Vous êtes sur la page 1sur 21

THE POLAR FRONT JET STREAM

PART I STRUCTURE OF THE JET STREAM

I. GENERAL

The Jet Stream, like some of the major ocean currents, is a permanent feature vary-ing in strength and
location in both hemispheres. Since it is a velocity phenomenon it is best described in terms of isovels
or isotachs, which are lines of equal wind velocity without regard to wind direction. These new terms
have in recent years become almost as common as isobars, contours and streamlines, which have been
used to describe the wind field in the atmosphere.

Jet Stream Terminology:

Isovels
Lines on a vertical cross section drawn through points of equal wind velocity.

Isotachs
Lines on a horizontal weather chart drawn through points of equal wind velocity.

Isobars
Lines on a constant level weather chart drawn through points of equal pressure.

Contours
Lines on a constant pressure chart drawn through points of equal height of the constant pressure sur-
face above sea level.

Isotherms
Lines of horizontal charts or vertical cross sections drawn through points of equal temperature.

ZOMCOT
Zone of maximum Concentration of temperature

Tropopause (Polar and Tropical)


Boundary between Troposphere and the Stratosphere.

Shear
Rate of change of wind velocity (or temperature), vertically or horizontally.

Core
The center line of maximum wind velocity in the Jet Stream.

Jet Stream
The three dimensional volume of air moving with velocities in ex-cess of an arbitrary value, usually
taken as being greater than 70 KTS.

1
Jet Stream Wind
On any particular horizontal chart, the lines of maximum wind MAX velocity.

II. EXPLANATION

The simplest explanation of the Polar Jet Stream is Confluence: the bringing together of cold and warm
air masses, which, under certain conditions, results in a sharp fron-tal boundary aloft. The greater the
frontal temperature discontinuity, the stronger the jet. This frontal zone aloft has been labeled the ZOM-
COT, which means “Zone of Maximum Concentration of Temperature.” This temperature contrast brings
about a strong pressure gradient that causes an acceleration in the warmer, lighter air adja-cent to and
usually south of the ZOMCOT.

This wind flow extends aloft, increasing in strength as long as the polar air remains colder than the adja-
cent air mass. Once the Polar Tropopause is reached, and at some level above, the horizontal thermal
difference ceases. Here, the core of the jet is reached. Above the jet core the temperature field reverses,
which causes an abrupt decelera-tion of the wind. This deceleration above the core is approximately
three to four times as great as the acceleration below the core.

III. THERMAL STRUCTURE

Horizontal
Much has already been written and said about temperatures in and near jet streams, but usually only in
connection with the medium altitudes (15 25,000 feet). Here, an at-tempt will be made to cover thermal
aspects of the jet from 700 mbs. to 100 mbs. By this approach it is thought a better understanding of the
jet can be developed.

Each Polar Jet has an associated surface front, because any front aloft will have its foundation some-
where on the surface, even though at times it may be weak and bare-ly discernible. Surface fronts usu-
ally have marked ZOMCOTS, but are so much out of phase with the isobars cross flow that the neces-
sary conditions are not met for jet flow at levels below 10,000 feet. This is due mostly to the rapid
change of wind direction at low altitudes. Occasionally, under freak conditions along strong cold fronts,
the COMCOT/Contour relationship necessary for jet formations (in phase) materialize at very low levels
(4 8,000 feet), giving winds up to 80 knots as low as 2,O00 feet and even 60 70 knots on the surface.

700 mbs. is usually about the lowest level in which a ZOMCOT, in phase with the con-tours, is found.
Generally, this is too low for marked ZOMCOTS and almost always a slight cross flow is noticeable.

Notice the crowding of the isotherms at the 500 400 mb. level. Roughly, the core of the polar jet is
found approximately 12,000 feet above the level of maximum horizon-tal temperature gradient. For very
strong jets in winter, this level is near 500 mb., with the core at 30 32,000 feet. For moderate to weak
streams, this level will be nearer the 400 mb. Ievel with the core 34 36,000 feet. For the average North
Pacific jet, wind maximum at 500 400 mbs, the temperature will decrease approximately one degree
centigrade in 10 miles through the ZOMCOT toward the cold air, but will be as much as one degree cen-
tigrade in six miles along strong jet streams. The temperature gra-dient is much weaker south of the
ZOMCOT, averaging approximately one degree cen-tigrade in 35 miles.

2
3
The horizontal temperature gradient decreases in magnitude fairly rapidly above 400 mb, giving gener-
ally on a weak ZOMCOT at 300 mbs, averaging approximately one degree centigrade in 20 miles. Occa-
sionally with strong jets, whose cores are near 300 mbs, a horizontal temperature gradient is absent and
the temperature is uniform over a wide area. Thus, from slightly above 300 mbs to approximately 250
mbs, near the jet core, a ZOMCOT is not evident. At 250 mbs the jet is usually accompanied by a nar-
row band of colder air that is only one to two degrees centigrade colder than the air to north and south.
See Figure 1A. Above the jet core, the ZOMCOT re-establishes itself, but in reverse order, i.e, warm air
north—cold air south. The ZOMCOT above the core is more pronounced from the 200 to 150 mb Ievel
and averages approximately one degree centigrade in 10 miles. Above 150 mbs it begins to weaken and
generally fades out at 100 mbs. Above the core, the cold air contains the maximum wind belt.

Vertical
Five to 10 degrees of latitude, south of the jet core, a typical temperature sounding (RAOB) shows a
fairly large, steep lapse rate in the shallow cold air below the frontal surface with a sharp inversion upon
penetrating the frontal surface aloft. This inver-sion averages 6 8°C. Within the warm air above the
front, a near normal lapse rate, 2°C/1000 feet, is found until the warm air tropopause is reached at ap-
proximately 40,000 feet, where a sharp increase is usually found. Figure 4, sounding A is a typical soun-
ding at this point. Farther north toward the jet at sounding B, the cold air becomes deeper and the inver-
sion at the frontal intersection is found higher and weaker— averaging approximately 2 4°C. Once the
warm air is reached, a normal lapse rate is again established until the warm air tropopause is
intercepted. At C, near the jet core, the frontal intersection is found at still higher levels, with the inver-
sion becoming a near isothermal condition. This high level frontal inversion is often mistaken for a tro-
popause surface. Within the cold air, north of the jet, a typical sounding is represented by D. It is safe to
assume that an inversion above 30,000 feet is a tropopause surface rather than a frontal surface, espe-
cially if a wind decrease is noticed at this height.

IV. WIND STRUCTURE

Horizontal
While no two jet streams are exactly alike in respect to strength, wind shear, etc, most of them conform
to a typical pattern to the extent that some generalizations can be made in regard to wind shears.

All Polar jet streams in the Pacific have much greater wind shear towards the warm air below the jet
core. In most cases, the shear to the north (cold air) is approximately six to seven times the shear to the
sough (warm air) near the 500 400 mb Ievel. At jet core level, the shear to the north might be as much
as 15 times greater than towards the south. At present operating levels, 15 25,000 feet the shear north
will average approximately one knot per mile at 500 MB and up to near 1.5 knots per mile at 25,000 feet.
To the south (warm air), the shear is much less, averaging approx-imately one knot in six miles at 500
mb, to one knot in four miles at 400 mb.

4
Wind Shear, Knots per Unit Distance
COLD SIDE WARM SIDE
500 mb. 1 kt/mi. 500 mb. 1 kt/6 mi
400 mb. 1.5/mi. 400 mb. 1 kt/4 mi.
250 mb. 3 kt/mi. 250 mb. 1 kt/2 mi.

The above figures are only general averages and variations are great along the same jet stream and
from day to day. The importance of the extreme wind shear found to the north of the average jet stream
cannot be stressed too much.

Vertical
The vertical wind shear through the frontal surface increases as the polar front aloft becomes higher.
The maximum increase will be found in the 15 25,000 feet altitude range as the frontal surface is pene-
trated. Above the jet core, the decrease in wind is always much faster than the increase below the core
and may amount to as much as 30 to 40 knots per 1000 feet (see Figure 3). Below are listed averaged
increases in wind velocity with altitude through the frontal zones.

ALTITUDE INCREASE IN WIND (kts) per 1,000 feet


10,000 FEET 5 KNOTS
15,000 FEET 8 KNOTS
20,000 FEET 12 KNOTS
25,000 FEET 15 KNOTS

V. THE JET STREAM MODEL


Figure 3 shows a jet stream model drawn to a scale of 1 to 20, that is, one mile in the vertical equals 20
miles in the horizontal. Therefore, the true picture would be seen if the model was stretched 20 times its
length. This would give a ribbon appearance rather than a turbular shape. The wind shear shown to the
left of the core in Figure 3 is approximately twice that of the average jet. This model in appearance is
typical of most jet streams. All will follow the same pattern insofar as relative wind shear, temperature
shears, tropopauses etc, are concerned.

At least 90% of the jet streams will give cloud free, smooth flight above 15,000 feet. The typical cloud
structure is also shown in Figure 3. All frontal clouds are well south near the surface front. The cold air
cumulus, when over warmer water, occasionally becomes towering cumulus and sometimes even minia-
ture CBs are found near posi-tions A with tops 20 30,000 feet. Only one case is on record where fairly
large CBs were found just south of the jet near the Azores and quite far north of the surface front. Jet
stream cirrus is often found approximately 100 150 miles south of the jet core near the level where the
vertical wind shear reverses. This distance, of course, varies with individual jets. Quite frequently, the
skies are completely clear.

South of the jet core there is ascending air, north descending air, as shown in Figure 4. Figure 4 shows
the areas of probable clear air turbulence due to wind shear. Air-craft are constantly losing airspeed in
the cold air due to the subsidence. This can amount to 10 15 knots when holding constant pressure
altitude. This descending air seems greatest just north of a cyclonic curved jet stream. The warm pocket
of air found just north and above the jet core is proof of the subsidence below.

5
6
VI. CLIMATOLOGY

Since the polar jet stream is associated with the polar front, its seasonal and geographical variations are
similar to those of fronts.

Jet steams follow fronts around Lows. The mean positions of troughs and Lows is perhaps a better cli-
matological guide to the mean positions of jets than the North/South displacement of the Polar front.
Some seasons (generally winter) are dominated by large stagnant Lows in the North Pacific, which
cause horseshoe shaped jets. In other seasons (summer, generally), straight

VI. CLIMATOLOGY

Since the polar jet stream is associated with the polar front, its seasonal and geographical variations are
similar to those of fronts.

Jet steams follow fronts around Lows. The mean positions of troughs and Lows is perhaps a better cli-
matological guide to the mean positions of jets than the North/South displacement of the Polar front.
Some seasons (generally winter) are dominated by large stagnant Lows in the North Pacific, which
cause horseshoe shaped jets. In other seasons (summer, generally), straight East/West (zonal) flow
predominates, which produces jets aligned more closely to our tracks even though they are usually
weaker than the horseshoe types.

VII. SUMMARY

The Jet Stream is a year round phenomenon, being much stronger in winter and far-ther south in sum-
mer.

Summer jets are relatively weak and move northward with the polar front. The strongest North Pacific
jets will be found in the western Pacific off the Asian continent and are apt to be of the cyclonic type.

The Jet Stream and the Tropopause Boundary


The tropopause is that boundary between the tropopause (lower atmosphere) and the stratosphere
(higher atmosphere). It is actually the absolute height at which the at-mosphere ceases to be affected by
surface heating. Over the equator, where there is a terrific amount of heating possible, it is reasonable to
expect that a deeper layer of air is affected. Therefore, this boundary must necessarily be higher, and as
we pro-ceed towards the poles, the boundary between the "affected" and "non affected" air is lower.

The height of the tropopause is established by definition. It is defined as the lowest level at which the
lapse rate of the air becomes less than 2°C/3000 feet. (Roughly 1 °F/1000 feet) low lapse rates imply
stability, or a resistance of air to ascend or des-cend. In effect, the tropopause marks the boundary
where due to stability the at-mosphere is refusing to accept any vertical currents. The level is deter-
mined from accents made by the Radiosonde Balloons. In Figure 2, we show typical soundings made in
the polar and equatorial regions. It is easy to see from these soundings, and because of the definition
used to define what is considered the boundary, that the air over the equator, which refuses vertical cur-
rents, is much colder than that over the poles.

7
Research meteorologists also discovered that this boundary was not always definitely defined. In other
words, the change of lapse rate from something greater than 1 °F/1000 feet to something less than 1°
F/1000 feet in some instances requires considerable depth, while in other cases it was very short and
abrupt, even to the extent that an inver-sion may be present. In other words, resistance to vertical cur-
rents at the boundary is either spread out throughout a considerable vertical depth or it becomes highly
resis-tant in a relatively short vertical depth.

Supposing then we take a situation where the boundary is very sharp, that is, sudden-ly there exists an
altitude where the atmosphere accepts absolutely no vertical cur-rent. At the same time this boundary
exists, considerable vertical accelerations are introduced in the lower layers of the atmosphere. To make
the picture simple, we can imagine a terrific convective cell forming someplace over the equator with a
terrific volume of air moving vertically. It is easy to see that as soon as these currents reach the tro-
popause, the current must necessarily begin to flow laterally beneath the stable air. Now if we imagine a
little further and permit the convective call to move northward, it is easy to see that if the tropopause
boundary lowers, we are in effect running into a “squeeze" play with the only alternative being that the
lateral movement must in-crease in velocity.

This explanation satisfies the fact that high velocities exist, but it doesn't satisfy the existence of the nar-
row bands. Further research made the discovery that the tropopause was not continuous, but that there
were areas where it just seemed to stop and take over again at some lower attitude. These regions
marked the so called jet stream. It seems quite clear then, that as the energy is passed northward, it is
not a smooth running affair but goes in jumps. There is the possibility that what we are seeing then is a
critical point at which the resistance of the air above the tropopause is overcome and the jet is the
breaking through action. The polar front very likely is acting as a block to further prevent northward
movement of the energy causing higher concentra-tions of energy to exist in a localized area. Notice in
the cross section how the tropopause boundary breaks and the location of the jet core in relation to that
break.

The Tropopause and Flying Conditions


The tropopause boundary is a place where vertical accelerations must stop and begin to flow beneath
the stable air of the stratosphere. Furthermore, it can be of a sharp boundary or a diffuse boundary. It is
natural to assume that if the change is abrupt, it cannot be accomplished without creating some turbu-
lence. It has already been pro-ven that with most penetrations some turbulence does exist, and, at
times, it is severe enough to be operationally important. The degree of turbulence is subject to two
items: the sharpness of the boundary and the amount of vertical acceleration present. It is also reason-
able to expect that even if no jet activity is evident, flight close to the boun-dary ahead of the surface
intersection of the polar front has a good chance of encounter-ing turbulence. The lack of a jet merely
indicates that the stable stratosphere is holding, but it doesn't eliminate the possibility that many forces
are at work.

In some cases, nature will help you locate the exact boundary of the tropopause. The stratosphere
above the tropopause is very stable air, so any pollution, which has had occasion to be carried aloft to
that altitude, will form a layer just below the stable air. The boundary, therefore, is in the haze layer. The
presence of the haze layer has been verified many times and apparently is more dense than we would
expect. Pilot reports of zero horizontal visibility are numerous. Very often, this layer cannot be seen from
the ground.

8
Jet Operations Meteorological Notes for Pilots

CLEAR AIR TURBULENCE


I. CLEAR AIR TURBULENCE — DEFINITION AND GENERAL DISCUS-
SION
The term clear air turbulence is self defining. The use of this term came along naturally with the devel-
opment of higher flying aircraft. Turbulence at non pressurized aircraft levels occurred mainly within
clouds. As aircraft began to operate at higher levels, turbulence not associated with clouds was more
frequently encountered. Initially these encounters were considered anomalous because clouds were not
present, and the term clear air turbulence was used to emphasize the fact that the turbulence was not
associated with clouds.

Turbulence, as experienced in an aircraft, is generally caused by a combination of atmospheric instabil-


ity (resulting in rising and descending currents in close proximity to one another) and excessive wind
shear, i.e., a large change of wind direction and/or wind speed with altitude.

There are no routes where clear air turbulence may not be encountered. However, there are preferred
conditions (jet streams) and regions (mountainous areas) where the frequency of clear air turbulence is
greatest. The cause of the clear air turbulence in each situation is discussed below.

II. CAUSES OF CLEAR AIR TURBULENCE


The main causes of clear air turbulence are:

High rates of change of wind speed with altitude. This is called vertical wind shear. It is this effect which
produces clear air turbulence in the tropopause, upper fronts and trough lines.

Disturbed air flow over mountainous terrain. This is referred to as the mountain wave.

III. CLEAR AIR TURBULENCE DUE TO WIND SHEAR


Since the wind shears necessary to produce clear air turbulence are most often met under jet stream
conditions, this phenomenon will be discussed first. The typical distribu-tion of wind speed and vertical
wind shear which produces clear air turbulence is il-lustrated in Figure 1.

9
Note the large changes in wind speed with altitude (vertical wind shear) through the Front in the tro-
popause and through the tropopause. The change in wind speed through the tropopause at times may
equal 100 knots in 5,000 feet. In general, it has been found that when the wind direction remains fairly
constant with altitude (less than 30° change per 1,000 feet), an increase in wind speed in excess of 6
knots per 1,000 feet is sufficient to produce clear air turbulence.

Clear air turbulence occurs whenever the necessary wind shear conditions exist, mainly along frontal
boundaries and long the tropopause.

10
Usually the large wind speed increases (positive shear) at any given level below the core of the jet
stream are concentrated into narrow regions, mainly in frontal boun-daries. At any given level above the
core of the jet stream, and especially near the tropopause, the area of large wind speed decreases
(negative shear) is more widespread. Consequently, clear air turbulence at higher altitudes (35,000 feet
and above) is also apt to be more widespread, but not necessarily more intense than at levels below the
core of maximum winds.

Through the core, usually located between 30,000 and 35,000 feet (see Figure 4), the wind speeds are
very high with very little change in wind speed from level to level. This lack of vertical wind shear is due
to the very small thermal gradient at jet stream core levels. Therefore, the incidence of turbulence in the
core of stronger winds is relatively small when compared to the incidence of turbulence in those regions
out-side the core possessing less strong winds, but having strong horizontal thermal gra-dients and as a
result, large vertical shear.

Cigar air turbulence may be present along a trough line where the wind direction changes rapidly. How-
ever, there is little evidence that the rapid change in wind direc-tion alone results in clear air turbulence.
Since a sharp upper air trough is a preferred area for jet stream activity (with associated strong vertical
wind shear) it is felt that the cases of clear air turbulence, which have been reported in the area shown
in Figure 5, were the result of undetected vertical wind shear.

IV. CLEAR AIR TURBULENCE DUE TO MOUNTAIN WAVES


A mountain wave or standing wave is a disturbance in the atmosphere set up by a mountain barrier and
characterized by a wave like airflow, producing strong vertical currents and severe turbulence. This phe-
nomena is present over the Polar Route in the western U.S. and Canada. The clear air turbulence asso-
ciated with the mountain wave is caused by the Rockies disturbing the normal wind flow.

Mountain waves may form when three conditions are satisfied:

Air flow is normal to the range with a speed exceeding 25 knots and, more com-monly 50 knots, at
mountain top level.

Wind speed increases with altitude at mountain top level.

The temperature sounding for upwind stations usually shows a shallow inversion just above the surface
and there is a tendency towards stability between 10,000 and 13,000 feet.

Figure 6 illustrates the main features of the mountain wave. The most turbulent area is centered around
the roll cloud at levels mainly below 20,000 feet. When jet stream conditions are present, which is mainly
in the winter months, there is also a turbulent layer at the tropopause level. In between these levels, in
the laminar flow characteriz-ed by lenticular clouds, the turbulence consists mainly of relatively smooth
up and down drafts.

Fortunately, characteristic cloud patterns give visual evidence to the pilot in the estimated 90% of the
cases. Most characteristic is the standing lenticular (lens shaped) cloud in bands to the lee of the moun-
tain. The lenticulars are generally found at 20,000 feet or above and may be in massive layers. They
appear stationary, but are undergo-ing formation on the windward edge and dissipation on the leeward.

11
12
V. INTENSITY, AVOIDANCE, AND OTHER NOTES
CIear air turbulence will be most intense in the tropopause surface just above the jet core and at the up-
per front on the cold side of the jet stream below the core level.

Initial flight planning will take into account avoidance of areas of forecasted clear air turbulence. How-
ever, to avoid or minimize clear air turbulence, the following courses of action are suggested:

Make a lateral diversion if close to the core level.

Change altitude if along the front or the tropopause.

Use a combination of both along steeply sloping portions of the front or tropopause.

Reduce speed if unable to accomplish above.

Other factors of importance in minimizing the effects of clear air turbulence are:

In the vicinity of strong winds when climbing through cirrus clouds, be prepared for clear air turbulence
just on top of the layer or a few thousand feet above the cloud layer.

When approaching a cirrus deck from the low pressure side of the jet stream, or leaving cirrus clouds
toward the low pressure side of the jet stream, be prepared for turbulence in the clear air at some
distance (50 100 miles) away from the edge of the cloud deck. This zone corresponds to the zone
of strongest wind shear on the cold side of the jet axis.

To assist the meteorologist in locating and identifying the causes of areas of clear air turbulence, please
report all occurrences of clear air turbulence (even light) and associated temperatures.

Similarly, lateral deviations must be carefully evaluated since the maximum wind is generally found just
below the tropopause, and a change of course on a flight in a condition of the tropopause, which is
sharply slanted downward, could result in en-countering strong tropospheric winds, aiding and retarding,
depending upon direc-tion of flight in relation to the location of the pressure system. (See FLIGHT in Fig-
ure 7.) If the FLIGHT continues on course, the strong winds (J) are avoided by flight through the strato-
sphere. A heading change to the south avoids the tropopause, but encounters the strong winds. How-
ever, in a flight from west to east, in the opposite direction from the FLIGHT in Figure 7, an original flight
plan or a change in course could well take advantage of the opportunity to remain in the troposphere
without a tropopause penetra-tion and take full advantage of the jet stream winds.
In those regions where the tropopause is everywhere above 40,000 feet, this boun-dary is of little con-
cern insofar as jet aircraft operations in the very near future are concerned.

• In those areas over the globe where the temperature at 300 mbs (close to 30,000 feet) is warmer
than about 35°C, the tropopause is usually everywhere above 43,000 feet.

In those areas over the globe where the temperature at 300 mbs (close to 30,000 feet) ranges from
about 36°C to 41 °C, the tropopause lowers gradually from about 42,000 feet to 38,000 feet.

13
In those areas over the globe where strong westerly winds (jet streams) are associated with intense
horizontal temperature gradients between temperatures of about 42°C to 50°C at 300 mbs. the
tropopause will drop abruptly from about 38,000 feet at the warmer temperature to about 30,000 feet
in the colder temperature.

In those areas over the globe where the temperature at 300 mbs is 50°C and col-der, the tropopause
boundary is close to 30,000 feet or below.

The rate of change of temperature decrease with altitude diminishes as aircraft climb into the tro-
popause. On occasion, the temperature may even increase slightly as one goes from the tropo-
sphere into the tropopause.

A sharp rise in temperature, however, is experienced by an aircraft in level flight in the troposphere as
the aircraft crosses a tropopause boundary, which is sloping down to and below the flight altitude.

Since it is impossible to show the three dimensional aspects of the slope of the tropopause on a fore-
cast chart, a visual picture of the slope must be kept in mind at all times.

14
JET 3 OPERATIONS METEOROLOGICAL NOTES
FOR PILOTS
TROPOPAUSE

I. TROPOPAUSE—DEFINITIONS, GENERAL LOCATIONS AND CON-


FIGURATIONS
The Troposphere is the lower part of the atmosphere which normally is characterized by a regular de-
crease of temperature with altitude.

The Tropopause is the boundary layer that separates the troposphere from the stratosphere. The regu-
lar decrease of temperature with height changes at the tropopause, and this change, is the prime mete-
orological factor of identifying the tropopause. By meteorological definition, the tropopause is located at
that initial height in the atmosphere where a vertical sounding shows that the lapse rate becomes one
third or less of the normal lapse rate and continues at that rate for an increment of at least 6,000 feet.

The Stratosphere is the upper part of the atmosphere in which normally the temperature no longer de-
creases with altitude. Temperatures may remain fairly cons-tant or even increase with altitude in this
region.

• COAST/HONOLULU (particularly along the eastern half of the route and in the winter half of the
year)

• POLAR ROUTES and ALASKA (all seasons)

• TOKYO/WAKE, TOKYO/HONOLULU and TOKYO/HONG KONG (particularly in the TYO area in the
winter half of the year).

• SYDNEY and AUCKLAND/NANDI (particularly along the southern half of the route in the southern
hemisphere winter)

• NORTH PACIFIC ROUTES (all seasons)

The tropopause will usually not be penetrated on the CENPAC route west of Honolulu, the SOPAC route
Honolulu to Handi and the Asiatic Routes. The tropopause exists over these routes, but will be found
above the maximum flight level. (See Figure 8.) As with other meteorological conditions the height of the
tropopause varies with the seasons generally being lowest in winter and highest in summer.

15
Meteorological phenomena associated with the tropopause and pertinent to flight operations:

• Clear Air Turbulence at or near the tropopause due to large wind shear.

• Increase in wind velocity with altitude up to the tropopause and decrease in veloci-ty above the tro-
popause.

• Decrease of wind velocity above the tropopause is usually much greater than the increase below the
tropopause.

• Decrease in temperature with altitude up to the tropopause and then decrease in temperature at a
slower rate, no change in temperature, or increase in temperature above the tropopause.

• Clouds of vertical development are limited by the tropopause.

Since all of these implications are interrelated with regard to flight planning and con-duct, no one should
be considered without full evaluation of the others. These con-siderations are discussed in detail in the
following paragraphs.

• The tropopause layer is generally continuous as shown in Figure 8, but it rises in steps from the arc-
tic to the tropics as shown in Figure 9.

• As shown in Figure 9, jet streams may be imbedded in or below the tropopause leaves between the
various air masses and in the regions of jet streams the tropopause may be discontinuous or dip
down sharply. See Figure 10 which is an enlarged sec-tion of the far right portion of Figure 9.

16
• The effect of the tropopause layer, which separates the troposphere from the stratosphere, is similar
to that of a frontal surface with respect to temperature changes and wind shear. In fact, the tro-
popause often merges with the upper portions of the lower level fronts. See Figure 10.

• The tropopause if found at lower altitudes in troughs than in ridges. In such cases, the low as it
moves within the synoptic patterns carries its associated low tropopause with it. Thus, penetration
through the tropopause into the stratosphere may be ex-pected at lower altitudes near the centers of
lows or in troughs. Also, on a given horizontal plane, which is not in the tropopause at the outset of a
flight, the tropopause may be encountered even if the flight proceeds southward in the vicinity of
lows or deep troughs. See Figure 11.

• Light to moderate (with possible heavy) clear air turbulence may occur in the tropopause layer. This
may occur during climb, descent or cruise if flight plan is along or in the tropopause layer. See
Cross Section portion of Figure 11. Evasive action to avoid turbulence associated with the tro-
popause layer can involve either a change in altitude or a change in course or both. Some points to
remember are:

♦ The tropopause may be a surface, a thin layer, or even a quite deep layer (as much as
6,000 feet) and other operational considerations may not per-mit complete vertical avoid-
ance of the deeper layers.

♦ The tropopause may slope upward in the direction of flight. Climbing may prolong exposure
if flight is along or in the surface. The converse is also true. (See Figure 12.) Trip ONE flying
from north to south under the tropopause elects to cruise climb at A and due to upward
slope of tropopause, climbs into the tropopause and then continues to fly in the tropopause.
No climb at A would have kept Trip ONE in the troposphere for the entire flight.

♦ On maximum performance flights, a climb into the stratosphere may bring warmer tempera-
tures with loss of fuel economy. Appreciable increases in temperature in the stratosphere
above the tropopause may occur in high level lows where sinking in the stratosphere causes
warming and a warm pocket is formed. Along the surface of the tropopause temperatures
will normally be colder than just above or just below, and flights may be planned in this area
to effect best performance.

♦ Loss of an aiding wind or increase of a retarding wind may occur during a climb or descent
to avoid turbulence associated with the tropopause. It is necessary to know the distribution
of wind with height since in some cases this distribution is not normal and a double wind
maximum below the tropopause could exist instead of the normal single maximum.

17
18
19
20
21

Vous aimerez peut-être aussi