the CRP-5 corrections
obtained through the use of this window are not sufficiently accurate, so it is not to be used in
examinations. The method to be used will be taught from the Meteorology notes.

6-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 7 The Triangle of Velocities

1 Velocities
There is often confusion between the words ‘speed’ and ‘velocity’. A vehicle may be described in
terms of its speed, but this gives no indication of its direction, which may be changed without its
speed being affected. A velocity, on the other hand, is a combination of both direction and speed
or, in other words, speed in a particular direction. An aircraft which is flying on a heading of 045°
(T) at 350 kt, for instance, has a different velocity from an aircraft which is also flying at 350 kt but
on a heading of 135° (T).

2 Vectors
Velocities may be represented as a straight line of a given length and drawn in a specified direction,
and are then known as ‘vectors’.

Figure 7-1 A Vector

3 The Vector Triangle


Three of these vectors are commonly used in navigation to define the movement of an aircraft.

They are:
a. The Wind Velocity (W/V) Vector. This is a vector which represents the speed and direction
of the wind over the ground°. It is written as 080°/25, which means a wind which is
blowing from a direction of 080° at a speed of 25 kt. Note that the direction is ‘from’ where
the wind is blowing, not ‘to’.
b. The Heading/TAS Vector. This vector represents the movement of the aircraft through the
air mass in which it is flying. The aircraft flies at the TAS in the direction defined by the
heading. Note that heading and TAS always go together in this vector.
c. The Track/Groundspeed Trk/GS Vector. This is the vector which represents the movement
of the aircraft at the groundspeed in a particular direction over the ground. Note that track
and groundspeed always go together in this vector.

General Navigation (NPA25) - Edition 4 – 110630 7-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Each vector is given a symbol by convention; a single arrowhead for the heading/TAS vector, a
double arrowhead for the Trk/GS vector, and a triple arrowhead for the W/V vector. Together, the
three vectors form a ‘Vector Triangle’ which represents all these variables, and shows in
diagrammatic form how and why the aircraft moves over the ground in a particular direction at the
groundspeed. The aircraft is steered from point A in the direction of point B, but because of the
movement of the air (the wind), the aircraft actually moves towards point C.

Figure 7-2 The Vector Triangle

The vector triangle can be drawn to scale, and if any two of the vectors are known, the third may be
found by measuring. In the above drawing, an aircraft which leaves an aerodrome at point A flying
at a TAS of 120 kt on a heading of 063° will arrive at point B after 1 hour. Point B, however, is a
point in the air mass through which the aircraft is flying (known as an ‘Air Position’).

During this 1-hour flight, the W/V of 014°/37 has been acting on the aircraft, moving the whole air
mass in the direction B to C. The aircraft is therefore actually at point C (the aircraft’s ground
position) at the end of its 1 hour travel. It is particularly important to note that the wind vector
always goes from the end of the Hdg/TAS vector to the end of the Trk/GS vector.

This vector triangle can be used in practice, with the W/V perhaps being obtained from a wind
forecast, to predict the aircraft position C after 1 hour. The track and groundspeed can be measured
directly from the drawing – this one works out as a track of 079° and a groundspeed of 100 kt.

A more practical use of the triangle of velocities is likely to be that the aircraft leaves point A, and,
after 1 hour, the pilot fixes his position at point C by map reading. In this case the two vectors AB
and AC are known, AB because the pilot is presumably aware of his heading and airspeed, and AC
because the track and groundspeed can be measured from the chart. Plotting these two vectors on a
chart, using a suitable scale, the pilot can find the W/V BC by measurement.

3.1 Drift.

The angle between the aircraft’s heading and its track is known as ‘Drift’. It is always measured
from heading to track, and is annotated ‘Port’ (left) or ‘Starboard’ (right).

7-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 7-3 Drift Angle Measurement

4 Examples
Using graph or other suitable paper, draw out the following examples to check that you have
understood the process. Drawings in this manual are not likely to be precisely to scale because of
possible distortions during printing.

Example 1: An aircraft flies from A to B at a TAS of 180 kt on a heading of 350°. It is subject to a


W/V of 060°/25. What is its track, groundspeed and drift?

Figure 7-4 Example 1

Draw the Hdg/TAS vector as shown above, using some vertical line on the paper as a North datum.
Draw the vector AB at a length equivalent to the airspeed. Now from B draw in the W/V vector at
the same scale, taking care to get the wind direction right – from B. Finally, join A to C and
measure the track angle from your North datum line, and measure the distance from A to C to find
the groundspeed. Your answer should be very close to Trk 342°° and GS 173 kt. The drift is the
angle between the heading (350°) and the track (342°) measured from the heading. The answer is
in this case 8°° port.

General Navigation (NPA25) - Edition 4 – 110630 7-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example 2: After a 30-minute flight from your base at A you note that you have flown a distance of
82 nm. along a direction of 173°. Your TAS has averaged 195 kt and your heading was 184°.
What is the wind velocity?

Figure 7-5 Example 2

Draw in the two vectors for Hdg/TAS and Trk/GS, carefully noting the trap that the ground distance
is flown in 30 minutes, not a full hour. The groundspeed is therefore 164 kt. Now join B to C,
which is the wind vector. The distance is easy, but what is the wind direction? Remember that the
wind blows from heading to track, so that the wind direction is from B to C. Always do a double-
check that your answer is sensible. Your heading is to the right of your track, so if the wind blows
from heading to track the wind must be coming from somewhere on your right (seen from your
pilot’s seat). Also, your groundspeed is around 30 kts less than your TAS, so your wind cannot be
less than 30 kt and it must have a strong headwind component no matter which direction it is
coming from.

Your answer should be close to 227°°/46.

Example 3: You are required to fly from point A, leaving at 1000 hrs, and to arrive overhead a
VOR beacon at point B at 1018hrs. The track from A to B is 229° and the distance between them is
126nm. If the W/V is 290°/70, what heading and TAS must you fly?

First find the required groundspeed. If you can’t do this easily in your head (and it is easy!), use
your CRP-5, and you should find it is 420 kt. Plot on your paper the track and groundspeed of 229°
and 420kt. Now plot in the wind, taking great care to plot its direction the right way round.

7-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
.

Figure 7-6 Example 3

Finally join A to C in your diagram to find the third side of the vector triangle and measure the
direction and distance to find the hdg & TAS.

Your answer should be close to 236°° & 457 kt.

Example 4: You are required to fly a track of 279° in your aircraft, which has a cruising TAS of
330 kt. If the W/V is 210°/60, what is your required heading, and what will be the groundspeed?

This is a somewhat different problem. In this case you are told the W/V, but instead of being given
one other vector you are only given one value of each of the other two sides. First draw in the track
of 279° (side AB in the diagram below), but without setting any particular length to the vector. Plot
in the W/V as normal, (side BC), once again taking care to plot in the correct direction. Now take
your compasses, and set them to a scale distance equivalent to the TAS of 325 kt. Mark off from C
an arc which crosses AB at D, where AD is the TAS. The direction of DC is the required heading,
and the distance DB gives the groundspeed. This is by far the most common type of navigational
problem for which the triangle of velocities needs to be resolved.

Figure 7-7 Example 4

Your answer should be close to Hdg 269°° GS 305 kt.

5 Directional Datums
Different directional datums have not been mentioned in the examples above, but navigational
problems will be given in future which use a confusing combination of differing datums. It is
likely, for instance, that you will be flying a magnetic heading but have to use a true wind to
General Navigation (NPA25) - Edition 4 – 110630 7-5
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
calculate your track. It is essential that you convert all your information to a single datum before
you begin any of these calculations. If you have both magnetic and true information, it is usually
(but not always) best to convert everything to true.

7-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 8 The Navigation Computer (Part 2)
We have so far looked only at the circular slide rule side of the navigation computer. The other
purpose for which the computer is designed is to solve vector triangle problems. The important
parts of the vector triangle are drawn on the face of the computer, enabling the pilot to quickly find
the unknown vector or parts of the vector required for navigation.

1 The Slide
The moving slide has two sides, one for high and the other for low speeds. The method of
operation is the same for both, so we will use only the low-speed side for this first illustration.

Figure 8-1 The CRP-5 Slide, Showing the Wind Point

The slide is marked with arcs which show the speed, and radial lines which represent direction. In
the diagram above, for example, the Hdg/TAS vector has been drawn to show a TAS of 150 kt, and
the Trk/GS vector shows a groundspeed of 140 kt with a drift angle of 15° Port. In most cases, the
only thing which actually needs to be marked on the CRP-5 is the ‘Wind Point’. In the above
illustration there is of course no indication of what the heading and track actually are, but the
rotating compass rose fitted over the slide give a direction to the triangle.

2 Use of the Computer

2.1 Marking the Wind Point

There are various ways of using the computer, but we will standardise on a single method which
will be used throughout these notes. In a typical problem in which the pilot is given the wind to

General Navigation (NPA25) - Edition 4 – 110630 8-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
solve the vector triangle the first requirement is to mark the wind point on the computer. Given a
W/V of, say, 020°/30 kt, mark the wind point as follows:

a) Rotate the circular scale so that 020° is aligned with the ‘True Heading’ index mark.
b) Move the slide so that the ‘Zero’ line on the slide is positioned below the centre circle on the
rotating CRP-5 face.
c) Mark the face (a soft pencil is best) with a small, accurate cross at the 30 kt position.

Figure 8-2 Marking the Wind Point

Now that the wind is marked on the computer we can use it to find the effect of that wind on a range
of aircraft headings and airspeeds. Computers vary slightly, but you should find that you are able to
be accurate to within 1° and 2 kt.

2.2 Finding Track and Groundspeed

Example: Given a Hdg/TAS of 335° and 140 kt, find the Trk/GS, using the wind of 020°/30 kt
already set.

Set the slide so that the TAS of 140 kt appears under the centre circle, and rotate the compass rose
to set the heading of 335° to the heading index. (Note that the TAS is always set under the centre
circle.) You will now find that the wind point has moved to the new position illustrated in the next
diagram. Remember that the centre circle is the end of the Hdg/TAS vector, and is therefore at the
upwind end of the W/V vector. The wind point is therefore at the downwind end of the W/V vector,
so the wind point also defines the end of the Trk/GS vector. Now you can read off the groundspeed
under the wind point – you should read close to 120 kt.

8-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Note that the wind point is on a direction line which is marked at 10° left of the centre circle. This
shows that your drift on this heading of 335° is 10° port, making your track 325°°. You can work
this out in your head, but it is probably easier to use the drift scale either side of the ‘True Heading’
index.

Figure 8-3 Finding Track and Groundspeed

Example: Given a W/V of 250°/25, find the Trk/GS of an aircraft flying on a Hdg of 178° at a
TAS of 170 kt.

Go through the procedure of setting the wind point again, then find the Trk/GS as shown in the
above example. Your answer should be close to 170°° and 164 kt.

Example: Given a W/V of 070°/90, find the Trk/GS of an aircraft flying at 475 kt TAS on a
Hdg of 272°.

This example requires the use of the high-speed side of the slide, but it is otherwise carried out in
the same way as the first problem. When you have rotated the compass rose to the correct heading
you will find that you have to take some care with your readings. The drift falls at about 3¾°, but
you will never need to read it to better than ½°. Your answer should be close to Trk 268°° GS 560
kt .

2.3 Finding the Wind Velocity

Another velocity triangle problem is to find the W/V if the other two vectors are given. Consider
the following problem:

Example: Given a Hdg/TAS of 270° & 120 kt, and a Trk/GS of 288° & 146 kt, find the W/V.

General Navigation (NPA25) - Edition 4 – 110630 8-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
This is a low-speed slide problem, so first make sure that you have the correct side of the slide
facing you. Set up the computer so that you have the TAS under the centre circle, and set the
heading of 270° against the heading index. In order to find the W/V, we have to find out where the
wind point is. But remember that the wind point is at the end of the Trk/GS vector. We are given
the Trk, so we know that the drift is 18°S. (You don’t have to work it out; you can read it off the
drift scale either side of the heading index mark.)

Figure 8-4 Finding the Wind Velocity

Mark the drift line on the computer face, and also mark the given groundspeed of 146 kt. These
two lines intersect at the wind point. To find the W/V it is now necessary only to rotate the
compass rose until the wind point is vertically below the centre circle. You can now read the wind
direction against the heading index, and the wind speed by interpreting the scale on the slide. Your
answer should be close to 157°°/48 kts. Note, however, that it is quite difficult to obtain accurate
results using this procedure, and an accuracy of 5° and 3 kt is acceptable.

2.4 Finding Heading and Groundspeed

This is the most difficult, but also the most common of all the CRP-5 velocity problems. You are
likely to have a forecast W/V, you will know your required track and your intended TAS. Notice,
however, that you know only one of the values of each of the air and ground vectors. Some degree
of jiggery-pokery is therefore required when finding the answers to these problems.

Example: You are required to fly along a track of 278° at a TAS of 470 kt. The forecast W/V
is 050°/90 kt. What is your heading and groundspeed?

First, mark the wind point, noting that this is a high-speed problem. Now set the slide to show the
TAS of 470 kt under the centre circle and the Track (Note!) against the heading index. Now look

8-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
at the new position of the wind point. It is indicating a drift of about 7° port, so we must rotate the
compass rose until the track of 278° appears under the 7° port mark on the heading index scale.
Now look again at the drift indicated under the wind point, and you will see that it has now
increased to 8° port. Move the compass rose again until the required track of 278° appears under
the drift index mark shown by the position of the wind point.

When you have the required track set against the correct drift index mark as shown under the wind
point, you can now read off the groundspeed under the wind point and the required heading against
the heading index.
Your answers should be close to Hdg 286°° and GS 526 kt.

Try another example, this time on the low-speed side of the slide.

Example: Required Trk 355°, TAS 100 kt, W/V 040°/30. Find the Hdg & GS.

Your answer should be close to 007°° & 77 kt.

2.5 Multi-Drift Wind-Finding

A now obsolete method of wind finding is by measuring the drift on a number of different headings
(usually 3 headings 60° apart). It was reasonably accurate, and was once in common use by large
piston-engined aircraft which either carried a tail gunner or were equipped with a means of visually
measuring the drift (a ‘drift sight’).

Example: A navigator notes the following drift angles on three different headings. If the
aircraft is flying at a TAS of 140 kt, what is the W/V?

Hdg Drift
320° 11°P
260° 10°P
020° 3°S

Using the low-speed slide, first set the TAS under the centre circle and the first heading of 320°
against the heading index. Draw a pencil line down the 11°Port line. (Diagram (a) below.) Now
rotate the compass rose to place the second heading of 260° against the heading index, and draw
another line along the 10° Port line. (Diagram (b)). Finally set the third heading of 020° against the
heading index and draw a third line along the 3° Starboard line. (Diagram (c)).

You should now have 3 pencil lines intersecting at something close to a single point. This is the
wind point. Turn the compass rose until the wind point is vertically below the centre circle and read
off the wind velocity. The wind you have found should be approximately 010°°/30 kt.

General Navigation (NPA25) - Edition 4 – 110630 8-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
(a) (b)

(c)

Figure 8-5 Multi-Drift Wind Finding

2.6 Wind Components

As shown in an earlier chapter, it is possible to find head/tail/crosswind components using scale


drawing. It is much quicker, however, to use the navigation computer for the purpose.

Example: Air Traffic Control passes you a surface W/V of 350°(M)/32 kt. What are the
head/tail and crosswind components if the QDM of the runway in use is 020°(M)?

First, a note of caution. All the examples we have used so far have assumed that we are using only
true direction. Now, because we are operating on an aerodrome we use magnetic directions. The
runway QDM is, of course, Magnetic, and the W/V we use with it to calculate our wind effects
must also be Magnetic.

8-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Mark the wind point of 350°/32 kt on the computer in the usual way. Now, without moving the
scale, leaving the zero line under the centre circle, rotate the compass rose to place the runway
QDM of 020° (M) against the heading index.

Figure 8-6 Crosswind & Head/Tailwind Components

i. Headwind/Tailwind.
Now read off the headwind component, reading vertically down from the zero line to the wind
point. In this example the headwind is about 27 kt. If the wind point comes above the zero
line there is a tailwind component, and you will have to set the reciprocal (a difference of
180°) of the runway QDM against the heading index to be able to read the strength of the
tailwind.
ii. Crosswind
The crosswind is read left or right from the central vertical line of the slide. In this example it
is about 16½ kt. But – there’s a trap here. Which direction is the wind coming from? You
must remember that the wind blows from the centre circle towards the wind point. The wind
is blowing from left to right. It should be obvious, but it’s a common mistake.

A worked example is shown on the next page.

General Navigation (NPA25) - Edition 4 – 110630 8-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example: You are approaching a runway, QDM 042°(M). The surface wind is 270°(M) at
21kts. What are the head/tailwind and crosswind components?

Mark the wind point in the usual way, then set the runway QDM of 042° (M). Decide first on the
direction, remembering that the wind blows from the centre circle to the wind point. You can see
clearly that you have a tailwind, and a crosswind from the left. Now reset the compass rose to the
runway QDM reciprocal of 222°(M) and read off the windspeeds. You should have answers close
to 14 kt tailwind and 16 kt crosswind from the left.

8-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 9 Pilot Navigation Techniques (Part 1)
A modern well-equipped aircraft may carry some very sophisticated area navigation equipment to
enable the pilot to define the aircraft’s position with great accuracy. The CPL and ATPL also allow
the holder to fly light aircraft with no navigation assistance and to earn money for any instruction
the holder is qualified for. Therefore these techniques are still useful. Computers have also been
known to fail, so it is important for a pilot to be able to carry out a few simple, basic procedures to
enable the aircraft to reach its destination safely. Although these procedures are not very precise,
they are adequate for many types of flight under normal circumstances, especially in light aircraft
outside controlled airspace. They are largely based on estimates using the ‘One in Sixty Rule.’

1 The ‘One in Sixty Rule’.


The ‘One in Sixty Rule’ is based on the very simple fact that, at a range of 60 nm, an angle of 1°°
subtends an arc of about 1nm.

Figure 9-1 1 in 60 Rule (1)

In the diagram above, the tangent of the angle at A is CB over AB.

tan A = CB so CB = AB tan A
AB

But tan 1° is 0.017, so if AB is 60 nm, CB = 60 x 0.017 = 1.02 nm

There are, therefore, small errors in assuming that at a range of 60 nm CB will be 1 nm, but they
may usually be ignored unless a high degree of accuracy is necessary.

If AB remains at 60 nm but angle A is increased to 2°, the side CB becomes approximately 2nm. If
the angle is increased to 5°, CB becomes about 5 nm, and so on. The errors in this assumption
gradually increase, becoming about 5% at 10°.

If the distance AB is doubled to 120 nm, CB becomes about 2 nm. See the next diagram.

Figure 9-2 1 in 60 Rule (2)


General Navigation (NPA25) - Edition 4 – 110630 9-1
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
We can reduce these calculations to a simple formula which fits most circumstances. In the above
diagram, to find the side D we can use the formula:

R x A = 60 x D

where R is the range and A is the angle. R and D must be in the same units, and A is the angle in
degrees.

Example: A ground radar unit finds that an aircraft is at a range of 75 nm and at an angle of 3°
above the horizontal. What is the height in feet of the aircraft above the radar horizon?

R x A = 60 x D, so 75 x 3 = 60 x D D is therefore (75 x 3) ÷ 60 = 3.75 nm.

The answer is required in feet, so 3.75 x 6080 = 22 800 ft.

Example: A pilot notes that a landing ground is at a range of 15 nm. If he is now flying at
4500 ft above his landing point, what continuous angle of descent must he maintain to touchdown?

First, convert 4500 ft to nm by dividing by 6080: answer 0.74 (although ¾ is probably quite
accurate enough.)

R x A = 60 x D, so 15 x A = 60 x 0.74 ∴ A = (60 x 0.74) ÷ 15

Answer: 2.96°° (If this type of calculation is required in an examination the ‘correct’ answer would
almost certainly be 3°)

2 Tracking Errors and Corrections

2.1 Track Error

One of the most common uses of the 1 in 60 rule in pilot navigation is to solve the problem of what
to do when the pilot finds himself ‘off track’ during a flight. In the next diagram, the aircraft is
planned to fly along the track AB, but after a period of time the pilot has found his position to be at
C. The angle between the ‘Required Track’ (or ‘Planned’ or ‘Desired’ Track) and the actual track
(usually known as the ‘Track Made Good’) is the Track Error (TE). It is simply calculated using
the 1 in 60 rule:

Example: A pilot has flown for 12 minutes at a GS of 120 kt. He then finds himself to be 2 nm left
of track. What is his track error?

9-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 9-3 Track Error Angle

The aircraft has flown for 12 minutes at 120 kt, so it has covered 24 nm. In the formula
R x A = 60 x D, R is 24 nm and D (the distance off track) is 2 nm.

∴ 24 x A = 60 x 2. So A = (60 x 2) ÷ 24 Answer: 5°° Port

Note: Do not confuse track error angle with drift. Drift is the angle between heading and
track, whereas track error angle is the angle between the required track and the track
actually flown (‘Track Made Good’ or TMG).

2.2 Regaining Track

Having discovered that he is off track, track may be regained in the same time it took for the error
to arise. Consider the same aircraft as in the problem above, having arrived at point C and
calculated that the TE is 5°Port.

Figure 9-4 Regaining Track

On reaching point C the pilot turns right by double the track error angle (in this case 10°) to regain
track at point D. Since it took 12 minutes to fly from A to C, track should be regained after another
12 minutes. When point D is reached the aircraft will need to be turned left again, this time by the

General Navigation (NPA25) - Edition 4 – 110630 9-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
track error angle of 5°. This should leave the aircraft on a heading 5° further right than the original
heading to correct for the wind error which blew the aircraft off the required track in the first place.

Note that this technique only works if the fix position C is less than half way along the track to the
destination E.

2.3 Altering Heading for the Turning Point.

If it is not necessary to regain track before the turning point, an alteration of heading direct to the
turning point may be sufficient. Consider the diagram below.

Figure 9-5 Alteration of Heading for the Destination

If the aircraft has arrived at C and if the pilot turns right through the 5° of TE we found, the aircraft
would then fly parallel to the required track along CD. A further right turn through the ‘Closing
Angle’ would take the aircraft direct to the destination E. If the distance BE is known, because we
have already found the distance off track BC, we can find the closing angle using the 1 in 60 rule.
Suppose the remaining distance to the destination BE is 40 nm. In our formula:

R x A = 60 x D, D is 2 nm (found previously) and R is 40nm

So 40 x A = 60 x 2 ∴A = (60 x 2) ÷ 40 Answer: 3°°

Note that this is the closing angle. The total angle through which the aircraft must be turned to
reach the destination is the closing angle plus the track error angle, in this example 8°°.

3 Other Common Applications of the 1 in 60 Rule


Other popular uses of the 1 in 60 rule is for the calculation of rate of descent (RoD) on a glideslope,
the measurement of the height of cloud tops using a weather radar display, or estimating distance
from a small change in bearing.

9-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3.1 Aircraft Rate of Descent (Rod)

Example: What RoD is required to maintain a 2.5° glide slope at a groundspeed of 140k?

Figure 9-6 Rate of Descent Calculations

Consider the diagram above. The aircraft is 1 minute from touchdown, and during its final 1 minute
in the air it must travel a forward distance R for 1 minute at 140 kt. Since 140 kt is 140 nm per
hour, in one minute it will travel (140 ÷ 60) nm. However, since we measure the RoD in feet per
minute, we have to convert this distance to feet by multiplying by 6080. (We will use 6000, which
is close enough for this purpose.)

In our 1 in 60 – rule formula R x A = 60 x D, therefore, the range R is (140 ÷ 60) x 6000.


The aircraft must descend through the vertical distance D (in the diagram) in 1 minute, which is the
RoD we want to find.

Rearranging the formula to: D = (R x A) ÷ 60, we now have:

RoD = GP Angle x Groundspeed in knots x 6000 100


60 60

We can cancel as shown above, leaving the formula in this form:

RoD = GP Angle x Groundspeed in knots x 100


60

= (2.5 x 140 x 100) ÷ 60

= 583 ft/min

General Navigation (NPA25) - Edition 4 – 110630 9-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3.2 Change of Rate of Descent

Another problem sometimes encountered is the change of rate of descent required if the
groundspeed changes. The rule is: Increase Groundspeed = Increase RoD
Decrease Groundspeed = Decrease RoD

The above formula can be used for this problem too, simply substituting Change of RoD and
Change of Groundspeed in the formula.

Example: During a landing approach procedure at RAF Leuchars, which has a glideslope of
2.5°, an aircraft experiences an increased headwind of 20 kts. How will the Rate of Descent
change?

Change of RoD = GP Angle x Change of Groundspeed in knots x 100


60

= 2.5 x 20 x 100
60

= 83 fpm Decrease.

3.3 Approximate Range to a Beacon or Point of Land

It must be stressed that this technique only works when tracking approximately 90° to the beacon’s
bearing and when the two bearings taken are within about 15° of each other. If the bearing to a
VOR or NDB is given at two points along a track when the beacon is abeam the track, then the 1-in-
60 rule can be used to estimate the range to the beacon. In this case for every degree of bearing
change the ground distance covered is about 1/60 of the range to the beacon.

In the formula previously used, R x A = 60 x D, R is the range to the beacon, A is the bearing
change between the two fixes and D is the ground distance run in the time between the two fixes.

Example: Flying on a track of 010°M at groundspeed 150 kts you take two fixes on a VOR.
The first fix is 092° from the aircraft, and the second fix 6 minutes later is 100°. What is the
approximate range to the beacon?

The change in bearing is 8°. At 150 kts (2½ miles per minute) the distance moved across the ground
in 6 minutes is 15 nm.

R x A = 60 x D, so R x 8 = 60 x 15 ∴R = (60 x 15) ÷ 8

Answer : 112.5 nm

Exactly the same technique can be used to calculate range from a point of land, given radar
information and a similar geometry.

Example: Your aircraft is flying on a track of 360° (M) at a groundspeed of 240 kt, and you
note that the bearing of a radio beacon is 088° (M). Four minutes later you notice that the bearing
has changed to 094° (M). Use the 1 in 60 rule to estimate your range from the beacon.

9-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
You are flying at 240 kt which is 4 nm per minute, so your distance flown in the 4 minutes of your
observation is 16nm. During that time the bearing has changed by 6°
R x A = 60 x D, so R x 6 = 60 x 16 ∴R = (60 x 16) ÷ 6

Answer : 160 nm

3.4 Speed factor

Speed factor is defined as Speed/60 and can be used in a number of mental flight path calculations.

For example, if an aircraft is flying with a groundspeed of 120kts, the speed factor would be:

120/60 = 2

For example, to calculate the wind correction angle (WCA) given the current crosswind component,
the following formula can be used:

WCA = Crosswind component


Speed factor

For an aircraft flying with a groundspeed of 120kts, with a crosswind component of 10kts, the wind
correction angle would be:
10 = 5 degrees
2
Another useful formula is:

Rate of descent = Speed Factor * glidepath angle * 100

For an aircraft flying with a groundspeed of 120kts (speed factor = 2), the rate of descent required
to maintain a 3 degree glideslope would be:

2 * 3 * 100 = 600fpm

Compare this to the alternative formula which works for a 3 degree glideslope:

ROD = G/S * 5 = 120 * 5 = 600fpm

4 Navigation in climb and descent


This topic is covered fully in the Aeroplane Performance and Flight Planning modules but key
items are included in this module as Pilot Navigation techniques.

4.1 Airspeed to be used in climb and descent

Aircraft climb schedules are usually given as IAS/CAS and/or Mach numbers. When an aircraft
climbs at a constant CAS as per the climb schedule, the TAS will increase with altitude. TAS is
required in order to find groundspeed and therefore leg times – so what TAS should be used in a
climb?

General Navigation (NPA25) - Edition 4 – 110630 9-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
The technique used in this case is to calculate the TAS that would be achieved at the climb CAS at
an altitude 2/3 of the way up the climb.

As we are using TAS 2/3 of the way up the climb, we also use the wind velocity at this altitude for
any calculations affecting the climb segment.

An aircraft is usually descended at a constant IAS/CAS also, so the same problem exists – what
TAS and wind velocity should be used? In a descent, the technique is to use the TAS and the wind
velocity at an altitude 1/2 way down the descent.

4.2 Other useful climb/descent techniques

The speed at which an aircraft travel vertically when climbing or descending – the rate of climb or
rate of descent - is usually stated in feet per minute (fpm).

When comparing the vertical distance travelled in a unit of time to the horizontal distance travelled
in the same unit of time, a climb/descent gradient or climb/descent angle can be calculated.

A climb/descent gradient is an expression of the vertical distance travelled as a percentage of the


horizontal distance travelled.

A climb/descent angle is a measurement in degrees of the aircraft flight path from the horizontal.

Some useful formulae when tackling problems related to climb and descent are shown below:

Rate of climb or descent (fpm) = Gradient (%) * Groundspeed (in Kts)

Climb or descent gradient (%) = altitude change (in feet) * 100 / ground distance covered (ft)

Climb or descent angle (degrees) = TAN-1( altitude change (ft) / ground distance covered (ft) )

9-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Pilot Navigation Techniques (Part 2)

5 Topographical Aviation Mapping


One of the essential skills which must be possessed by a pilot is the ability to navigate visually.
Although most modern aircraft are equipped with sophisticated navigation equipment which is both
reliable and accurate, a pilot must still be able to cope with navigation in the unlikely event of
equipment failure. A range of maps specifically designed to be suitable for visual navigation are
produced by aviation authorities, and one of the most common is the ICAO 1: 1,000,000
topographical map, the ICAO World Aeronautical Chart. Most UK GA pilots are more familiar
with the CAA 1:500,000 charts. In Flight Planning you will be using the ED6 1:515,000.

5.1 Scale.

It is important to use a map which is of a suitable scale for the task. The scale of a map is defined
as the ratio of the distance on the map to the distance on the earth:

Scale = Chart Length


Earth Length

For example, a map of scale 1 : 5,000,000 or 1


5,000,000
means that 1 cm on the map is equal to 5,000,000 cm on the ground. This method of representing
scale is known as a ‘Representative Fraction’. The top figure in this fraction is always 1.

Other methods of scale representation may be used, such as the graduated scale:

Figure 0-1 Graduated Scale

Alternatively the scale may be expressed in words, such as “1 inch to one statute mile”. This would
often simply be described as a “one-inch map”.

General Navigation (NPA25) - Edition 4 – 110630 9-9


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5.2 Scale Problems.

It is occasionally necessary to answer scale problems in examinations. These can take a number of
forms, but the simpler problems are shown below.

Example: On a 1:250,000 map, how long a line in cm would represent 15 nm on the ground?

Scale = CL So ___1___ = CL
EL 250,000 15nm

∴CL = 15 x 1852 (conversion to metres) x 100 (conversion to cm)


250,000

Answer: 11.11cm.

Example: What is the chart scale of a 1 inch to the statute mile map?

Scale = CL So scale = ___1 inch___ = _______1 inch_________


EL 1 statute mile 5280 (feet) x 12 (inches)

= __1__
63,360

Some maps and charts vary in scale from one place to another on the map, so it is very important
for the pilot to be well aware of the properties of the particular map he is using. If the map is
‘Constant Scale’, a graduated ruler may be accurate enough for distance measurement. Where there
is scale variation over the chart, the technique for measuring distance will be covered later.

6 Topographical Features
One of the most important uses of a topographical map is to enable the pilot to determine the lowest
altitude at which it is safe to fly. Topographical features are represented in various different ways:

a) Maximum Elevation Figure (MEF). The MEF is printed on the chart in the quadrangle bounded
by the lat & long graticule every half degree. It is based on the best information available on the
highest known feature in the quadrangle, including both terrain and obstacles and an allowance
for features which are unknown (which might be up to 299’ high) by adding 300’ to any ground
elevation, although not to marked obstacles. It is printed as a two-figure group, the first (large)
figure representing thousands and the second (small) figure hundreds of feet above mean sea
level. It is not a safe height to fly.

b) Spot Elevations. Important hilltops are shown and marked in feet, and the elevation is printed
inside a white rectangle with a black border.

c) Contours. A contour is a line which joins all places of the same elevation above sea level. The
intervals at which they are drawn varies from chart to chart, and they may be marked in either
feet or metres. It is vitally important that you are familiar with the system employed on the
map you are using. Closely-spaced contours indicate steep slopes, whereas widely-spaced
contour lines are an indication of gentle gradients.

9-10 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
d) Layer Tinting. Ground elevation is usually indicated by a change in the colour used on the map.
The darker colours are used for higher ground, and a key is printed on the map border.

7 Reading the Map


Pre-flight preparation and planning is always time well-spent, and the pilot should carefully
consider which features will stand out during the flight. Roads and rivers can often be difficult to
see, especially in well-wooded areas. Railway lines often stand out well because of their relative
straightness. Woods are sometimes very useful, but the marking of wooded areas on a map may
well become outdated very quickly as trees are cut down and new plantations grow. Map reading is
the ability to accurately relate what is printed on the map to what is seen on the ground, and doing it
well takes practice.

7.1 Map-to-Ground.

The normal method of visual navigation is to look at the ground features printed on the map which
are close to the planned track, decide what is likely to be recognisable and within visual range, then
look out of the aircraft to find them. This is known as navigation by Map-to-Ground.

7.2 Ground–to-Map.

From time to time, all pilots become uncertain of their position. In these circumstances it is usually
best to look for some major ground feature within visual range, then find it on the map. This is
known as Ground-to-Map.

7.3 Major Features.

i. Water Features.
The most prominent features are usually the best for navigational use. Coastlines are always
very useful, and it is almost always quite easy to fix the aircraft position by using them. Other
water features are also reliable if they are large enough and unique enough. Navigation in the
UK using lakes and reservoirs is often easily possible, whereas trying to use lakes and rivers
in many parts of Canada is very risky because there are so many of them.
ii. Roads and Railways.
Railways and motorways are usually large and distinct enough to be very useful for visual
navigation. Ordinary roads are often too numerous for reliable use in areas of high population
density, whereas a rough track over an otherwise desolate desert landscape is an excellent
navigational feature.
iii. Towns and Cities.
Towns are often too large to be used for precise navigation, but a part of a town, or a major
feature within that town (York railway station, for instance,) may be very useful when map
reading.
iv. Vertical Features.
Any large feature which has a significant vertical extent is likely to be useful, again provided
that it is sufficiently unique to be unambiguous. (Blackpool Tower?) Hills, monuments set
on hilltops, ranges of hills etc., can be very helpful, and can often be identified at long ranges.

General Navigation (NPA25) - Edition 4 – 110630 9-11


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
v. Snow and Ice.
Snow, ice, and other seasonal changes can play havoc with a pilot’s ability to read a map.
Many features become indistinguishable from their surroundings after a heavy snowfall, and
small lakes and reservoirs tend to disappear when they are frozen over then covered in a layer
of snow. Snow-covered ground combined with a grey-white sky is also extremely dangerous,
resulting in a condition known as ‘white-out’, in which the horizon is lost and the terrain
clearance is very difficult to judge visually. Flooding can also make for difficult visual
navigation as rivers and roads can disappear under water.

8 Visual Check Points.


Easily identifiable points on or close to the planned track should be chosen at the planning stage,
separated by about 5 to 10 minutes at the groundspeed. A pilot has more to do in the air than just
navigate. The visual check point is intended to be found and identified easily enough for the pilot
to be able to concentrate on monitoring the aircraft systems, fuel state and lookout, but then to look
out and identify the check point without having to try too hard. Ideally these points should be on
track so that an accurate fix can be obtained, but it is more important that they points should be
easily and accurately identified, even if it is not possible to fly overhead. Night navigation can be
quite difficult, and a completely different set of visual check points will probably have to be chosen
for a flight after dark. Motorways (if they are busy) show up very well at night, but unlit natural
features may be impossible to see.

9 Lost Procedure.
It is not difficult to get lost. As soon as you have any doubt about your position, follow the logical
‘Lost’ procedure to resolve the uncertainty as soon as possible.

a) Check airspeed and, especially, heading. Carry out a compass comparison if you have more
than one compass in the aircraft. Heading errors are the most common cause of becoming lost.
b) Use all possible navigation aids to establish your position.
c) Climb to increase both visual and radio range.
d) Establish a DR (dead reckoning) position based on the last time you were certain of your
position. Use the best track and groundspeed you have available in calculating your DR
position.
e) Draw a ‘circle of uncertainty around the DR position. Make the circle of radius 10% of the
distance flown since the last known fix, and look for a prominent feature which you can see on
the ground and which you think you are likely to be able to find on the map, looking within the
plotted circle of uncertainty.
f) If you still have difficulty, and fuel and other considerations permit, turn towards a line feature
and follow it until you reach a visual check point which you can identify.

If you still find difficulty in identifying your position, bearing in mind your fuel state and weather
conditions, set the transponder to ‘Emergency’ (squawk 7700) and call the Air Traffic emergency
service on 121.5 MHz.

9-12 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section Three Charts and Plotting

Chapter 10 A Flat Earth?


Navigational charts are representations of an almost spherical earth on a flat surface. If we imagine
the problem of trying to flatten a hemisphere onto a table, it is clear that it cannot be done without
some stretching and distortion. All flat paper charts therefore have some degree of distortion built
into them, and various methods have been tried to minimise the effects of this flattening process.

1 Chart Projections.
The usual way of drawing a navigational chart is to begin by preparing a ‘Reduced Earth’. This is a
scale model of the earth, and the scale chosen is that of the finished chart. For instance, if a 1 :
250,000 scale chart is required, the reduced earth will be at the same scale. A light source is then
placed at some point inside or outside the reduced earth, and the latitude and longitude graticule is
projected onto a sheet of paper. Charts produced by this method are known as ‘Perspective Charts’.

Most charts today are produced by computer models rather than light projection, and they are
known as ‘Non-perspective charts’. The differences may not always be obvious to the user.

2 Scale Factor.
When the paper chart and the reduced earth touch, the scale on both will be the same.

Scale Factor (SF) = ____Chart Length___


Reduced Earth Length

Where the scale factor is 1, the scale of the chart is ‘correct’, meaning that it is the same as the
nominal scale printed on the chart.

3 Types of Projection.
There are three basic types of projection:

a) Plane, or Azimuthal.
b) Conic.
c) Cylindrical.

4 Plane Projections.
The ‘Plane Projection’ type is produced when a flat sheet of paper is placed against a point on the
reduced earth as shown in the next diagram. The light source in the projection illustrated projects
the graticule onto a sheet of paper where the scale factor is 1 at the South Pole. The scale expands
as the chart distance from the pole increases.

General Navigation (NPA25) - Edition 4 – 110630 10-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 10-1 Plane Projection

Figure 10-2 Plane or Azimuthal Graticule

10-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5 Conic Projections
Conic projections are produced by placing a cone of paper over a reduced earth. A perspective
conic projection is illustrated below.

Figure 10-3 Conic Projection

The paper cone is placed over the reduced earth, then the cone is ‘developed’ (opened out) as shown
below, producing a conical graticule.

Figure 10-4

General Navigation (NPA25) - Edition 4 – 110630 10-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
6 Cylindrical Projections
The third type of projection in the JAR syllabus is one produced by wrapping a sheet of paper
around the reduced earth in the form of a cylinder. The one illustrated below is a perspective
geometrical cylindrical projection.

Figure 10-5 Geometrical Cylindrical Projection

Note how the scale expands as the position on the chart moves further away from the equator. This
produces distortions as shown in the next illustration.

Figure 10-6 Geometrical Cylindrical Projection

10-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
7 Chart Properties

7.1 The Ideal Chart.

The perfect chart for all purposes does not exist. Ideally, our charts would possess the following
features for accurate representation of the earth’s surface:

a) Shapes would be correct.


b) Areas would be correct. (Known as ‘Equivalence’.)
c) Scales would be correct, and constant over the whole chart.
d) Angles would be correct.

For navigation, the ideal chart would also have the following characteristics:

a) Great circles would be straight lines.


b) Rhumb lines would be straight lines.
c) Adjacent sheets would fit together.
d) World-wide coverage would be possible.
e) Plotting of positions would be straightforward.

All these properties can only exist together on a globe. Note that:

• It is never possible on the whole of any chart to have shapes correct, or for the scale to be
everywhere constant and correct.
• It is possible to have either angles correct or areas correct, but it is never possible to have both
angles and areas correct on the same chart.

By using non-perspective projections (i.e., mathematically modifying the chart), it is possible to


come acceptably close to some of these ideals on a single projection. Over small areas, for instance,
some charts will have an almost constant scale and represent shapes reasonably accurately, but this
cannot be done over large areas on one projection.

7.2 Orthomorphism/Conformality

The primary requirement for a navigational chart is that it should be ‘Conformal’ or ‘Orthomorphic’
– note: these two words are interchangeable when referring to charts.

Definition: An orthomorphic or conformal chart is one in which angles are correctly represented at
every point on the sheet. This will mean that:

• Meridians and parallels will cross at right angles.


• At any given point, either scale will be the same in all directions, or it will change at the same
rate in all directions.
When a chart is drawn so as to be orthomorphic or conformal, it will possess additional important
properties:
• Rhumb lines will always be on the equatorial side of great circles.
• Small shapes will be reasonably correct, but larger shapes will be distorted.
• Areas will not be correctly represented.

General Navigation (NPA25) - Edition 4 – 110630 10-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

10-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 11 The Mercator Projection.

1 Description
The Mercator chart is a non-perspective cylindrical projection derived from the geometrical
cylindrical projection illustrated in the previous chapter. It is mathematically modified to make it
conformal. The basic chart, known as the Normal or Direct Mercator Chart consists of a cylinder
wrapped around the earth at the equator, but two other Mercator projections are also in common
use.

1.1 Properties of the Normal Mercator Chart.

Yes, achieved by mathematical modification of a basic cylindrical


ORTHOMORPHISM
geometrical projection.
EQUIVALENCE No.
SCALE Expands from the equator according to the secant of the latitude.
GREAT CIRCLES Curves concave to the equator (or convex to nearest pole)
RHUMB LINES Straight lines.
MERIDIANS Parallel straight lines, which cut all parallels of latitude at 90degrees.
PARALLELS OF LATITUDE Straight lines, parallel to the equator.
GRATICULE Rectangular.

This is one of the most important of all navigational charts, although its use has reduced
considerably in recent times.

Note that the meridians do not converge on this chart. Although the departure is decreasing with
increasing latitude, (Remember the formula: Departure = ch. Long x 60 x cos lat) the distance
between the meridians drawn on the chart is not decreasing. The scale in an E-W direction is
therefore increasing with latitude. In order to keep the scale expansion the same in all directions,
(in order to maintain orthomorphism,) the distance between the parallels of latitude must also
increase as the latitude increases.

1.2 Disadvantages of the Mercator Projection.

a) Apart from the meridians and the equator, great circles are complex curves. Radio bearings
(which are great circles) have to be converted to rhumb lines before they can be plotted.
b) The projection is unusable at high latitudes (North of about 75°N or S) because of the increasing
scale expansion. It is impossible to project the poles on a standard Mercator chart.
c) The scale is not constant, so great care is required in the measurement of distances. Whenever a
track has a N/S component, it must be divided into parts of no more than about 100nm for the
purpose of distance measurement. Each part of the track must then be measured, using the
latitude scale at that latitude to measure the distance.

1.3 Advantages of the Mercator Projection.

a) Rhumb line tracks are straight lines and are very easy to follow, since a constant track may be
flown. For short distances the difference between the great circle and rhumb line distance is
negligible.
b) Rectangular graticules makes the plotting of bearings and positions easy.
c) Sheets fit together if the meridian spacing is the same.

General Navigation (NPA25) - Edition 4 – 110630 11-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.4 Uses of the Normal Mercator Projection.

The primary use of the basic Mercator chart is for plotting up to latitude of about 75°N or S.

Figure 11-1 The Mercator Graticule

1.5 Great Circles and Rhumb Lines.

As stated above, rhumb lines are straight lines and great circles are concave to the equator. The
rhumb line is always on the equatorial side of the great circle, and the angle between the two at the
extremities of the leg is the conversion angle. (½Ch. long x sin mean lat., as covered in an earlier
chapter.)

The illustration below shows a rather extreme case of a very long E-W track at about 50°N, and
although 50°N is well within the useful latitude of the basic Mercator, another type of chart would
be probably be chosen for this type of flight. The difference between the rhumb line and great
circle tracks close to the equator, however, is very small, and the Mercator is an almost ideal chart
for use in equatorial regions. Between the latitudes of 8°N & S a great circle may be assumed to be
a straight line and the scale may be assumed to be constant.

11-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 11-2 Great Circle and Rhumb Line on Mercator Chart

1.6 Scale Problems on the Mercator Chart.

As stated in the previous chapter, the scale on a Mercator chart is the same at the equator as that of
the reduced earth on which it is based, because reduced earth and paper cylinder touch at that point.
To find the scale at any other point on the chart, however, it is necessary to multiply the equator
scale by the secant of the latitude. Remember that the secant is . 1 .
cosine

Example: The equator scale of a Mercator chart is 1:250,000 . Find the scale at 55°N.

We have already seen that: Scale = Chart Length


Earth Length
And, in our example,

Scale at equator = Chart Length at equator = 1 .


Earth Length at equator 250,000

But:
Scale at 55°N = Chart Length at equator x sec 55°
Earth Length at equator
So:
Scale at 55°N = 1 x sec 55°
250,000

General Navigation (NPA25) - Edition 4 – 110630 11-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
= 1 x 1 = 1
250,000 cos 55° 143,394

Example: The scale of a chart is 1 : 200,000 at 60°N. Find the scale at the equator.

Scale at 60°N = Scale at equator x sec 60°

1 x 1 = Scale at equator
200,000 sec 60°

= 1__ .
400,000

There are, however, simpler ways of doing these Mercator problems by remembering the formula:

___A_ = _ B__
Cos. a° Cos. b°

Where the scale at latitude a° = _1_


A

And the scale at latitude b° = _1_


B

In other words, remembering that the numerator of the scale fraction is always 1, the
denominator of the scale is A at latitude a° and B at a latitude of b°.

Example: Given the scale at the equator of 1:500,000 calculate the scale at 50°N.

Assume the equator to be latitude a° and 50°N to be latitude b°.

∴ 500,000 = __B__
Cos 0° Cos 50°

So B = 500,000 x Cos 50° = 321,393


Cos 0°

The scale at latitude 50°N is therefore: 1:321,393

11-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example: Your chart scale at 45°S is 1 : 100,000. Find the scale at 55°S.

Let lat A° be 45° and lat B° be 55°S. Substituting in the formula

___a_ = _ b__
Cos A° Cos B°

100,000 = ___b__
Cos 45° Cos 55°

∴ b = 100,000 x Cos 55°


Cos 45°

= 81116

So the scale at latitude 55°S = ___1___


81,116

1.7 Problems Involving Departure

Some Mercator scale problems involve the use of the departure formula

Departure (in nm) = Ch. long x 60 x Cos. Lat

Example: The scale on a Mercator Chart at 50°N is 1 : 3,000,000. Find the length in inches of
a straight line drawn from 58°N 4°W to 58°N 2°E.

Scale = Chart Length so CL = Scale x EL


Earth Length

So in order to find the CL in our problem we have to first find both the EL at the correct latitude
and the Scale, also at the correct latitude.
i. Earth Length
Substituting in the departure formula:

Departure = [6° (ch long) x 60 (to convert to minutes)] x Cos. 58°

= 191 nm
ii. Scale
Now we must find the correct scale at 58°N. Let Lat ‘A’ be 50°N and Lat ‘B’ be 58°N.
b = a Cos. B = 3,000,000 Cos. 58° = 2,473,224
Cos. A Cos. 50°

So the scale at 58°N = ___1____


2,771,658

But CL = Scale x EL
General Navigation (NPA25) - Edition 4 – 110630 11-5
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
= ___1____ x {191(nm) x [6080 x 12 (conv. to ins.)]}
2,473,224

= 5.62 inches

However this calculation is simpler remembering on fact about the Direct Mercator projection: all
the meridians are parallel. This means that on the chart the distance between two meridians is
constant regardless of the latitude. The distance can be calculated wherever it is easiest, in this case
at 50°N, where the scale is already known.

First calculate the Earth length between the correct meridians at this latitude, using the departure
formula.

Departure = [6° (ch long) x 60 (to convert to minutes)] x Cos. 50°

= 231 nm

Scale at 50°N is 1: 3,000,000 as given in the question.

Chart length = Scale x Earth Length

= ___1____ x {231(nm) x [6080 x 12 (conv. to ins.)]}


3,000,000

= 5.62 inches

Giving exactly the same answer.

11-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 12 Other Mercator Charts
Two other types of Mercator charts are in common use. They are constructed in the same way as
the standard chart we have so far considered, but the cylinder is wrapped around the reduced earth
in a different way.

1 The Transverse Mercator.


The Transverse Mercator is a variation of the standard Mercator chart. Instead of the chart being
wrapped around the equator, it is wrapped around a selected meridian of tangency. This chart is
very popular for countries which have a large ch. lat but small ch. long. It is used as the basis for
the UK Ordnance Survey charts where the selected central meridian is 2°W, although in this
particular case the chart scale is adjusted so that it is accurate 180 km E & W of 2°W.

Figure 12-1 The Transverse Mercator Chart

The chart suffers from the same distortions as the standard Mercator, (see the illustration above) but
provided it is not used outside the 500nm band either side of the selected central meridian it is an
excellent chart. Note that, unlike the standard Mercator chart, the Transverse Mercator is capable of
showing the poles, and is frequently used for this purpose. The Polar Stereographic chart (covered
later) is more commonly used in polar regions, however.

General Navigation (NPA25) - Edition 4 – 110630 12-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.1 Properties of the TM Chart.

CONFORMALITY Yes.
SCALE Scale correct along the Central Meridian (CM), expands away from it.
EQUIVALENCE No.
Straight lines along, and at right angles to, the CM, but otherwise
GREAT CIRCLES
curved concave to that meridian.
RHUMB LINES Complex curves.

2 The Oblique Mercator


This is another variation on the standard Mercator projection. The chart cylinder is wrapped around
the reduced earth at an oblique angle, producing a rather peculiarly-shaped graticule. This chart
once again has only small scale expansion within 500nm of the ‘false equator’ where the chart and
the reduced earth touch, and great circles can be assumed to be straight lines within this area.

Figure 12-2 The Oblique Mercator

12-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2.1 Properties of the Oblique Mercator

CONFORMALITY Yes.
Scale can be considered constant within 500nm either side of the
SCALE ‘false equator’ (the great circle of tangency). It expands away from
that great circle.
EQUIVALENCE No.
Great circles are concave to the ‘false equator’ unless they coincide
GREAT CIRCLES with it or are at right angles to it. Great circles within 500nm of the
false equator can be considered to be straight lines.
RHUMB LINES Complex curves.

2.2 Uses of the Oblique Mercator.

The Oblique Mercator is in common use as a chart for great circle air routes, when the false equator
is aligned with the route. It may also be used as the basis for topographical maps of countries of
great length but little width, e.g., Malaysia. The false equator would be aligned with the
approximate oblique axis of the country concerned.

General Navigation (NPA25) - Edition 4 – 110630 12-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

12-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 13 Navigation – Plotting

1 The Manual Track Plot


There are various ways of plotting the aircraft’s position on a navigational chart. The most
common of the basic ways is the ‘Manual Track Plot’, which involves the marking of navigational
information such as positions and fixes on the chart, then using the information to update wind
velocities, calculate ETAs and maintain the aircraft’s position on or close to the planned track.
More sophisticated navigation equipment carries out this function automatically, when it is known
as the ‘Automatic Track Plot’.

1.1 Symbols.

The symbols in common use are shown here, and their meanings and use will be progressively
explained.

Figure 13-1 Plotting Symbols

1.2 Fixes.

A fix is a position of the aircraft, which is found from references outside the aircraft. It is not an
estimate of position, although it may not be very precise. Where navigation aids and aircraft
equipment enable continuous fixes to be obtained, ‘Rapid Fixing’ is said to be possible. Accurate
visual fixes obtained when an aircraft is positioned overhead a known ground feature are known as
‘Pinpoints’, and fixes obtained from navigation aids such as GPS would be marked on the chart
using the ‘Radar Fix’ symbol.

If fixes have to be obtained only from bearings, using such navigation aids as the ADF or VOR, that
is known as ‘Non-Rapid Fixing’.

General Navigation (NPA25) - Edition 4 – 110630 13-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.3 DR Positions.

A ‘Dead Reckoning’ (DR) Position is an aircraft position obtained by calculation, based on


information obtained only from data systems carried within the aircraft. Since it is normally
derived from a knowledge of the Hdg/TAS of the aircraft combined with the most accurate known
wind velocity, the accuracy of the DR position depends very much on the accuracy of the weather
forecast. Some automatic systems (such as military Doppler navigation equipment or inertial
navigation equipment) provide a continuous automatic track plot of the DR position.

The manual track plot required by the JAA syllabus is a very simple manual method of keeping a
‘Plot’ of the aircraft DR position, and by finding accurate fixes wherever possible, recalculating the
W/V to enable the pilot to navigate to the destination or turning point.

1.4 Local and Mean Wind Velocities.

A ‘Local’ W/V is that found in the aircraft’s immediate area. It is extremely useful as the basis for
navigation, but is often not particularly accurate because it is measured over a short period of time
and may be affected by local conditions. A ‘Mean’ W/V is measured over a much longer period
(anything from perhaps 12 to 45 minutes) so is usually much more accurate. However, because it is
the W/V which has affected the aircraft in the past, it is of limited use in determining what the W/V
will be on the route ahead.

2 The Track Plot – Plotting Exercise.


A common method of carrying out a track plot in a rapid-fixing area is to simply plot a series of
fixes on the chart at perhaps 3 – minute intervals. Track-keeping is carried out by visually assessing
the aircraft’s position in relation to the required track and altering the heading to stay on or close to
the track. The 1-in-60 rule may be used to regain track, but calculations are kept to a minimum.
ETA changes are estimated visually from the aircraft position in relation to timing marks on the pre-
prepared chart.

The manual track plot may also be used in non-rapid fixing areas and this is described below. A fix
is used to find an average wind over a period of around 30 minutes and the method shown is known
as finding a ‘Track & Groundspeed Wind’, using the CRP-5 navigation computer in a way that has
already been covered. The events which permit the wind calculation are logged on a ‘Navigation
Log’.

Event Track Hdg M.No/ Press


W/V Hdg (T) Var Observations TAS G/S Dist Time ETA
Time (T) (M) RAS Alt/OAT

Figure 13-2 The Navigation Log

Example: You are required to fly from position ‘A’ at 5230N 0230W to position ‘B’ at 5350N
0140E, flying at a press alt of 2000ft, Temp +12°C, CAS 140kt. The forecast W/V is 180/30. (We
will use this as shorthand for 180° and 30kt.)

This is all the information you need to begin the ‘Pre-Flight’ preparation. Carry out the work in the
following order:

13-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
a) Mark the start point and the destination on the chart, annotating the track line with the
conventional double arrow. Measure the track and distance, and record them on the Nav Log.
Take great care to measure the distance correctly, remembering that because of the scale change
on the Mercator chart the distances must be measured in portions at the latitude of that part of
the track..

b) Now calculate the TAS (from the given information of CAS, pressure altitude, temperature) and
groundspeed. Record in the log.

c) Next, mark the wind on the CRP-5 and calculate the required heading and resultant
groundspeed. Record in the log.

d) Now complete the first line of the log, assuming a take-off time of 1000hrs and taking the
variation at the chart value for position ‘A’ to calculate the Hdg (M).

Your figures should look something like this:

Event Track Hdg M.No/ Press


W/V Hdg (T) Var Observations TAS G/S Dist Time ETA
Time (T) (M) RAS Alt/OAT
1000 062 180/30 073 8W 081 A to B 140 2000 +12 145 156 171 66.2 1106.2

You are now ready to begin the ‘flight’.

Now let us suppose you obtain a radar fix at 1036 at a position 5305N 0000E/W. Record this in the
log and plot the fix on the chart.

At this point we have to assume that the pilot has been flying the correct heading and CAS and that
the forecast temperature is correct. Because the fix is not on track, the only other possible cause of
the track error is that the wind is not as forecast.

Draw in the Track Made Good (TMG) by joining the start point A to the fix, and measure the track
angle, recording it in the log. Measure the distance from A to the fix. This distance has been
covered in 36 mins, so calculate the actual groundspeed, recording it in the log. You should have
found that the TMG is about 069°(T) and the GS is about 162kt.

Now you have two sides of a new vector triangle. You have the original Hdg and TAS (which you
are still flying,) and a new Trk & GS. Use the CRP-5 to calculate the new W/V. Enter it in the log.
The nav log should now look something like this:
M.No
Event Track Hdg Press
W/V Hdg (T) Var Observations / TAS G/S Dist Time ETA
Time (T) (M) Alt/OAT
RAS
1000 062 180/30 073 8W 081 A to B 140 2000 +12 145 156 171 66.2 1106.2
1036 069 220/20 Fix 5305N 0000E/W 162

This wind calculation has all taken some time, during which the aircraft has continued to fly in the
‘wrong’ direction. It is now necessary to plan ahead (DR ahead) to a point from which a new track,
distance, heading and ETA can be calculated.

General Navigation (NPA25) - Edition 4 – 110630 13-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
.

Figure 13-3 Trk & GS W/V and DR ahead.

Extend the track line between A and the fix at 1036 for some suitable time ahead (in this example 6
minutes) at the new known groundspeed of 162kt to a DR position. From here, draw in a new track
to the destination B. Measure the new track and distance, fill in the log, then recalculate the
required heading and new ETA at B. Your log should now read like this:

M.N
Event Trac Hdg Hdg Press
W/V Var Observations o/ TAS G/S Dist Time ETA
Time k (T) (T) (M) Alt/OAT
RAS
2000
1000 062 180/30 073 8W 081 A to B 140 145 156 171 66.2 1106.2
+12
1036 069 220/20 Fix 5305N 0000E/W “ “ “ 162

1042 048 “ 049 7W 056 ∆ 5311N 0028E A/H B “ “ “ 166 59 21.4 1103.4

This is the complete track plot for this exercise.

13-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 14 Plotting Using Position Lines.

1 The Radio Magnetic Indicator (RMI)


Bearing information from ADF and VOR systems can be shown on the Radio Magnetic Indicator
(RMI), and also on some Horizontal Situation Indicators (HSI). The RMI is a very common
instrument and is frequently referred to in JAA examinations.

The RMI is provided with the following


facilities:

1.1 Rotating Compass Card

Aircraft heading from the compass system


is usually automatically fed to the RMI
compass card so that heading is shown at
the top of the indicator by a heading
index. On some RMIs, the card has to be
rotated by hand.

1.2 Bearing Needles

One or two bearing indicator needles with


arrow heads are provided, with a switch
for each needle to select the radio aid
which is to provide the bearing
information to that particular needle. If
two needles are provided one is narrow Figure 14-1 The RMI
and the other is wide, and the needles are
often coloured differently.

Selector switches are also provided, and are generally marked with the thin or thick arrow and
colour coded, with their selections labelled ADF1, ADF2, VOR1, VOR2, etc, where dual radio aid
systems are provided in the aircraft.

Some older bearing indicators have a fixed compass card, which indicates only the relative bearing
of the beacon from the aircraft. These indicators are called Relative Bearing Indicators (RBI), and
the QDM and QDR of the beacon must be calculated by the pilot.

The RMI needle shows the Magnetic bearing of the


beacon from the aircraft (the QDM) and also the
position of the beacon relative to the aircraft's nose.
The aircraft in the diagram is flying on a heading of
055°(M) and has tuned in an NDB or VOR. The RMI
is showing a bearing (QDM) of 105°(M). The needle
also shows that the beacon is on a bearing of 050°
relative to the nose of the aircraft. The aircraft is on
the 285° radial (QDR) from the beacon.

General Navigation (NPA25) - Edition 4 – 110630 14-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Plotting is usually done, however, with reference to True direction except in high latitudes. Where
it is not possible to use a magnetic compass because the earth’s magnetic field is too weak, gyro
steering and grid navigation is likely be used. Grid navigation techniques may be used at any
latitude, however, and will be described later.

The bearing to plot will be the reciprocal (±180°) of the bearing noted in the aircraft, but the plotted
bearing will often have to be corrected for variation differences between the positions of the aircraft
and the radio beacon. Meridian convergence will also have to be allowed for unless the aircraft and
the beacon have no significant ch. long between them, but this will be covered a little later.

2 Plotting the Bearings

Figure 14-2 Plotting a bearing as a Position Line

In the diagram shown here, the bearing of the radio beacon from the aircraft measured on the RMI
is about 075° (M) which has been converted to 065°(T). This is plotted on the chart as a position
line of 245°(T) from the beacon.

2.1 DME Ranges

Ranges from DME beacons are circular position lines and are plotted as an arc of a circle centred on
the DME position. The range may occasionally have to be corrected for the fact that the DME
measures slant range, whereas a plan range has to be plotted on the chart.

14-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 14-3 Plotting a Circular Position Line

3 Transferring Position Lines.

3.1 Transferring Straight Position Lines.

Figure 14-4 Transferring a Straight Position Line

Although it would be an advantage to take bearings on different beacons at the same time it is often
necessary to take bearings one after the other. In order to find a fix position it is then necessary to
‘Transfer’ one or two position lines to a later time, in effect holding them until they can be used in
combination with another position line. The lines are moved along the track in the direction of
flight by the best known groundspeed. Note the convention of the ‘double arrow’ to denote the
transferred line.

Occasionally it may be found that a position line does not cut the track in a sensible position. In
this case a suitable construction line is used for the transfer. The next diagram shows how this is
done.

General Navigation (NPA25) - Edition 4 – 110630 14-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 14-5 Transferring Using Construction Line

3.2 Transferring Circular Position Lines.

Circular (range) position lines are transferred by ‘moving’ the origin along in the direction of flight
parallel to the track.

Figure 14-6 Transferring Circular Position Lines

14-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4 Construction of a Fix by Transferring Position Lines.
It is possible, and often necessary, to construct a fix from two position lines. It is more usual,
however, and considerably more reliable, to construct a fix from 3 position lines.

Figure 14-7 Transferring Position Lines to Form a Fix

In the above diagram, three position lines have been taken at 6-minute intervals, and at angles to
each other deliberately chosen to provide a good ‘cut’. We normally aim to have an angle of at
least 30° between position lines to improve the accuracy of the fix. The 1000 & 1006 position lines
have been transferred forward by 12 and 6 minutes of groundspeed respectively, then combined into
a single ‘fix’ at 1012.

4.1 The ‘Cocked Hat.

Note that the position lines do not cross each other at a single point, forming a small triangle known
as a ‘cocked hat’. This is caused primarily by inaccuracies in bearing measurements, but will also
occur if the groundspeed used to transfer the lines is inaccurate – a very common situation. A large
cocked hat indicates that considerable errors may exist in the final fix position, and the fix must be
treated with caution. The fix position is usually accepted as being in the centre of the cocked hat,
but in some circumstances the relative accuracies of each of the position lines may be compared
when making a judgement about the final fix position.

5 Application of Variation when Plotting.

5.1 ADF Bearings.

ADF bearings read from an RMI will normally be magnetic because they are based on a magnetic
compass. In order to plot them, it is therefore necessary to convert them to True bearings by the
application of Variation. Since the bearing is measured at the aircraft, it is the variation at the
aircraft position which must be used to convert from magnetic to true direction.

General Navigation (NPA25) - Edition 4 – 110630 14-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5.2 VOR Bearings.

Although the presentation of a VOR bearing on an RMI looks identical to that of an ADF bearing,
the reading is produced in a completely different way. The VOR beacon is aligned with magnetic
North at the beacon position, and the aircraft’s radial is transmitted to the aircraft as a phase
difference which is only interpreted (not measured) at the aircraft. If it is necessary to convert the
bearing to true for plotting, the variation at the beacon position must be used, since it is the
beacon’s magnetic north on which the radial is based. Many navigators, however, either mark the
local magnetic north on the chart at the beacon position, or use the magnetic orientated compass
rose marked on the chart to define magnetic North. This enables them to plot the radial directly.

6 Conversion Angle on Mercator Charts.

6.1 ADF Bearings

You will no doubt remember that straight lines on Mercator Charts are rhumb lines. Radio waves,
however, travel over the earth as great circles, so bearings have to be corrected before they are
plotted whenever there is a significant difference between the great circle and the rhumb line
bearing..

Figure 14-8 Plotting ADF Bearings on Mercator Charts

In the diagram above, in which the conversion angle has been exaggerated for illustration purposes,
the ADF will measure a true bearing of 035°. If this were to be plotted from the beacon as a
reciprocal, the plotted bearing would be incorrect at 215°. In order to arrive at the correct answer of
235°, the conversion angle must be applied to the measured bearing.

Bearings measured in the Southern Hemisphere will of course also need correction, but since the
meridians converge in the opposite direction, the rhumb line will be on the opposite side of the great
circle. Conversion angle correction will therefore need to be applied the opposite way in the
Southern Hemisphere.

14-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
6.2 VOR and DF Bearings.

The same problem arises when plotting VOR and DF bearings on Mercator charts.

Figure 14-9 Plotting VOR & DF Bearings on Mercator Charts.

The radial or bearing received by the aircraft is a great circle bearing measured at the ground
station, so the bearing is in need of conversion angle correction before it can be plotted on a
Mercator chart.

7 Finding Conversion Angle.


An early chapter of these notes gave the conversion angle formula as:

Conversion Angle = ½ Ch. Long x Sin. Mean Lat

This is a formula which must be remembered, but in practice the conversion angle may often

Figure 14-10 The ABAC Scale.

be found more quickly from a scale printed on the side of the chart known as the ABAC Scale. In
the ABAC scale illustrated above, scale A must be used with scale C, and scale B with scale D. A

General Navigation (NPA25) - Edition 4 – 110630 14-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
ruler is placed across the scale to join the mean lat and the ch. long, and the conversion angle is read
off the centre scale. In the illustrated example, the mean lat is 50°, the ch. long is 16°, and the
conversion angle found is 6°. Other types of scale may be provided for specific charts.

8 Rules for Application of Variation and Conversion Angle.


Various correct methods may be taught for applying variation and conversion angle. This school
teaches the following:

1. Always apply corrections for variation and conversion angle where the bearing is measured,
i.e., at the aircraft for ADF bearings, and at the ground station for DF and VOR bearings.

2. If the above rule is followed, always apply the conversion angle in such a way as to move the
bearing line towards the equator.

Example: Your aircraft is positioned at 46°S 12°E, and you take an ADF bearing on an NDB which
is positioned at 44°S 20°E. The aircraft variation is 7°W and the variation at the beacon is 12°W.
The RMI reading is 084°(M). What true bearing would you plot on a Mercator chart?

First, apply the variation to correct the bearing. Since this is an ADF bearing, we use the variation
at the aircraft position. (The variation given here at the beacon is not needed in this calculation.)
With a variation of 7°W the bearing of 084°(M) therefore becomes 077°(T).

Next, find the conversion angle (CA). The mean lat in this problem is 45° and the ch. long is 8°.
Use the ABAC scale printed on the previous page to find the answer, which is just less than 3°.
Now apply the CA in accordance with rule 2 above. The bearing of 077° must be moved towards
the equator, and because the problem is set in the Southern Hemisphere, our bearing becomes
074°(T).

Finally, take the reciprocal of the calculated bearing to find the correct bearing to plot. Your
answer should be 254°°(T)

9 Relative Bearings.
A small number of aircraft are still fitted with Relative Bearing Indicators (RBIs) instead of the
RMI. This instrument, used in older ADF installations, has a fixed compass card, so gives bearings
measured relative to the aircraft’s nose. (If VOR Radials are presented on the same instrument,
which is very rare, the radials are read normally.) Bearings are calculated simply by adding the
aircraft’s true heading to the RBI reading to find the bearing of the beacon from the aircraft.

Example: Your aircraft is flying in the Southern Hemisphere on a heading of 144°(M) where the
variation is 3°E. You take a bearing on an NDB of 302°(R) on the RBI. If the CA is 4°, what true
bearing would you plot from the beacon?

First, apply the variation to the aircraft heading to find the Hdg (T), then add the Hdg and the RBI
reading together. (Alternatively, you could add the Hdg(M) and the RBI reading, then apply the
variation. The answer is the same.) Your answer is 144° + 3°E + 302° = 449°° Because the answer
is more than 360°, subtract 360° to give the answer of 089°.

14-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Now apply the CA to move the bearing towards the equator, remembering again that the problem is
set in the Southern Hemisphere. This changes the bearing to 085°, and the reciprocal which you
have to plot is 265°°.

10 Current Use of Mercator Charts.

The type of Mercator chart covered in this chapter is gradually falling into disuse because of an
ICAO Standard which requires that aviation charts should be selected on which a great circle
approximates to a straight line. In equatorial regions, however, (roughly from 8°N to 8°S) close to
where the reduced earth and the chart cylinder meet, the Mercator chart remains very popular
because scale expansion is very small and great circles can be assumed to be straight lines.

General Navigation (NPA25) - Edition 4 – 110630 14-9


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

14-10 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 15 The Lambert Conformal Conic Projection.
The Lambert Conformal Conic Projection is a non-perspective projection in which orthomorphism
is achieved by mathematical construction.

Figure 15-1 The Lambert’s Projection

Note carefully that this is not a normal conic projection. The cone passes inside the reduced earth at
the two standard parallels. This results in the scale being reduced between these two parallels,
precisely accurate only on the standard parallels, and expanded outside them.. The parallel midway
between the two standard parallels is known as the ‘Parallel of Origin’.

Figure 15-2 Lambert Chart – Area of Coverage.

A typical Lambert’s chart covers not more than 14° between the standard parallels and. illustrates
not more than one-sixth of the chart outside them. Provided that the chart construction is restricted
in this way, it can be assumed to be a constant-scale chart with a maximum scale deviation of less
than 1%. The greatest scale deviation occurs at the parallel of origin.

Note that not all Lambert’s charts are constant in scale. A number of special-purpose charts exist
which do not conform to the above rules, some of which are published by ICAO.

General Navigation (NPA25) - Edition 4 – 110630 15-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1 Properties of a Lambert’s Conformal Conic Chart.

1.1 The Graticule.

Figure 15-3 The Lambert Graticule

The graticule on a Lambert chart consists of parallels of latitude as arcs of circles which are centred
at the poles, and meridians which are straight lines converging at the pole and cutting all parallels of
latitude at right angles.

1.2 Conformality.

The Lambert’s chart is mathematically constructed to be orthomorphic.

1.3 The Parallel of Origin.

The parallel of origin of the chart may be assumed to be half-way between the two standard
parallels.

1.4 Rhumb Lines.

Meridians are straight lines, but all other rhumb lines are curves concave to the nearer pole.
Parallels of latitude are, of course, examples of rhumb lines. As in the Mercator chart, the angle
between a great circle and a rhumb line is the conversion angle.

1.5 Great Circles.

Meridians are straight lines. All other great circles are curves concave to the parallel of origin, and
a straight line joining points on the parallel of origin is almost exactly a great circle. For practical
use, however, great circles may be assumed to be straight lines on this chart. Remember that this is
an assumption because the difference between them is small.

1.6 Scale.

The scale is the least at the parallel of origin and expands away from it, becoming correct at the
standard parallels. The scale is greatest at the North and South extremities of the chart.

15-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.7 Chart Convergence and Conversion Angle.

Convergence of the meridians is constant on a Lambert’s chart, because all the meridians are
straight lines. You will remember the convergence formula:

convergence = ch. long x sin lat


When applied to a Lambert’s chart, the meridian convergence over the whole chart is calculated
based on the latitude at the parallel of origin (P of O). At latitudes greater than the P of O the earth
convergence will be greater than the chart convergence since the Sine of the earth latitude will be
greater than the Sine of the Parallel of Origin. At latitudes lower than the P of O the earth
convergence is correspondingly less than the chart convergence.

1.8 Constant of the Cone or Convergence Factor.

The sine of the P of O is printed on the chart, and is known as either the ‘Convergence factor or the
‘Constant of the Cone’.

1.9 Conversion Angle.

Conversion angle calculations on Lambert’s charts are always based on the latitude of the parallel
of origin whatever the position on the chart.

1.10 Summary of Properties of a Lambert’s Conformal Conic Chart.

CONFORMALITY Orthomorphic by mathematical construction.


Correct at the Standard Parallels. Scale is expanded outside them and
SCALE contracted between them. Provided that the 14° and one-sixth rules are
followed, the scale may be assumed to be constant over the whole chart.
EQUIVALENCE No.
Normally assumed to be straight lines, but in fact they are concave to the
GREAT CIRCLES
parallel of origin.
RHUMB LINES All rhumb lines except meridians are curves concave to the nearer pole.

2 Calculations.
Example:
A straight line is drawn on a Lambert’s chart from position A at 50N 020W to position B at 60N
015E. The direction of this straight line is 050º(T) at A. The P of O is 54º N. Determine the
direction of the track from B to A measured at B and the Rhumb line bearing of B from A.

The straight line track from A to B is 050º(T) when measured at A, but we are required to find the
track from B to A measured at B. In order to do this we need to find out the chart convergence
between the two positions.

Chart Convergence = ch..long x Sin P of O


From 20°W to 15°E = 35° x Sin 54°
Chart Convergence = 28.5º
Now draw this out to see clearly what is required.

General Navigation (NPA25) - Edition 4 – 110630 15-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Transferred
True North
North
28½°

60°N B

50° 180°
50°
50°N
A

020°W 015°E

Figure 15-4 Calculation Example

The dotted line is parallel to the meridian at A, so the angle between the dotted line and the
meridian at B is the chart convergence. Remember to take the reciprocal because the track is in the
opposite direction. The track from B to A is thus:

050° + 180° + 28½° = 258½°°


To find the rhumb line track, we need to know the conversion angle. The CA is 14¼°, half the
chart convergence between the two points. The rhumb line is always on the equator side of the
great circle, so the rhumb line bearing of B from A is

050° + 14¼° = 064¼°°

2.1 Track Measurement

Consider the aircraft in the diagram below flying from A to C via B. If the aircraft flies by the
shortest route, following the great circle track, the track required is constantly changing as it crosses
the converging meridians.

Figure 15-5 Track Angle Measurement on Lambert’s Charts.

15-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
You can see in the diagram that the initial track in the illustration is about 280°(T), whereas on
arrival at C the track will have changed to about 260°(T). In measuring the track for flight planning
purposes, therefore, the track angle should be measured at the mid-point B, which is the average
track for the leg. When departing from A, however, the track measured at A should be used to
calculate the initial heading.

Note that the change in direction from A to C is the same as the convergence between the meridians
at A and C.

3 Advantages and Disadvantages of the Use of Lambert’s Charts.

3.1 Advantages.

a) Straight lines may be assumed to be great circles, so it is easy to draw the line of shortest
distance between two points.
b) The application of conversion angle to radio bearings is unnecessary.
c) Distance measurement is easy if the chart is of a constant scale. Graduated rulers may be used.

3.2 Disadvantages.

a) Plotting positions requires care if accuracy is to be achieved.


b) Chart convergence must be applied to ADF bearings, or the aircraft’s meridian transferred to the
beacon position..
c) Constant heading change is required to stay on any track with an E/W component drawn as a
straight line on a Lambert’s chart.

General Navigation (NPA25) - Edition 4 – 110630 15-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4 Plotting of Tracks

The drawing of a great-circle track onto a Lambert’s chart is simple. On a Lambert’s chart a great
circle plots approximately as a straight line as long as the standard parallels are not much more than
14° apart (remember that the great circle is actually slightly curved, concave to the parallel of
origin. That curve increases if the standard parallels are chosen further apart).

When considering the course of a track line, the position at which the course is to be measured must
be considered. Unlike Mercator charts Lambert charts have converging meridians, therefore the
course will depend to a certain amount on the meridian from which it is measured. In most cases the
mean course will be required, and should be measured by aligning the protractor with the central
meridian.

Consult figure 17-1 where the course from A to B is 079°T as measured against the central
meridian. The protractor shown on the diagram is aligned with the mid meridian and demonstrates
this. However if the course was measured from the meridian at A, for example if the initial course
was required, then that would be 075°T.

x B

A
x

Figure 15-6 Track plot

The mean track A to B is 079°T, but the initial track is 075°T.

15-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5 Plotting of Position Lines

Plotting on a Lambert’s chart is generally more straight-forward than it is on a Mercator chart.


Radio waves follow great circles, and as discussed these can be plotted as straight lines on a
Lambert’s chart with little error. Therefore the conversion angle is not used; rhumb lines are never
plotted. The only additional consideration is convergence when plotting a radio bearing from an
NDB.

5.1 Plotting from a VOR

The bearing measured by VOR equipment will usually be given as a reading from an RMI. This is a
reading from the aircraft to the VOR, so the reciprocal must be taken to find the radial. If given as a
reading from an OBS it could be TO or FROM; if TO then again the reciprocal must be taken.

The bearing is magnetic so there are then two ways of plotting. The variation (isogonals are found
as ticks around the edge of the E(LO)1 chart used in the exam) can be applied, and the protractor
lined up with the meridian closest to the VOR. A simpler method is to line the protractor up with
the magnetic North tick that is printed on the VOR compass rose on the chart.

327°M

Figure 15-7 VOR Position Line

Example

The VOR needle of an RMI reads 147°. This is the magnetic bearing to the VOR, so to plot from
the VOR the reciprocal must be taken, which is 327°M. This can be plotted from the magnetic north
arrow of the VOR, as in figure 17-20.

Note that plotting the whole line from the VOR can lead to cluttering of the map, so the best
technique is just to plot one or two inches of the most relevant part of the position line.

General Navigation (NPA25) - Edition 4 – 110630 15-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5.2 Plotting from an NDB

The ADF needle of an RMI shows the magnetic bearing from the aircraft to an NDB, measured
against the aircraft’s magnetic meridian. When the variation at the aircraft’s position is taken into
account the bearing is a true bearing, related to the aircraft’s meridian. The reciprocal gives the true
bearing from the beacon to the aircraft, but still measured from the aircraft’s meridian.

There are two choices. The first is to apply convergence between the aircraft and the NDB to plot
from the NDB’s meridian. This would give the actual bearing to be plotted. In practice however it is
easier to draw a line parallel to the meridian at the aircraft’s position through the NDB, and plot
from that line.

Note that because the variation and meridian at the aircraft must be used, the magnetic north tick on
the NDB symbol can only be used if there is little East/West difference between the NDB and the
aircraft.

A useful mnemonic for the factors to be taken into account is VCR, standing for Variation,
Convergence and Reciprocal.

Transferred meridian parallel


to the meridian through the
aircraft position.

Chart convergence 5o

102°T 097oT to plot from


beacon local meridian

Figure 15-8 NDB Position Line

Example

The ADF needle of an aircraft in the northern hemisphere is reading 290°. Variation at the NDB is
9°W, at the aircraft is 8°W.
The bearing means that the direction from the aircraft to the NDB, measured at the aircraft, is
290°M. To plot this bearing it must be corrected for variation, convergence and reciprocal.
Variation must be taken into account at the aircraft, i.e. 8°W. This gives a bearing of 282°T. To
remove the effect of convergence the aircraft’s meridian must be used as a North reference, so a line
15-8 General Navigation (NPA25) - Edition 4 – 110630
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
should be drawn through the position of the NDB, parallel to the nearest meridian to the aircraft.
The reciprocal, which is 102°T, must be plotted as shown in figure 17-3.

5.3 Plotting from a DME

The position lines from a DME are the most straight-forward to draw, but there is one complication
in finding a position only from two DME fixes.

Ordinary compasses are used to plot position lines. The arms of the compasses are opened so that
the distance from the point to the pencil tip is the same as the DME range reading at the chart scale.
The point is then placed on the location of the DME, and an arc is drawn with the pencil tip. Of
course this would constitute an entire circle, but minimise clutter on the chart by only plotting the
most relevant part of the circle.

Example

The DME reads 47.0.

Open a pair of compasses to 47 nm using either the chart scale or the latitude markings on the
meridians, of which each minute represents 1 nm or a degree 60 nm. Caution: you cannot use the
longitude markings on the parallels, as they are less than 1 nm to the minute. The compasses must
be opened north to south up the meridians if you are using this technique.

Place the point on the DME and mark an arc.

47 nm

0 nm 10 nm 20 nm 30 nm 40 nm 50 nm

Figure 15-9DME Position Line

6 Fix from Position Lines

Of course in order to fix the position of the aircraft a minimum of two position lines is required.
These can be of any of the above types, as long as the geometry is correct, so the angle between
position lines is sufficient. 90° between two position lines, or 60° between each of three is ideal,
whereas if the angle is only small the position will not be well-defined.

General Navigation (NPA25) - Edition 4 – 110630 15-9


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
6.1 Two-position-line fix

Two position lines can come from any two navigation aids, of the same type or different. The
important factor is that the geometry allows for a good fix.

A two position line fix from a VOR or NDB and a co-located DME is a common fix with the ideal
geometry. The range position line from the DME will always cross the radial position line from the
VOR or NDB at a right angle.

DME range

VOR Fix Position


bearing

Figure 15-10 VOR / DME Fix

6.2 DME / DME fix

A DME/DME fix is the most accurate available in ground-based radio navigation. However there is
a problem with using such a fix with two position lines. They will cross at 2 points, giving an
ambiguity. More information is required to decide which of the two points is the aircraft position.

Without further
information each fix
position is equally valid

Figure 15-11 Ambiguous DME / DME Fix

15-10 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
In the air this ambiguity would usually not arise, as the known approximate position of the aircraft
would rule out one of the two locations. However there are other methods of deciding which
position is correct, including knowing whether each DME range is increasing or decreasing with the
aircraft on a known track. For example in figure 17-6 if the aircraft is tracking south-easterly and
both DME ranges are increasing, the aircraft cannot be at the more northerly fix location.

6.3 Three-position-line fix

If three position lines are used instead of two then a more accurate fix can be obtained. However it
is unlikely that they will all intersect at a single point. This will result in a small, triangular region
between the three position lines known as a “cocked hat”. The fix should be plotted in the middle of
the cocked hat, as shown below.

Figure 15-12 A Cocked Hat

General Navigation (NPA25) - Edition 4 – 110630 15-11


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

15-12 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 16 The Polar Stereographic Projection.

1 The Projection.
The Polar Stereographic is a perspective chart, produced by shining a light source through the
reduced earth from the opposite pole as shown in the diagram below. It is the only perspective
chart in the JAA syllabus.

Figure 16-1 The Polar Stereographic Projection.

2 Properties

2.1 Graticule.

Figure 16-2 The Polar Stereographic Graticule.

The graticule consists of meridians as straight lines radiating out from the pole, cutting all parallels
of latitude at right angles, and parallels of latitude as concentric circles centred on the pole. Note
how the parallels appear further apart towards the edges of the chart because of the scale expansion.

General Navigation (NPA25) - Edition 4 – 110630 16-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2.2 Scale.

The scale expands along the meridians away from the point of tangency according to the formula:

Scale at any Lat = Scale at pole x sec2 colat or denom A = cos2 ½ colat A
2 denom B cos2 ½ colat B
The colat = (90°- lat). The scale deviation increases from zero at the pole (where the chart and the
reduced earth touch) to 1% at a latitude of 78½°, so this chart is normally accepted as being of
constant scale at latitudes greater than 75°.

2.3 Great Circles.

It can easily be seen from the above graticule diagram that meridians are straight lines. Any great
circle passing through the pole is therefore a straight line. Note, however, that it is possible to show
the equator (another great circle) on this chart, and that it is a circle centred on the pole. Great
circles which do not pass through the pole are therefore concave towards the pole.

If the chart is used within its normal maximum coverage of latitudes greater than 75°, the difference
between a great circle and a straight line is so small that it can be ignored for plotting purposes.

2.4 Rhumb Lines.

Rhumb Lines are concave to the point of tangency. Note how the parallels of latitude (all rhumb
lines) appear on the chart.

2.5 Chart Convergence

Chart convergence on a polar stereographic chart is simply ch. long. This is clear from the fact that
the meridians are straight, so convergence is independent of latitude, and that convergence is correct
at the pole. At the pole sin Lat is sin 90°, which is 1, so convergence, ch. long. x sin lat = ch. long.
x 1.

2.6 Summary of Properties:

CONFORMALITY Yes.
SCALE Correct at the pole, expands away from it.
EQUIVALENCE No.
Curves concave to pole of projection, but assumed to be straight lines
GREAT CIRCLES
near the pole.
RHUMB LINES Curves concave to the projected pole.

3 Uses.
The Polar Stereographic chart is used as a plotting chart in high latitudes, and for topographical
maps.

4 Calculations
The geometry of a polar stereographic chart allows for certain calculations which are not possible
on other charts.

16-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4.1 Track to Fly

If the origin and destination of the flight are at the same latitude then it is possible to calculate the
track to fly at any point along the route for an approximate great circle. This relies on the fact that a
straight line approximates to a great circle.

A triangle can be draw with the start and end points of the route and the pole as its corner. This will
be an isosceles triangle, as the distances from the pole to the start and end points are the same. Once
the angle between the start and finish meridians (two sides of the triangle are on these meridians)
has been found, the other two internal angles of the triangle can be calculated, as they are equal and
the three angles must sum to 180°. From this the initial track and final track can be calculated.
From the chart convergence track at any given longitude can be found.

Example

Find the initial and final tracks on a great-circle route from 83°N 037°E to 83°N 121°E.

Initially draw a diagram of the problem, a simplified graticule with a sketch of the meridians on
which the route starts and ends. See figure 18-3.

48°
121°E

48°
NP
84°

48°

037°E
Greenwich Meridian

Figure 16-3 Example calculation

The most common mistake is in the local true north. For every meridian true north is toward the
centre of the graticule if it is a north-polar stereographic chart, as marked on the Greenwich

General Navigation (NPA25) - Edition 4 – 110630 16-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Meridian on figure 17 – 3, and toward the outside of the chart for a south-polar chart. In this case
the angle between the two meridians is 84°. The three angles in the triangle must add up to 180°, so
the sum of the other two angles must be

180° - 84° = 96°

Since they are equal each must be 48°.

The centre of the graticule is the North Pole, so true north is toward the centre. This means that the
initial track, measured round from the 037°E meridian’s northerly direction, is 048°. The final track
again measured round from north on the 121°E meridian is 48° less than true south, so the track is

180° - 48° = 132°

The track can be found at any longitude by finding the ch. long. from the start location then either
adding this to or subtracting it from the initial track. A diagram can be used to determine whether it
is to be added or subtracted. Alternatively, in this case as the final track is higher than the initial
track the track increases all along, so simply add ch. long. to the initial track.

4.2 Great Circle Distance

If the start and end points are at the same latitude we can also estimate the great circle distance
(N.B. this is an approximation that only holds at high latitudes where polar stereographic charts are
used). The triangle formed by the track and the start and end meridians can be split into two equal
right-angle triangles, divided along the central meridian of the track.

The latitude at which the flight is to depart from and arrive at is known (and both must be the same)
so the distance of this from the pole can be calculated from the formula

60 x (90°-Lat)

which is a specific case of the ch. lat. to distance formula. This distance is the hypotenuse of the two
right-angled triangles, and the angle at the pole is half the ch. long. for the route. From this the
distance along track to the half-way point is simply

hypotenuse x sin (½ x ch. long.)

So the complete formula for the total distance along the route is

2 x [ 60 x ( 90° - Start/end Latitude ) ] x sin ( ½ x ch. long.)

16-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example

Using the same flight as the last example, routing from 83°N 037°E to 83°N 121°E. The distance
from the pole to the start point is a 7° arc of a great circle, which at 60 nm per degree is 420 nm.
The angle between the two meridians is 84°, so to the half-way meridian is 42°. See figure 17 – 4.
The side of the triangle half the length of the route is opposite this angle, so use the formula

Opposite = Sine x Hypotenuse


= sin(42°) x 420 nm

That distance is 281 nm. The distance for the whole route is twice this figure, 562 nm.

83°N

121°E

NP

42°

281 nm
420 nm

037°E

Figure 16-4 Example of Great Circle Distance

General Navigation (NPA25) - Edition 4 – 110630 16-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

16-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 17 Grid Navigation.

1 The Navigation Problem.


Navigation in high latitudes requires different and somewhat more complicated techniques than the
procedures used below about 70°.

• Firstly, the very considerable meridian convergence makes it very difficult to steer accurate true
headings or measure accurate tracks anywhere near the poles. Plotting using the procedures we
have seen up to this point is virtually impossible.
• Secondly, the magnetic directive force is insufficient to give reliable magnetic compass
readings, so it is often necessary to fly using a directional gyro as the heading datum.

2 The Solution.
There are two linked solutions to the problem. A reasonably sophisticated remote indicating
compass may be operated in the ‘unslaved’ (DG) mode, in which magnetic monitoring of the
compass readout is cut off. It is also possible in many systems to switch off compensation for
astronomical drift and, if provided, automatic compensation for transport drift. Many gyro-based
compass systems have drift setting control which can then be used to make the gyro drift at a fixed
rate appropriate to the route.

A suitable chart, such as the polar stereographic, is then given a square grid overlay in addition to
the lat and long graticule. The gyro is then ‘slewed’ (turned) to a suitable reading aligned with the
grid datum. The pilot is then able to fly ‘gyro’ headings instead of true or magnetic headings.

Figure 17-1 Polar Stereographic Chart with Grid Overlay.

The above diagram illustrates a typical polar stereographic chart with a square grid overlay, in
which the Grid North is aligned with the Greenwich meridian.

General Navigation (NPA25) - Edition 4 – 110630 17-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2.1 Grid Convergence.

Note that on the above chart illustration of a North Polar Chart, True North direction is always
towards the centre of the chart, whereas Grid North is at the top of the chart. (On South Polar
charts, True North is all directions away from the Pole.) The angle between grid north and true
north therefore varies over the whole chart. Grid Convergence is defined as the angular difference
between True North and Grid North measure East or West of Grid North. At position ‘A’, for
example, the grid convergence is 45°E.

2.2 Grivation and Isogrivs.

Figure 17-2 Grivation

It is possible to use grid navigation techniques at lower latitudes than the 75° appropriate for Polar
Stereographic charts, and it is not uncommon to find a grid overlay on some Lambert’s or Oblique
Mercator projections. In these cases it may be possible to use a magnetic compass as a heading
reference, and instead of carrying out the more normal conversion between True and Magnetic
references it becomes necessary to convert from Grid to Magnetic and vice versa.

17-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Grivation is defined as the angular difference between the Grid North and Magnetic North
measured East or West of Grid North.

Lines of equal Grivation are known as Isogrivs

Grivation is easily found by algebraically adding together the variation and grid convergence, as
shown in Figure 18-2. In summing variation and grid convergence East and West can be thought of
as opposites: easterly variation or convergence can be thought of as negative westerly variation or
convergence. Conversely westerly variation or grid convergence can be treated as negative easterly
variation or grid convergence. This allows the addition of variation and grid convergence where one
is east and the other west.

3 Grid Calculations on a Polar Stereographic Chart


On a polar stereographic chart it is possible to calculate grid convergence at any point knowing the
meridian with which the grid is aligned. Convergence consists of two parts, a magnitude (angle) and
a direction, East or West.

The magnitude of convergence is simply the difference in longitude, (d. long or ch. long) between
the meridian with which the grid is aligned and the location where the convergence is to be
calculated. Remember that true north and grid north are the same on the meridian with which the
grid is aligned.

The most reliable way of finding the direction of convergence is a simple diagram. Sketch in the
approximate grid direction and the meridian on which the grid convergence is required. Make sure
you mark the correct True North on the meridian (towards the centre of the graticule on a North-
Polar chart, outward on a South-Polar chart) and a grid line crossing the meridian in question. For
example see figure 18 – 3, for finding the convergence at 065°E using a grid aligned to 030°W.
Looking along the Grid North, if true North is to the left (west) then convergence is West. If true

NP

065°E

030°W Greenwich
Meridian
North is to the right (East) then convergence is East. In figure 18-3 convergence is West.

Figure 17-3 Example Sketch of Grid Convergence Diagram

General Navigation (NPA25) - Edition 4 – 110630 17-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

17-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section Four Space and Time

Chapter 18 The Solar System


The solar system consists of the Sun and all the bodies in its immediate gravitational influence, i.e.
in orbit around the Sun.

Within the solar system are 9 known planets and a total of 63 natural satellites to these planets. In
order of mean distance out from the Sun the planets are Mercury, Venus Earth, Mars, Jupiter,
Saturn, Uranus, Neptune and Pluto. There are estimated to be around 2000 smaller bodies –
asteroids and comets – in addition to the many asteroids in the ‘asteroid belt’ beyond Mars.

1 Orbits and Kepler’s Laws

1.1 An Ellipse

All the planets orbit the sun anticlockwise as viewed from above the Sun’s North pole in an elliptic
orbit. An ellipse is a closed, smooth curve similar to a circle but elongated in one direction. See fig
20 – 1. It has two foci (singular ‘focus’) and one long (major) and one shorter (minor) axis. Half the
length of one axis is sometimes described as a radius, so it has two different radii. A circle is a
specific type of ellipse, with the foci at the same point and both axes the same length.

Minor axis

Major axis

Foci

Figure 18-1 An ellipse

1.2 Planetary Paths – Kepler’s first law

The planets and minor bodies of the sun orbit in elliptical paths with the sun at one focus of the
ellipse (the other focus is simply a point in space, there is nothing there). This is Kepler’s first law.
Most of the planets’ orbits are only slightly elliptical, they are almost circular. The exceptions are
Mercury and Pluto, Pluto having an orbit so elliptical (described as ‘eccentric’) that it spends much
of the time inside the orbit of Neptune.

General Navigation (NPA25) - Edition 4 – 110630 18-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
The orbits of most of the planets lie more or less within the same plane, approximately 7º to the
Sun’s equator. See fig 20 - 2. Pluto is again the exception with an inclination to the Earth’s orbital
plane of around 17º. For this reason and because of the eccentricity of its orbit some astronomers
prefer not to define Pluto as a planet.

Figure 18-2 Relative Orbits Around the Sun

1.3 Changes in Speed – Kepler’s second law

At certain points in an elliptical orbit a planet is further from the Sun than at other times. This
means that some energy must be used to take the planet further from the sun, just as energy is used
to lift a weight further from the Earth. This is an increase in potential energy. In the case of the
planets the only source of energy is their kinetic energy, so as a planet moves away from the sun its
potential energy increases, reducing its kinetic energy and thus its speed. So at the further points in
the orbit the planet is moving more slowly. In fact the speed is such that in a given length of time a
line connecting the centre of the planet and the centre of the sun sweeps out the same area
regardless of where the planet is in its orbit. See fig 20 - 3. This is Kepler’s second law.

Area A = Area B

Time x A B Time x

Figure 18-3 Kepler’s Second Law

18-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.4 The Planetary Year – Kepler’s third law

Kepler’s third and final law relates to the length of a planetary year, the period of time that a planet
takes to make one complete orbit of the Sun. The planets further from the sun take longer to do so,
and those closer take less time. Our planet, the Earth, completes one orbit of the Sun every 365.25
days.
Kepler’s third law relates the distance from the Sun to the time period of orbit by stating that:
d3
t2 is a constant.

1.5 Summary of Kepler’s Laws

• Kepler’s first law: the planets are in elliptical orbits with the sun at one focus of the
ellipse.
• Kepler’s second law: the area described in a planetary orbit by a straight line joining
the centre of the planet and the centre of the sun is equal for
equal time spans.
• Kepler’s third law: the ratio of the distance of a planet from the sun cubed to its
orbital period squared is constant, i.e. d3/t2 is constant.

2 Rotation of Planets
At the same time as orbiting the Sun all the planets also spin on their own axes. Seven of the
planets, including the earth, orbit anticlockwise as viewed from above their north poles, with each
axis at a large angle to the plane of orbit. Venus has an axis at large angle to its plane of orbit, but
rotates clockwise as viewed from above its north pole. Uranus has an axis at a small angle to its
plane of orbit.

3 The Earth’s Motion

3.1 Perihelion and Aphelion

The Earth is just another planet, so has a standard, elliptical orbit. The point in the Earth’s orbit at
which it is closest to the Sun is known as the Perihelion, and the point where it is furthest from the
Sun is called the Aphelion. The Perihelion occurs on approximately the 3rd January each year, the
Aphelion on approximately the 3rd July. A tip for remembering these is “Aphelion” > Afar (further
away).

Using Kepler’s second law it can be shown that at the Perihelion the Earth’s speed of movement is
greatest, and at Aphelion it is least.

General Navigation (NPA25) - Edition 4 – 110630 18-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 18-4 Closest Approach and Furthest Point

3.2 Obliquity of the Ecliptic

There is an angle of 23½° between the equator and the plane of its orbit (the plane sometimes called
the Plane of the Ecliptic – defined later in these notes), which is known as the obliquity of the
ecliptic. This means the Earth’s axis is at 66½° to the plane of orbit and spins about this axis
anticlockwise from above the North Pole. The tilt of the axis away from the normal to the plane of
orbit, the axis pointing in the same direction in space all year, causes changes to the amount of
sunlight received and the concentration of solar energy at any given point through the year. This
causes our seasonal climates.

3.3 Solstices, Equinoxes and Tropics

The North Pole points most towards the sun on 21st June, which is known in the Northern
Hemisphere as the Summer Solstice, and the sun is then overhead 23° 30’ N. This latitude is called
the Tropic of Cancer. The date marks the middle of the Northern-Hemisphere summer and the
Southern-Hemisphere winter. On the 21st December the North Pole points most away from the Sun,
and this is known in the Northern Hemisphere as Winter Solstice. The sun is overhead 23°30’ S,
known as the Tropic of Capricorn.

The Sun passes overhead the equator on about March 21st, known as the Vernal or Spring Equinox.
The Sun passes overhead the equator again on about 23rd September, known as the Autumnal
Equinox (of course in the Southern Hemisphere these will also be the other way around). As will be
seen later the day and night are the same length for an observer not on the equator only on the
equinoxes. This means that on the spring and autumnal equinox, the sun will rise and set at
approximately the same time regardless of the latitude of the observer.

18-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Figure 18-5 Soltices and Equinoxes

In a diagram showing the solstices it may look like they occur at the Perihelion and Aphelion. This
is not the case, but by coincidence the Perihelion occurs within 2 weeks after the Winter Solstice,
and Aphelion within 2 weeks after Summer Solstice (in the Northern Hemisphere; the solstices are
of course reversed in the Southern Hemisphere!).

General Navigation (NPA25) - Edition 4 – 110630 18-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4 The Celestial Sphere

The Celestial Sphere is a theoretical device for more easily describing the position of heavenly
bodies in relation to the Earth and in relation to an observer. The Celestial Sphere is a spherical
surface with an infinite radius concentric with the Earth (centred on the centre of the Earth). Any
celestial body, either within or outside the solar system can then be projected onto the celestial
sphere, as can the observer, and relative positions can then be sensibly described.

The mapping of a body onto the Celestial Sphere is similar to chart projection. It is treated as if a
light in the centre of the Earth was being used to project each body onto an infinitely distant
spherical surface. See fig 20 – 6. This means that its position on the Celestial Sphere becomes two-
dimensional, being described by only two angles. These angles are the declination and the hour
angle, and are analogous to latitude and longitude on the Earth’s surface.

The Celestial
Sphere

The Earth

Figure 18-6 Projecting celestial bodies and an observer onto the Celestial Sphere

The Celestial Sphere does not spin with the Earth, so of course an observer tracks across its surface
as the Earth spins.

The Celestial Sphere has North and South Poles as does the Earth, directly overhead the Earth’s
poles, and also a Celestial Equator directly overhead the Earth’s Equator. This Celestial Equator is
often known as the Equinoctial.

18-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4.1 Declination

Declination is analogous to latitude. Circles of declination are small circles on the celestial sphere
parallel to the equinoctial (or celestial equator), declination being defined as the angular distance to
a body on the celestial sphere measured north or south through 90º from the equinoctial along the
hour circle of the body. See Fig 20 – 7. Declination of a body is the same as the latitude of a point
on the Earth if the body is directly overhead that point.

4.2 Hour Angle

Hour Angle is the celestial co-ordinate corresponding to longitude, and is defined as the arc of the
equinoctial intercepted between a datum hour circle and the celestial meridian of the body,
measured westward from 0º to 360º. In this of course it differs from longitude, which is measured
180º east or west. See fig 20 - 7. A common hour angle datum is that of Greenwich, so the position
of the body in the E-W direction is defined by the Greenwich Hour Angle (GHA).

NCP

Circle of Declination Specified datum


celestial meridian

Declination

Hour Equinoctial – the


angle Celestial Equator

Hour circle – the celestial


meridian of body
SCP

Figure 18-7 Declination and Hour Angle

4.3 Hour Circles and Celestial Meridians

A celestial meridian is analogous (equivalent) to a meridian on Earth. An hour circle (actually a


half-circle) is similar to a celestial meridian, but is stationary with respect to the Earth, whereas the
celestial meridian is stationary on the celestial sphere.

General Navigation (NPA25) - Edition 4 – 110630 18-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
5 Relative Co-ordinates
So far we have looked at the co-ordinates of a body on the Celestial Sphere with reference to a
datum fixed with respect to the Earth, the Equinoctial or the Greenwich Hour Circle. Altitude and
azimuth are co-ordinates relative to the observer.

5.1 Altitude

Altitude is defined as the smaller arc of the vertical circle through the body, from the celestial
horizon to the body. See Fig 19-8. This is, then, the angle measured from the horizon to the star
using a sextant.
In terms of the Celestial Sphere, if the observer and the body are both projected onto the Celestial
Sphere altitude is the angle between the observer’s celestial horizon and the body measured along a
great circle that passes through the body and the observer, which will be the vertical circle. The
celestial horizon is a great circle 90° round the Celestial Sphere from the observer, and for
measuring at any distance that is very large relative to the radius of the Earth it is effectively the
same as the observer’s sensible Earth horizon (the sensible horizon is corrected for refraction).

5.2 Azimuth

Azimuth is the true bearing of the star from the observer. On the Celestial Sphere it is the angle
between the hour circle of the observer and the great circle passing through the observer and the
body, measured round from north. See fig 20 – 8

NCP
Azimuth
Altitude

Observer

Observer’s hour
circle

Observer’s celestial
horizon

SCP

Figure 18-8 Altitude and Azimuth on the Celestial Sphere

5.3 Zenith

The point on the celestial sphere vertically above an observer on the earth is called zenith.

18-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
6 The Ecliptic

The Ecliptic is defined as the path of the Sun on the Celestial Sphere. The Plane of the Ecliptic is
therefore the plane in which the Earth passes around the Sun. The latter is a simpler way of thinking
about this concept, as of course the path of the sun moves north and south within the Celestial
Sphere throughout the year.

Summer Solstice (Northern


Hemisphere) Winter Solstice
NP
NP
66½°

SP SP
The Plane of the Ecliptic

Figure 18-9 The Ecliptic

6.1 Seasonal Variation

As can be seen from fig 20 - 9 the plane of the ecliptic is not in the same position on Earth all year.
In Northern Hemisphere summer the sun is north of the Equator. In Northern Hemisphere winter the
sun is south of the equator. The Plane of the Ecliptic moves in the same way across the Celestial
sphere. The two points in the year when the Plane of the Ecliptic and is in the same place on the
Celestial Sphere as the Equinoctial (the Celestial Equator) are called the equinoxes, the Vernal (or
spring) Equinox and the Autumn Equinox.

General Navigation (NPA25) - Edition 4 – 110630 18-9


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

18-10 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 19 Time
Time of day is based on the rotation of the Earth and resulting passage of the Sun from east to west.
This is historically the most important and obvious recurrent event. Of course all the heavenly
bodies appear to pass from east to west across the sky as the Earth revolves west to east.

1 Definitions of Day

The simplest way of noting the passage of a heavenly body is by the point at which it is closest to
being overhead the observer. This will be when it passes overhead the meridian on which the
observer is located. In terms of the celestial sphere the body and the observer will be at the same
hour angle, and the passing of a body through a defined meridian like this is called a transit.

The basic definition of a day is the time between two transits of a heavenly body.

1.1 Sidereal Day

A sidereal day is the time between two transits of a fixed point in space, a point distant enough to be
considered at infinity. This means that the only factor determining the length of a day is the Earth’s
rotation. The point most commonly chosen is the first point of Aries. See fig 20 – 1.

A sidereal day is about 23 hours and 56 minutes. The reason that the familiar day is longer will be
made clear later.

Aries

Figure 19-1 The Sidereal Day

The sidereal day is of constant duration, because the movement of the Earth with respect to the first
point of Aries is negligible and the rotation of the Earth is at a constant rate.

General Navigation (NPA25) - Edition 4 – 110630 19-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1.2 The Apparent Solar Day

An apparent solar day is the time between two transits of the sun over a given meridian. Due to the
movement of the Earth around the Sun, at approximately one degree per day on average (360° in
365.24 days) the Earth must rotate through more than 360° between two transits of the sun. See fig
21 – 2.

Figure 19-2 The Apparent Solar Day

Since the Earth moves 360° around the sun in approximately 365 days the angle through which it
moves in one day is on average 360°/365 = 0.986°, about 1°. As can be seen in figure 20 – 2 this
angle is the same as the angle through which the Earth has to move in addition to the 360° of the
sidereal day (marked in fig 21 – 2). As the Earth turns at approximately 15° per hour the apparent
solar day is about 4 minutes longer than the sidereal day.

As has been noted in the chapter on the solar system the Earth does not pass around the Sun at the
same distance or at the same speed throughout the year. Close to the perihelion, when the Earth is
close to the Sun the angle moved through is greater for the same distance. In addition by Kepler’s
2nd law the Earth is moving more quickly, and so the distance is greater. This means that the Earth
has to spin further between transits of the sun, and the apparent solar day is longer. Close to
aphelion the reverse happens, giving slightly shorter days.

1.3 The Mean Solar Day

It would be rather inconvenient if the length of a day varied slightly depending on the time of year.
The average length of a solar day throughout the year has been calculated and defined as being the
mean solar day, consisting of 24 mean hours.

From this is defined the concept of a mean sun, a theoretical construction of the position of the sun
that agrees with mean time. To agree with mean time, the time between two successive transits of
the mean sun is constant, at one mean day. The Earth is assumed to be in circular orbit of the mean
sun, a circular orbit of the same period as the Earth’s orbit about the real (apparent) sun. Since the
mean sun is a mean over a year it is always over the equator. Therefore in the celestial sphere the
mean sun is always in the equinoctial plane.

Mean time is used for all practical navigation except astronavigation, for which the difference
between mean sun and apparent sun is tabulated in the Air Almanac, in a table ‘Meridian Passage
19-2 General Navigation (NPA25) - Edition 4 – 110630
©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
and Declination of the Sun at 12h UT’ . This difference is known as the ‘Equation of time’ . This is
not an equation in the most common mathematical sense, rather a difference between apparent and
mean time. The greatest difference between mean time and apparent time is in November, when
apparent time is about 16 minutes ahead of mean time. In February mean time is around 14 minutes
ahead of apparent time.

General Navigation (NPA25) - Edition 4 – 110630 19-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2 Local Mean Time
Mean time can also be related to longitude, and measured as an arc, because the mean sun is
assumed to travel round the Earth at 360º each mean day, or 15º an hour. Since time is defined in
relation to the passage of the mean sun then the time in different locations can be related to the
difference in longitude between the locations. The time so defined is known as Local Mean Time
(LMT) because it is time specific to a location (or rather to a meridian) and defined with relation to
the mean sun.

180 +/- 12 hrs


135 W
- 9hrs
Meridians 135 E + 9 hrs
of
longitude North
Pole
90 W 90 E + 6 hrs
- 6 hrs Greenwich
45
Arc
of 45 E + 3 hrs
longitude 45 W
- 4 hrs
-3 hrs 0
Prime meridian

Figure 19-3 Longitude to Time Conversion

Local mean time is defined as being midday when the mean sun is overhead the meridian of the
location, i.e. when the mean sun is in transit with that meridian. When the sun is in transit with the
anti meridian then the LMT is midnight, or 0000 hours or 2400 hours.

2.1 Arc to Time Conversion

The difference between local mean times at different locations is calculated by dividing the
difference in longitude, or arc of longitude by 15, as the mean sun moves at 15º per hour. See figure
20 – 3. Usually the calculations are made with respect to local time at the Greenwich Meridian, or
Greenwich Meant Time (GMT). If GMT is calculated from the known LMT, then the unknown
LMT calculated from GMT the complications involved with crossing the International Date Line
(discussed later) are avoided.

The time is always later to the east, so if converting LMT to a more easterly LMT then add the
difference, if converting to a more westerly location subtract the time difference. If the result is less

19-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
than 00 00 hours, then take one day from the date and add 24 hours. If the result is greater than 24
00 hours then add one day to the date and subtract 24 hours.

The table in figure 21 – 4 can be helpful in arc to time conversions.

Arc Time
15º 1 hour
1º 4 minutes
15’ (arc minutes) 1 minute
1’ 4 seconds

Figure 19-4 Arc/Time equivalence

Example 1
At a location 024º15’ west the local mean time is 1415. What is Greenwich Mean Time?
Solution 1

Dividing the arc by 15º:


24º15’ = 1.617 hours or 1 hour 37 minutes
15º
Alternatively using the table above the first 15º represents a 1-hour change, the remaining 9º
represent 4 minutes each for a total 36 minutes and the remainder of 15’ represents 1 minute. Sum
these times:
1 hour + 36 min + 1 min = 1 hour 37 minutes

The location is west of Greenwich, so we are changing time to the east and therefore add this figure
to LMT.
1415 + 1h37 = 1552
Example 2
The local mean time at 065º30’ east is 1923. What is local mean time at 018º15’ west?
Solution 2
First calculate Greenwich Mean Time. Time difference :
65º30’ = 4.367 hours or 4 hours 22 minutes
15º
The location of known time is east of Greenwich, so we are calculating to the west, and this must be
subtracted.
1923 – 4h22 = 1501
The calculation can then be repeated for calculating the time at 018°15’ .
18°15’ = 1.217 hours or 1 hour 13 minutes
15°
Still westward so
1501 – 1h13 = 1348
Example 3
Time is 0241 LMT at 004° 00 W on 12 July. What is the time and date at 148°45’ W?
Solution 3

General Navigation (NPA25) - Edition 4 – 110630 19-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
To find GMT
004°00’ = 0.267 hours or 16 minutes
15°
East to the Greenwich Meridian, so add
0241 + 0h16 = 0257 GMT
To find the time 148°45’ W
148°45’ = 9.917 hours or 9 hours 55 minutes
15°
West of Greenwich so subtract this from GMT
0257 – 9h55 is less than 0000, so add 24 hours
0257 + 24h00 – 9h55 = 1702
Because the sum came to less than 0000 then we must subtract a day from the date, so the answer is
1702 on 11 July.
Further Examples
1. If the time is 1842 LMT at 156° W what is GMT?
2. If the time is 1203 LMT at 014°30’ E what is the time at 077°15’ W?
3. If the time is 1814 LMT on 18 February at 153°00’ E what is the time and date at
101°30’ W?

19-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3 Standard Time
Local mean time is impractical for most purposes, as over a relatively short distance the time could
vary by a few minutes. A system of keeping time has been devised based on UTC (Universal Time
Co-ordinated). UTC is derived from atomic time whereas GMT is based on Earth rotation, but for
all practical purposes they are considered to be the same. The local government decides a time basis
to use, the LMT of a given standard meridian. In most cases this differs from UTC by a whole
number of hours, although sometimes with an odd half-hour in the difference. The time so defined
is known as Standard Time, and is what would be set on local clocks and watches.

Standard time is based on a series of time zones. Each time zone is centred on a meridian, at 15°
intervals. Therefore the time zone extends 7½° either side of this meridian, and the standard time
for that time zone is taken to be local mean time at the central meridian. However this is only
strictly true in international waters, because as stated the local government decides on standard
time, and although this is always in practice the same as a nearby time zone it may not be a time
zone in which any part of the country is located.

Figure 19-5 World Time Zones

General Navigation (NPA25) - Edition 4 – 110630 19-7


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
These time zones can be denoted in two ways. The most common is by a number. This number is
the hours to add to local standard time to calculate UTC. So –2 is 2 hours ahead of UTC, and +6 is
6 hours behind. –4:30 is 4½ hours ahead of GMT. In figure 20 – 4 the numbers have been used in
the opposite sense. The other notation is a letter (see figure 20 – 5) but the order of the letters is
complicated so it is not necessary to remember what each means. It is worth noting however that Z
is the letter to denote Greenwich Mean Time, thus sometimes known as ‘zulu time’ as well as UTC,
and for local flying notice that A denotes –1, which is British Summer Time.

STANDARD TIMES (Corrected to July 1998)


LIST II – PLACES NORMALLY KEEPING UTC
Ascension Island Ghana Irish Republic*✝ Morocco Sierra Leone
Burkino-Faso Great Britain✝ Ivory Coast Portugal*✝ Togo Republic
Canary Islands* Guinea-Bissau Liberia Principe Tristan de Cunha
Channel Islands✝ Guinea Republic Madeira* St. Helena
Faroes*, The Iceland Mali São Tomé
Gambia, The Ireland, Northern✝ Mauritania Senegal
* Summer time may be kept in these places.
✝ The European Union directive states that Summer Time, one hour in advance of UTC, is kept from
1999 March 28d 01h to October 31d 01h UTC.

LIST III – PLACES SLOW ON UTC (WEST OF GREENWICH)


The times given
Below should be } subtracted from UTC to give Standard Time
added to Standard Time to give UTC.

h m h m
Argentina ................................................... 03 Canada(continued)
Austral (Tubuai) Islands1........................... 10 Quebec, east of long. W. 63°3 ................................ 04
Azores* ..................................................... 01 west of long. W. 63°* ................................ 05
Saskatchewan3 ................................................................
06
Bahamas* .................................................. 05 Yukon* ................................................................
08
Barbados.................................................... 04 Cape Verde Islands ................................................................
01
Belize ........................................................ 06 Cayman Islands ................................................................
05
……… ………

Labrador3*.............................................. 04 Galápagos Islands ................................................................


06
Manitoba* .............................................. 06 Greenland4
New Brunswick* .................................... 04 General* ................................................................
03
Newfoundland* ...................................... 03 30 01
Scoresby Sound ................................................................
Northwest Territories3* Thule area ................................................................
04
east of long. W. 85°............................. 05 Grenada ................................................................
04
long. W. 85° to W. 102° ...................... 06 Guadeloupe................................................................
04
west of long W. 102°........................... 07 Guatemala................................................................
06
Nova Scotia ............................................ 04 Guyana, Republic of ................................................................
04
Ontario, east of long. W. 90°* .............. 05
west of long. W. 90°* ............. 06 Haiti ................................................................05
Prince Edward Island* ........................... 04 Honduras ................................................................
06
* Summer time may be kept in these places.
1
This is the legal standard time, but local mean time is generally used.
2
Most stations use UTC.
3
Some areas may keep another time zone.
4
Mesters Vig and Danmarkshavn keep UTC.

Figure 19-6 Standard Time Tables – List II and extracts from List III, with footnotes

19-8 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
For practical use of standard time the Air Almanac has three tables of Standard Times (ST). List I is
of places fast on UTC, most of which are East of the Greenwich Meridian. The adjustment noted on
the table is added to UTC to give LST. List II is of places that use UTC as standard time. List III
contains those places slow on UTC, where the adjustment given should be subtracted from UTC to
calculate LST. See figure 20 – 6.

For calculations involving local Standard Time the documentation must always be used to look up
the time difference. This is because ST is not predictable, being defined by the local government.

Since the ST data are tabulated relative to UTC each calculation will in effect be through the
Greenwich Meridian, so date changes are sorted out in the same way as with LMT calculations. It is
important to read any relevant footnotes. Many countries change the time during the summer
(known as “ daylight saving” ), an hour ahead of standard time, and this will be noted (you will only
be expected to take account of this if full details are given in the question or the table). In other
cases some area of a country may be an exceptional case, and this information will also be given by
a footnote.
Example 4
If the ST in Barbados is 1215 on April 20th what is the time and date UTC?
Solution 4

From the table in figure 21 – 5 add 4 hours :


1215 + 4 hours = 1615, UTC is 1615, 20th April

Example 5
Standard time in the Yukon Territory of Canada is 2105 on October 2nd. What is ST in the Azores?
Solution 5
Look up the Yukon, add 8 hours to find UTC.
2105 + 8 hrs = 2905
Over 2400 so add 1 to the date and take off 24 hrs
2905 – 24 hrs = 0505
rd
UTC is 0505 on 3 October, looking up Azores subtract 1 from UTC to find ST
0505 – 1 hr = 0405
So ST in the Azores is 0405 on 3rd October.

General Navigation (NPA25) - Edition 4 – 110630 19-9


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
4 The International Date Line

4.1 Greenwich Anti-Meridian

When transiting the Greenwich Anti-Meridian when using local mean time, the date changes. The
reason for this meridian being chosen is simply convention. The reason that the date needs to be
changed is that moving all round the world the local mean time changes by 1 hour per 15°, and
always in the same sense. This makes a total of 24 hours change, so returning to the same location
this would put a traveller’ s time out by 1 day. At some point in the circumnavigation the time must
be changed by 24 hours.

If the time is 1800 UTC on 20th June, then travelling 180° west the time is 12 hours earlier, 0600
LMT on the 20th. However travelling east by 180° the time is 12 hours later than UTC, or 0600
LMT on the next morning, the 21st. So stepping over the 180°E/W meridian eastward from 180°E to
180°W the date must change from the 21st to the 20th, and vice versa.

At all times LMT is the same each side of the Greenwich Anti Meridian, but the date is one day
earlier at 180°W than at 180°E. Travelling east across the 180° meridian (NB east goes from 180°E
to 180°W!) the date loses a day, travelling west the date gains a day.

4.2 International Date Line

For the purposes of standard time rather than mean time the date changes on the International
Dateline. This is a line fixed by international agreement, and to a large degree follows the
Greenwich Anti Meridian. However it deviates in several places to avoid separating islands in a
group.

In calculating time in a given location the dateline is already taken into account by making all
calculations by way of UTC. This is because then no calculation ever crosses the dateline.

5 The Year
The true year is approximately 365.24 days. The calendar year is 365 days. In order to correct for
the extra 0.24 days per annum years divisible exactly by 4 are leap years, and have an extra day
added in February. However this would of course compensate for 0.25 days per annum, so another
correction is required, of 3 days every 400 years, so a year that divides exactly by 100 is not a leap
year unless it also divides exactly by 400.

This system is known as the Gregorian calendar, and it will be correct to the nearest day for
thousands of years.

19-10 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 20 Sunrise and Sunset

1 Variation of Sunrise and Sunset


The time of sunrise and sunset varies with latitude and also varies in a given location depending on
the time of year. Both of these variations come about due to the tilt of the Earth’ s axis, and fig 21 –
1 shows a diagrammatic representation on the Celestial sphere of why sunrise and sunset vary as
they do.

The sun rises at the observer’ s location when the top of the sun rises above the horizon. For the
purposes of this diagram that is approximated to the observer’ s celestial horizon. The dashed arrow
shows the track of the Sun in the Celestial Sphere in June, the dotted line shows the track of the sun
in December. It can clearly be seen that the sun rises above the celestial horizon later in the day in
December than in June, and spends less time above the horizon. This is true for an observer in the
Northern hemisphere away from the equator. Of course the sun rises early in December and late in
June for an observer in the southern hemisphere.

North Pole

Celestial horizon

June

December

North Pole

Celestial horizon

June

December

Figure 20-1 Variation of Sunrise and Sunset Times

General Navigation (NPA25) - Edition 4 – 110630 20-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
If sunrise and sunset are calculated in local mean time rather than standard time then time of sunset
and sunrise is independent of longitude. All observers at the same latitude will experience sunrise
and sunset at almost exactly the same local mean time regardless of their longitude. This is because
the variation over a complete revolution of the Earth is simply the change in sunrise or sunset times
from one day to the next at a single point, which is a very small change.

The second part of fig 22 – 1 shows an observer who is close to the equator, within the tropical
region. As can be seen from this observer’ s celestial horizon the sun rises later and sets earlier in
summer than for an observer away from the tropics. There is also much less annual variation, and
indeed an observer on the equator would always see the sun rise at the same time throughout the
year, 0600 local mean time, and see the sun set at 1800 local mean time (to within 15 minutes).

Remember also that at the spring and autumnal equinox, the sun will rise and set at approximately
the same time regardless of the observer’ s latitude. It is also useful to note that the rate of change in
the length of the day (daylight hours) is greatest around the equinoxes.

20-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2 Sunset and Sunrise Tables
Sunrise and sunset times can be calculated from tables in the Air Almanac, which show the times at
specified latitudes on specified dates for an observer on the Greenwich Meridian. Tabulated sunset
and sunrise times indicate when the top of the sun rises above or falls below the observer’ s visible
horizon. This is not quite the same as the sensible horizon (a tangent to the Earth at the observer’ s
location) or the celestial horizon because of refraction. See fig 21 – 2.

Figure 20-2 Refraction Effects on Visible Horizon

As already mentioned the sun always rises at approximately 0600 local mean time and sets at
approximately 1800 local mean time near the equator. For other latitudes the tables in the Air
Almanac are as shown in figure 22 – 3, an extract from the Air Almanac. In these tables the open
box symbol (similar to ) indicates that on that day at that latitude the sun never sets. The filled
box (similar to ) indicates that the sun never rises on that day at that latitude.

Figure 20-3 Sunrise Table

As can be seen in figure 22 – 3 the sunrise (and sunset) times are tabulated every 3 days and at
certain latitudes. For the intermediate days, and for other latitudes, simple, linear interpolation is
possible to find with sufficient accuracy the sunset or sunrise time on a given day at a given
latitude. Note that sunrise occurs earlier and sunset later for any location with increased altitude.

General Navigation (NPA25) - Edition 4 – 110630 20-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example 1
What is the time of sunrise at 64° N on 16 July?

Solution 1

Look up 16 July on the table at 64°N. Sunrise is 0216.

Example 2

What is the time of sunrise at 57°N on 3 August?

Solution 2

Look up 3 August, at the latitudes either side of 57°N. Sunrise is 0356 at 58°N and 0406 at 56°. 57°
is half way in between, so half way between 0356 and 0406 is sunrise at 57°, which is 0401.

Example 3
What is the time of sunrise at 55°N on 3 July?
Solution 3
Take the sunrise on 1 July, midway between 54° and 56°: midway between 0332 and 0318 gives a
sunrise time of 0325. Sunrise on 4 July is midway between 0320 and 0334, which is 0327. So
sunrise is later by 2 minutes after three days, a change of about 40 seconds per day. Sunrise on 3
July would be about 40 seconds earlier than on 4 July. To the nearest minute sunrise on 3 July is
0326.
Further Examples

1. What time is sunrise on 12 August at 63°N?


2. What time is sunrise on 23 July at 53°N?
3. What time is sunrise on 29 June at 61°N?

20-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3 Civil Twilight

Of course when the sun sets the sky does not immediately darken. The time after the sun has set but
before darkness completely falls is known as twilight. There are three different definitions of
twilight. Although originally these were related to how dark the sky had become, they are now
defined in terms of how far the sun has fallen below the celestial horizon.
The only definition used in civil aviation is civil twilight. Evening civil twilight is defined to be the
period of time between sunset and the centre of the sun falling more than 6° below the celestial
horizon. Morning civil twilight is defined to be the period of time between the centre of the sun
ascending above 6° below the celestial horizon and sunrise.
The duration of twilight is not constant. It is shortest near the equator and increases in duration on
any one day with increasing latitude. The duration also changes with the seasons. The times of the
end of evening civil twilight and the beginning of morning civil twilight are tabulated in the Air
Almanac in much the same way as sunrise and sunset times. There is an additional symbol used, ////,
which denotes that although the sun sets it never falls more than 6° below the celestial horizon, so
civil twilight does not end until sunrise, and begins at sunset.

Figure 20-4 Civil Twilight Tables

The duration of civil twilight is calculated from two tables. To find the duration of morning civil
twilight first find the time of sunrise, then find the time of the beginning of morning civil twilight.
The time difference between the two is the duration of morning civil twilight. The duration of
evening civil twilight is the time difference between sunset and the end of evening civil twilight.

General Navigation (NPA25) - Edition 4 – 110630 20-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Example 4
What is the time of the end of evening civil twilight at 63°N on 6 August?
Solution 4

The end of evening civil twilight is 2205 at 64°N and 2137 at 62°N. 63°N is half way between, so
interpolate half way evening civil twilight ends at 2151.

Example 5
What is the duration of morning civil twilight at 58°N on 17 July?
Solution 5
Sunrise is 0321 on the 16 July and 0326 on the 19 July (use figure 21 – 3). Interpolating between
these dates, to the nearest minute sunrise is at 0323 on 17 July. Morning civil twilight begins at
0216 on 16 July and 0224 on 19 July. Interpolating between these dates the time changes by about 2
minutes 40 seconds each day. Adding 1 day’ s change to the time on the 16 July, the beginning of
morning civil twilight is 0219 to the nearest minute on 17 July.

The time period between 0219 and 0323 is 1 hour and 4 minutes, which is the duration of morning
civil twilight at 58°N.
Further examples
1. What time does evening civil twilight end at 67°N on 13 August?
2. What is the duration of morning civil twilight on 1 July at 57°N?
3. What is the duration of morning civil twilight on 13 August at 57°N?

20-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section Five Speed Time Distance Calculations

Chapter 21 Relative Velocity and Speed Adjustments

1 Relative Velocity
The relative velocity of two aircraft moving in the same or opposite directions can be critical to the
safe separation of those aircraft. Adjustment of speed may be required to keep to a minimum time
separation at a beacon or FIR boundary, or calculations may be required to allow an aircraft on an
opposite heading to pass through the same level keeping separation minima. Although the FMS of a
modern aircraft is capable of performing such a calculation on many older aircraft still in service the
pilot is required to do so.

1.1 Overhauling or following

Knowing the groundspeeds of the two aircraft their relative velocity is calculated by subtracting the
speed of the leading aircraft from the speed of the following aircraft. Of course if the leading
aircraft is faster than the following aircraft then the result will be negative, indicating a safe
situation where the following aircraft is not overhauling the leading aircraft. This might allow
reduced time separation minima at reference points such as beacons.
Example 1
A Boeing 767 flying at 385 knots groundspeed is followed by an A320 flying at 420 knots. What is
the velocity of the Airbus relative to the Boeing?
Solution 1 A-B: 420 kts - 385 kts = 35 kts
The Airbus is closing the Boeing at 35 kts.
Example 2
A DC-10 flying at 295 kts groundspeed is followed by a Citation flying at 228 kts. What is the
velocity of the Citation relative to the DC-10?
Solution 2 C-D: 228 kts - 295 kts = -67 kts
The DC-10 is moving away from the Citation at 67 kts.

1.2 Passing head-on

If the two aircraft are approaching on reciprocal tracks then their relative velocities can be calculate
by adding the groundspeeds.
Example 3
An Electra flying at 194 kts groundspeed is approaching an F-27 flying at 212 kts. What is the
relative velocity of the aircraft?
Solution 3 E+F: 194 kts + 212 kts = 406 kts
They are closing at 406 kts!
Further Examples
1. A Cessna 152 travelling at a groundspeed of 116 kts is flying ahead of a Beech Bonanza on the
same track flying at 164 kts. What is the velocity of the Beech relative to the Cessna?

General Navigation (NPA25) - Edition 4 – 110630 21-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
2. A Kingair at 212 kts groundspeed is approaching head on with a Seminole flying at 140 kts
groundspeed. What is their relative velocity?

1.3 Time and point of passing

Knowing the relative velocities and using an initial known distance between them at a given time
then the time at which converging aircraft (either closing from astern or approaching head-on) will
pass can be calculated using the formula : -
Time = Distance
Speed
The distance along the track at which this will take place can then be calculated from the same
formula rearranged to give :-
Distance = Speed x Time
Example 4
An HS 125 passes the BHD VOR beacon at 1607 UTC at a groundspeed of 360 kts. A Gulfstream
is estimating at the BHD at 1611z following the same outbound track at 392 kts groundspeed. At
what time will the Gulfstream pass the HS 125, and how far from the BHD?

1611z Gulfstream HS 125


24 nm

BHD

Figure 21-1 Example 4

Solution 4
Location of the HS 125, at 1611z which is 4 minutes after passing overhead
360 kts x (4 / 60) hrs = 24 nm beyond BHD
Velocity of the Gulfstream relative to the HS 125
G - H: 392 kts - 360 kts = 32 kts
Time to pass
24 nm ÷ 32 kts = 0.75 hours or 45 minutes
So the Gulfstream will pass the HS 125 at 1656z.
Distance to passing point for Gulfstream
0.75 hrs x 392 kts = 294 nm
They pass 294 nm beyond the BHD

Example 5

An Il-76 passes the KONAN waypoint at 1421 UTC, westbound towards the MIMBI waypoint, a
total of 112 track miles away, at 230 kts groundspeed. A Jetstream is flying eastbound towards
MIMBI for KONAN at 185 kts groundspeed, estimating to be at MIMBI at 1427. At what time do
they pass, and at what distance from MIMBI?

21-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
1427z
Jetstream Il-76
23 nm

MIMBI KONAN

112 nm

Figure 21-2 Example 5

Solution 5
Location of Il-76 at 1427z, 6 minutes after passing KONAN
230 kts x (6 / 60) hours = 23 nm from KONAN
Separation of the aircraft at 1427z
112 nm - 23 nm = 89 nm
The relative velocity of the aircraft is
I+J: 230 kts + 185 kts = 415 kts
Time until they pass from 1427z
89 nm ÷ 415 kts = 0.214 hours or 13 minutes (to nearest
minute)
The aircraft pass at 1440z.
Distance moved by Jetstream in this time
185 kts x 0.214 hrs = 39.6 nm
They pass 39.6 nm from MIMBI.
Further Examples
1. An S-61 flying at 90kts groundspeed passes the LIZAD reporting point at 1821z, en route to the
TUNIT reporting point 54 nm from LIZAD. A Jetstream estimates LIZAD at 1833, to track to
TUNIT at a groundspeed of 205 kts. At what time is the Jetstream expected to pass the S-61?
2. Aircraft A is passes reporting point X at 0254z, tracking to reporting point Y at a groundspeed of
423 kts. Aircraft B passes Y at 0248z travelling to reporting point X at a groundspeed of 370 knots.
If X and Y are 238 nm apart then how far from Y do the aircraft pass?

2 Speed Adjustments
Although not strictly relative velocity problems, these are most easily worked out using the same
type of calculation.

A speed adjustment is used to reach a given point - usually a waypoint, beacon or FIR boundary - at
a time specified by ATC. The calculation can either be to select a new speed at which to fly for a
given time, or to select a time at which to change to a given speed. The method used is very similar.

In both cases the first part of the problem is to calculate what would happen if the speed is not
altered, using the formula :-
Distance = Speed x Time

General Navigation (NPA25) - Edition 4 – 110630 21-3


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
The point reached is not where the aircraft should be at that time, and the difference is used to
calculate the speed relative to unaltered speed at which the aircraft must fly, or the time for which
that altered speed must be flown. In each case the same formula should be used, rearranged to give
either speed or time :-
Speed change = Distance
Time
Time = Distance
Speed change
This is most easily shown by example.
Example 6
An aircraft is inbound to the STU VOR at 340 kts TAS, with a headwind component of 14 kts. At
0902 UTC, while still 245 nm out, the Captain is told to reduce speed at his discretion to 270 kts
TAS to arrive at STU at 0953z. As First Officer you are tasked with calculating the time at which to
slow down (assume instantaneous deceleration).
Solution 6
If the aircraft did not decelerate before 0953z then at that time it would have travelled
340 kts - 14 kts = 326 kts groundspeed
326 kts x (51 / 60) hrs = 277 nm
thus overshooting by 32 nm. This means that you must fly 32 nm less by a speed reduction of 70
knots. This drop in distance will therefore take
32 nm ÷ 70 kts = 0.416 hours or 25 minutes
0953z – 25 mins = 0928z
So the speed needs to be reduced at 0928z to arrive at STU at the correct time and speed.
Example 7
An aircraft is approaching the UK at a TAS of 430 kts with a 28 kt tailwind. At 0523 UTC the
estimate for the FIR boundary is 0609. ATC instruct the aircraft to reduce speed to make good the
FIR boundary at 0613. The captain asks you to calculate the TAS required.
Solution 7
On current estimate you expect to be 4 minutes beyond the FIR boundary at 0613, flying at a
groundspeed of 458 kts (430 kts + 28 kts wind).
458 kts x (4 / 60) hrs = 30.5 nm
The time available to lose 30.5 nm is 50 minutes (from 0523z to 0613z). The speed reduction has to
be
30.5 nm ÷ (50 / 60) hrs = 36.6 kts ≈ 37 kts
So the required groundspeed is
458 kts – 37 kts = 421 kts
With a 28 kt tailwind component the TAS required is
421 kts – 28 kts = 393 kts
Further Examples
1. A Boeing 727 is approaching the hold at a groundspeed of 235 kts. The captain’ s current estimate
overhead the beacon is 1234. ATC ask the pilot to slow to a groundspeed of 205 kts to enter the
hold at 1237. How far out should he slow down, assuming instant deceleration?
2. Flying a Seminole at 140 kts TAS, with a 5-kt tailwind component you are cleared at 1515 to
enter an airway at 1543. If your current estimate for entering the airway is 1539 and assuming that
you slow down immediately what TAS should you slow to?

21-4 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
3 Summary
Relative velocity problems involving two aircraft are reasonably simple to solve. If they are
travelling in the same direction, subtract the lead aircraft’ s groundspeed from the trailing aircraft’ s
groundspeed, if they are approaching head-on then add the groundspeeds together.

If the time of passing is required then simply divide the distance between the aircraft at a known
time by this speed. To calculate the position at which they pass find the time before they pass and
calculate where one of the aircraft will be at that time.

Calculation of the speed or time requirement to reach a point at given time is more complicated.
First calculate what would have been the case with no change in speed, then use this to calculate the
change required. Then use the equation Distance = Speed x Time, rearranged to give the required
answer, either a speed change or a time at which to change speed.

Remember if the question is complicated draw a diagram to lay out the information in a logical
manner!

General Navigation (NPA25) - Edition 4 – 110630 21-5


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

21-6 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 22 Critical Point / Point of Equal Time

Refer to Chapter 15 of the Flight Planning and Monitoring manual from Part II of the course. This
subject is taught in the Flight Planning syllabus but may be examined in General Navigation, either
in practice papers during the course or in the final Exam.

General Navigation (NPA25) - Edition 4 – 110630 22-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

22-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 23 Point of Safe Return, Point of No Return and Radius of
Action

Refer to Chapter 15 of the Flight Planning and Monitoring manual from Part II of the course. This
subject is taught in the Flight Planning syllabus but may be examined in General Navigation, either
in practice papers during the course or in the final Exam.

General Navigation (NPA25) - Edition 4 – 110630 23-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

23-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Section Six Navigation Instruments

Chapter 24 Magnetic Instruments

Refer to Chapter 8 of the Instruments manual from Part I of the course. This subject is taught in the
instruments syllabus but may be examined in General Navigation, either in practice papers during
the course or in the final Exam.

General Navigation (NPA25) - Edition 4 – 110630 24-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

24-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 25 Inertial Navigation Systems

Refer to Chapter 9 of the Instruments manual from Part I of the course. This subject is taught in the
instruments syllabus but may be examined in General Navigation, either in practice papers during
the course or in the final Exam.

General Navigation (NPA25) - Edition 4 – 110630 25-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

25-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
Chapter 26 Flight Management System (FMS)

Refer to Chapter 8 of the Instruments manual from Part I of the course. This subject is taught in the
instruments syllabus but may be examined in General Navigation, either in practice papers during
the course or in the final Exam.

END OF NOTES

General Navigation (NPA25) - Edition 4 – 110630 26-1


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011
THIS PAGE IS INTENTIONALLY BLANK

26-2 General Navigation (NPA25) - Edition 4 – 110630


©BCFT, a trading name of Bournemouth Flying ClubTM Ltd 2011

You might also like