Vous êtes sur la page 1sur 35

2.

1 PRECIPITATION

Introduction

Engineering hydrology takes a quantitative view of the hydrologic cycle.


Generally, equations are used to describe the interaction between the
various phases of the hydrologic cycle. As shown in Chapter 1, the following
basic equation relates precipitation and surface runoff:

Q=P-L (2-1)

in which Q = surface runoff, P = precipitation; and L = losses, or hydrologic


abstractions. The latter term is interpreted as the summation of the various
precipitation-abstracting phases of the hydrologic cycle.

Rainfall is the liquid form of precipitation; snowfall and hail are the solid
forms. In common usage, the word rainfall is often used to refer to
precipitation. Exceptions are the cases where a distinction between liquid
and solid precipitation is warranted.

Generally, the catchment has an abstractive capability that acts to reduce


total rainfall into effective rainfall. The difference between total rainfall and
effective rainfall is the losses or hydrologic abstractions. The abstractive
capability is a characteristic of the catchment, varying with its level of stored
moisture. Hydrologic abstractions include interception, infiltration, surface
storage, evaporation, and evapotranspiration. The difference between total
rainfall and hydrologic abstractions is called runoff. Therefore, the concepts
of effective rainfall and runoff are equivalent.

The terms in Eq. 2-1 can be expressed as rates (millimeters per hour,
centimeters per hour, or inches per hour), or when integrated over time, as
depths (millimeters, centimeters or inches). In this sense, a given depth of
rainfall or runoff is a volume of water uniformly distributed over the catchment
area.

Meteorological Aspects

The earth's atmosphere contains water vapor. The amount of water vapor
may be conveniently expressed in terms of a depth of precipitable water.
This is the depth of water that would be realized if all the water vapor in the
air column above a given area were to condense and precipitate on that area.

There is an upper limit to the amount of water vapor in an air column. This
upper limit is a function of the air temperature. The air column is considered
to be saturated when it contains the maximum amount of water vapor for its
temperature. Lowering the air temperature results in a reduction of the air
column's capacity for water vapor. Consequently, an unsaturated air column,
i.e., one that has less than the maximum amount of water vapor for its
temperature, can become saturated without the actual addition of moisture if
its temperature is lowered to a level at which the actual amount of water
vapor will produce saturation. The temperature to which air must be cooled,
at constant pressure and water vapor content, to reach saturation is called
the dewpoint. Condensation usually occurs at or near saturation of the air
column.

Cooling of Air Masses. Air can be cooled by many processes. However,


adiabatic cooling by reduction of pressure through lifting is the only natural
process by which large air masses can be cooled rapidly enough to produce
appreciable precipitation. The rate and amount of precipitation are a function
of the rate and amount of cooling and of the rate of moisture inflow into the
air mass to replace the water vapor that is being converted into precipitation.

The lifting required for the rapid cooling of large air masses is due to four
processes [72]:

1. Frontal lifting,
2. Orographic lifting,
3. Lifting due to horizontal convergence, and
4. Thermal lifting.

More than one of these processes is usually active in the lifting associated
with the heavier precipitation rates and amounts.

Frontal lifting takes place when relatively warm air flowing towards a colder
(hence denser) air mass is forced upward, with the cold air acting as a wedge
(Fig. 2-1 (a)). Cold air overtaking warmer air will produce the same result by
wedging the latter aloft. The surface of separation between the two different
air masses is called a frontal surface. A frontal surface always slopes upward
toward the colder air mass; the intersection of the frontal surface with the
ground is called a front .

Orographic lifting occurs when air flowing toward an orographic barrier (i.e.,
mountain) is forced to rise in order to pass over it (Fig. 2-1 (b)). The slopes
of orographic barriers are usually steeper than the steepest slopes of frontal
surfaces. Consequently, air is cooled much more rapidly by orographic lifting
than by frontal lifting.
Fig. 2-1 (a) Frontal lifting. Fig. 2-1 (b) Orographic lifting.

Lifting due to horizontal convergence is also important in the production of


clouds and precipitation. Convergence occurs when the pressure and wind
(velocity) fields act to concentrate inflow of air into a particular area, such as
a low-pressure area (Fig. 2-1 (c)). If this convergence takes place in the lower
layers of the atmosphere, the tendency to pile up forces the air upward,
resulting in its cooling. Even when precipitation does not result from
convergence alone, subsequent precipitation triggered by other processes
may be more intense if convergence has occurred.

Thermal lifting is caused by local heating. As heated surface air becomes


buoyant, it is forced to rise, resulting in its cooling. If the local heated air
contains enough moisture and rises far enough, saturation will be reached
and cumulus clouds will form (Fig. 2-1 (d)). Thermal lifting is more
pronounced in the warm season. Rainfall associated with thermal lifting is
likely to be scattered in geographic extent. In flat country, the greatest
convective activity is over the hottest surfaces; in mountain country, it is
greatest over the highest peaks and ridges.
Fig. 2-1 (c) Lifting due to horizontal convergence. Fig. 2-1 (d) Thermal lifting.

Condensation of Water Vapor into Liquid or Solid Form. Condensation


is the process by which water vapor in the atmosphere is converted into liquid
droplets or, at low temperatures, into ice crystals. The results of the process
are often, but not always, visible in the form of clouds, which are airborne
liquid water droplets or ice crystals or a mixture of these two.

Saturation does not necessarily result in condensation. Condensation nuclei


are required for the conversion of water vapor into droplets. Among the more
effective condensation nuclei are certain products of combustion and salt
particles from the sea. There are usually enough condensation nuclei in the
air to produce condensation when the water vapor reaches saturation point.

Growth of Cloud Droplets and Ice Crystals to Precipitation Size. When


air is cooled below its initial saturation, so that temperature and condensation
continues to take place, liquid droplets or ice crystals tend to accumulate in
the resulting cloud. The rate at which this excess liquid and solid moisture is
precipitated from the cloud depends upon: (1) the speed of the upward
current producing the cooling, (2) the rate of growth of the cloud droplets into
raindrops heavy enough to fall through the upward current, and (3) a
sufficient inflow of water vapor into the area to replace the precipitated
moisture.

Water droplets in a typical cloud usually average about 0.01 mm in radius


and weigh so little at an upward current of only 0.0025 m/s is sufficient to
keep them from falling. Although no definite drop size can be said to mark
the boundary between cloud and raindrops, a radius of 0.1 mm has been
generally accepted. The radius of most raindrops reaching the ground is
usually much greater than 0.1 mm and may reach 3 mm. Drops larger than
this tend to break into smaller drops because the surface tension is
insufficient to withstand the distortions the drop undergoes in falling through
the air. Drops of 3 mm radius have a terminal velocity of about 10 m/s;
therefore, an unusually strong upward current would be required to keep a
drop of this size from falling.

Various theories have been advanced to explain the growth of a cloud


element into a size that can precipitate. The two principal processes in the
formation of precipitation are: (1) the ice crystal process, and (2) the
coalescence process [29]. These two processes may operate together or
separately. The ice crystal process involves the presence of ice crystals in a
supercooled (cooled to below freezing) water cloud. Due to the fact that
saturation vapor pressure over water is greater than that over ice, there is a
vapor pressure gradient from water drops to ice crystals. This causes the ice
crystals to grow at the expense of the water drops and, under favorable
conditions, to attain precipitation size. The ice crystal process is operative
only in supercooled water clouds, and it is most effective at about -15oC.

The coalescence process is based on the difference in fall velocities and


consequent collisions to be expected between cloud elements of different
sizes (Fig. 2-2). The rate of growth of cloud elements by coalescence
depends upon the initial range of particle sizes, the size of the largest drops,
the drop concentration, and the sizes of the aggregated drops. The electric
field and drop charge may affect collision efficiencies and may therefore be
important factors in the release of precipitation from clouds [71]. Unlike the
ice crystal process, the coalescence process occurs at any temperature, lts
effectiveness varying from solid to liquid particles.
Fig. 2-2 The coalescence process (cmmap.org).

Forms of Precipitation. Precipitation occurs primarily in the form of drizzle,


rain, hail, or snow (Fig. 2-3). Drizzle consists of tiny liquid water droplets,
usually between 0.1 and 0.5 mm in diameter, falling at intensities rarely
exceeding 1 mm/h. Rain consists of liquid water drops, mostly larger than
0.5 mm in diameter. Rainfall refers to amounts of liquid precipitation. Rainfall
intensities can be classified as: light, up to 3 mm/h; moderate, from 3 to 10
mm/h; and heavy, over 10 mm/h. A rainstorm is a rainfall event lasting a
clearly defined duration.

Fig. 2-3 Forms of precipitation.

Hail is composed of solid ice stones or hailstones. Hailstones may be


spheroidal, conical, or irregular in shape and may range from about 5 to over
125 mm in diameter. A hailstorm is a precipitation event in the form of hail.

Snow is composed of ice crystals, primarily in complex hexagonal form and


often aggregated into snowflakes that may reach several millimeters in
diameter. Snowfall is precipitation in the form of snow. A snowstorm is a
snowfall event with a clearly defined duration. Snowpack is the volume of
snow accumulated on the ground after one or more snowstorms. Snowmelt,
or melt, is the volume of snow that has changed from solid to liquid state and
is available for runoff.
Factors affecting precipitation. Table 2-1 shows the various factors
affecting precipitation and their effect on: (a) moisture availability, (b)
condensation, and (c) coalescence. Factors No. 1 to 7 are entirely of natural
origin and, therefore, not subject to anthropogenic control. Factor No. 8 may
be subject to either natural or anthropogenic control. Factor No. 9 is the only
factor which is subject to anthropogenic control.

Table 2-1 Factors affecting precipitation.

Effect on
Description/
No. Factor
Moisture Conden- Coales- Example
availability sation cence

The climate
is tropical,
temperate, or
polar, a
1 Latitude ✓ - -
function of
the Hadley
and Ferrell
cells.

For example,
Global and the ENSO (El
2 mesoscale ocean ✓ ✓ - Niño
currents Southern
Oscillation).

For example,
monsoon-
related
precipitation,
3 Atmospheric currents ✓ ✓ -
such as that
of the Bay of
Bengal,
India.

Ocean or
large inland
lake;
Proximity to moisture presence of
4 ✓ - ✓
source salt particles
(aerosols)
from the
ocean.

Continental
location with
respect to
Relative continental one or more
5 ✓ ✓ -
position moisture
sources;
lifting through
horizontal
convergence;
frontal lifting.

Summer, fall,
winter,
spring;
6 Season - ✓ -
determines
radiation
balance.

Mountain
ranges,
which act as
barriers to
Presence of the
7 - ✓ -
orographic barriers movements
of continental
air masses;
orographic
lifting.

Through
volcano
eruptions or
fires, which
Natural/anthropogenic increase
8 atmospheric - - ✓ atmospheric
particulates particulates,
favoring the
downwind
formation of
precipitation.

Determined
by land-
surface
albedo,
which
Land surface conditions
condition (texture, the near
9 ✓ ✓ -
color, moisture surface-
content) radiation
balance and
makes
possible
thermal
lifting.

Quantitative Description of Rainfall

A rainfall event, or storm, describes a period of time having measurable and


significant rainfall, preceded and followed by periods with no measurable rainfall. The
time elapsed from start to end of the rainfall event is the rainfall duration. Typically,
rainfall duration is measured in hours. However, for very small catchments it may be
measured in minutes, while for very large catchments it may be measured in days.

Rainfall durations of 1, 2, 3, 6, 12, and 24 h are common in hydrologic analysis and


design. For small catchments, rainfall durations can be as short as 5 min. Conversely,
for large river basins, durations of 2 d and longer may be applicable [78]. Rainfall
depth is measured in mm, cm, or in., considered to be uniformly distributed over the
catchment area. For instance, a 60-mm, 6-h rainfall event produces 60 mm of depth
over a 6-h period.

Rainfall depth and duration tend to vary widely, depending on geographic location,
climate, microclimate, and time of the year. Other things being equal, larger rainfall
depths tend to occur more infrequently than smaller rainfall depths. For design
purposes, rainfall depth at a given location is related to the frequency of its
occurrence. For instance, 60 mm of rainfall lasting 6 h may occur on the average once
every 10 y at a certain location. However, 80 mm of rainfall lasting 6 h may occur on
the average once every 25 y at the same location.

Average rainfall intensity is the ratio of rainfall depth to rainfall duration. For example,
a rainfall event producing 60 mm in 6 h represents an average rainfall intensity of 10
mm/h. Rainfall intensity, however, varies widely in space and time, and local or
instantaneous values are likely to be very different from the spatial and temporal
average. Typically, rainfall intensities are in the range 0.1-30 mm/h, but can be as
large as 150 to 350 mm/h in extreme cases.

Rainfall frequency refers to the average time elapsed between occurrences of two
rainfall events of the same depth and duration. The actual elapsed time varies widely
and can therefore be interpreted only in a statistical sense. For instance, if at a certain
location a 100-mm rainfall event lasting 6 h occurs on the average once every 50 y,
the 100-mm, 6-h rainfall frequency for this location would be 1 in 50 years, 1/50, or
0.02.

The reciprocal of rainfall frequency is referred to as return period, or recurrence


interval. In the case of the previous example, the return period corresponding to a
frequency of 0.02 is 50 y.

Generally, larger rainfall depths tend to be associated with longer return periods. The
longer the return period, the longer the historical record needed to ascertain the
statistical properties of the distribution of annual maximum rainfall. Due to the paucity
of long rainfall records, extrapolations are usually necessary to estimate rainfall
depths associated with long return periods.
These extrapolations entail a certain measure of risk. When the risk involves human
life, the concepts of rainfall frequency and return period are no longer considered
adequate for design purposes. Instead, a reasonable maximization of the
meteorological factors associated with extreme precipitation is used, leading to the
concept of Probable Maximum Precipitation (PMP). For a given geographic location,
catchment area, event duration, and time of the year, the PMP is the theoretically
greatest depth of precipitation. In flood hydrology studies, the PMP is used as a basis
for the calculation of the Probable Maximum Flood (PMF).

For certain projects, a precipitation depth less than the PMP may be justified on
economic grounds. This leads to the concept of Standard Project Storm (SPS). The
SPS is taken as an appropriate percentage of the applicable PMP and is used to
calculate the Standard Project Flood (SPF) (Chapter 14).

Temporal and Spatial Variation of Precipitation

Temporal Rainfall Distribution. Rainfall intensities for events of short duration (1 h


or less) can usually be expressed as an average value, obtained by dividing rainfall
depth by rainfall duration. For longer events, instantaneous values of rainfall intensity
are likely to become more important, particularly for flood peak determinations.

The temporal rainfall distribution depicts the variation of rainfall depth within a storm
duration. It can be expressed in either discrete or continuous form. The discrete form
is referred to as a hyetograph, a histogram of rainfall depth (or rainfall intensity) with
time increments as abscissas and rainfall depth (or rainfall intensity) as ordinates, as
shown in Fig. 2-4 (a).

The continuous form is the temporal rainfall distribution, a function describing the rate
of rainfall accumulation with time. Rainfall duration (abscissas) and rainfall depth
(ordinates) can be expressed in percentage of total value, as shown in Fig. 2-4 (b).
The dimensionless temporal rainfall distribution is used to convert a storm depth into
a hyetograph, as shown in the following example.
Fig. 2-4 (b) A dimensionless temporal rainfall
Fig. 2-4 (a) A hyetograph.
distribution.

Example 2-1.

Using the dimensionless temporal rainfall distribution shown in Fig. 2-5, calculate a hyetograph for a 15-cm,
6-h storm.

For convenience, a time increment of 1 h, or 1/6 of the storm duration is chosen. The cumulative rainfall
percentages (at increments of 1/6 of storm duration) obtained from Fig. 2-5 are the following: 10, 20, 40, 70,
90, and 100% . Therefore, the incremental percentages, per hour, are: 10, 10, 20, 30, 20, and 10%. For a 15-
cm total storm depth, the incremental (hourly) rainfall depths are the following: 1.5, 1.5, 3.0, 4.5, 3.0, and 1.5
cm.
Fig. 2-5 Dimensionless temporal rainfall distribution for Example 2-1.

Spatial Rainfall Distribution. Rainfall varies not only temporally but also spatially,
i.e., the same amount of rain does not fall uniformly over the entire
catchment. Isohyets are used to depict the spatial variation of rainfall. An isohyet is a
contour line showing the loci of equal rainfall depth (Fig. 2-6 (a)).

Fig. 2-6 (a) Isohyets or isohyetal curves. Fig. 2-6 (b) A storm eye.
Individual storms may have a spatial distribution or pattern in the form of concentric
isohyets of approximately elliptic shape. This gives rise to the term storm eye to
depict the center of the storm (Fig. 2-6 (b)). In general; storm patterns are not static,
moving gradually in a direction approximately parallel to that of the prevailing winds.

Isohyets are also used to show spatial rainfall patterns for a given time period. Figure
2-7 shows an example of spatial rainfall distribution for the month of July 2008 in
Taranaki, New Zealand.

Fig. 2-7 Monthly spatial distribution of precipitation (mm).

For regional rainfall mapping, isohyets are commonly referred to


as isopluvials. Isopluvial maps for the United States are published by the
National Weather Service [58, 59, 85-88]. These maps show contours of
equal rainfall depth, applicable for a range of durations, frequencies, and
geographical locations; see, for example, Fig. 2-8 for San Diego County,
California, and Fig. 2-9 for the contiguous United States.
Fig. 2-8 100-yr 24-h isopluvials for San Diego County, California (0.1 in) (Source: NOAA) (Click -here- to
display).

Fig. 2-9 100-yr 24-h isopluvials for the contiguous United States (in.) (NOAA) (Click -here- to display).

For large catchments, highly intensive storms (thunderstorms) may cover


only a fraction of the whole basin, yet they may lead to severe flooding in
localized areas. The role of thunderstorms in determining the flood potential
of large basins is usually assessed on an individual basis.

Average Precipitation Over an Area. A precipitation (or rainfall) amount is


measured with rain gages. During a given storm, it is likely that the depth
measured by two or more rain gages of the same type will not be the same.
In hydrologic analysis, it is often necessary to determine a spatial average of
the rainfall depth over the catchment. This is accomplished by either of the
following methods:

1. Average rainfall,
2. Thiessen polygons, and
3. Isohyetal curves.

In the average rainfall method, the rainfall depths measured by the rain
gages located within the catchment are tabulated. These rainfall depths are
then averaged to find the average precipitation over the catchment, as shown
in Fig. 2-10 (a).

Fig. 2-10 (a) Average rainfall method (Click -here- to display).

In the Thiessen polygons method, the locations of the rain gages are plotted
on a scale map of the catchment and surrounding area. The locations
(stations) are joined with straight lines in order to form a pattern of triangles,
preferably with sides of approximately equal length. Perpendicular bisectors
to the sides of these triangles are drawn to enclose each station within a
polygon called a Thiessen polygon, circumscribing an area of influence, as
shown in Fig. 2-10 (b). The average precipitation over the catchment is
calculated by weighing each station's rainfall depth in proportion to its area
of influence.

Fig. 2-10 (b) Thiessen polygons method (Click -here- to display).

In the isohyetal method, the locations of the rain gages are plotted on a scale
map of the catchment and surrounding area. Each station's rainfall depth is
used to draw isohyets throughout the catchment in a manner similar to that
used in the preparation of topographic contour maps. The mid-distance
between two adjacent isohyets is used to delineate the area of influence of
each isohyet, as shown in Fig. 2-10 (c). The average precipitation over the
catchment is calculated by weighing each isohyetal increment in proportion
to its area of influence.

Fig. 2-10 (c) Isohyetal method (Click -here- to display).


The isohyetal method is regarded as more accurate than either the Thiessen
polygons or average rainfall methods. This is particularly the case when
averaging precipitation over catchments where orographic effects have a
significant influence on the local storm pattern. The Thiessen polygons
method is generally more accurate than the average rainfall method. The
increase in accuracy is likely to be more marked when averaging
precipitation over catchments with widely varying rainfall depths or large
differences in areas of influence.

Storm Analysis

Storm Depth and Duration. Storm depth and duration are directly related,
storm depth increasing with duration. An equation relating storm depth and
duration is:

h = c tn (2-2)

in which h = storm depth, in centimeters; t = storm duration, in hours; c = a


coefficient; and n = an exponent (a positive real number less than 1).
Typically, n varies between 0.2 and 0.5, indicating that storm depth
increases at a lesser rate than storm duration. By analyzing storm data on a
regional or local basis, Eq. 2-2 can be used to predict storm depth as a
function of storm duration. The applicability of such an equation, however, is
limited to the regional or local conditions for which it was derived.

Equation 2-2 can also be used to study the characteristics of extreme rainfall
events. A logarithmic plot of depth-duration data for the world's greatest
observed rainfall events (Table 2-2) results in the following enveloping line:

h = 39 t 0.5 (2-3)

in which h = rainfall depth, in centimeters, and t = rainfall duration, in hours.


The data of Table 2-2 are plotted in Fig. 2-11, including the enveloping line,
Eq. 2-3.

Table 2-2 World's greatest observed rainfall events [53].

Depth
Duration Location Date
(cm)

1 min 3.8 Barot, Guadeloupe 1970 November 26

8 min 12.6 Fussen, Bavaria 1920 May 25


15 min 19.8 Plumb Point, Jamaica 1916 May 12

42 min 30.5 Holt, Missouri 1947 June 22

2 h 10 min 48.3 Rockport, West Virginia 1889 July 18

2 h 45 min 55.9 D'Hanis, Texas (17 mi NNW) 1935 May 31

4 h 30 min 78.2 Smethport, Pennsylvania 1942 July 18

9h 108.7 Belouve, Reunion 1964 February 18

12 h 134.0 Belouve, Reunion 1964 February 28-29

18 h 30 min 168.9 Belouve, Reunion 1964 February 28-29

24 h 187.0 Cilaos, Reunion 1952 March 15-16

2d 250.0 Cilaos, Reunion 1952 March 15-17

3d 324.0 Cilaos, Reunion 1952 March 15-18

4d 372.1 Cherrapunji, India 1974 September 12-15

5d 385.4 Cilaos, Reunion 1952 March 13-18

6d 405.5 Cilaos, Reunion 1952 March 13-19

7d 411.0 Cilaos, Reunion 1952 March 12-19

15 d 479.8 Cherrapunji, India 1931 June 24-30

1 mo 930.0 Cherrapunji, India 1861 July

3 mo 1637.0 Cherrapunji, India 1861 May-July

6 mo 2245.0 Cherrapunji, India 1861 April-September

1 yr 2646.0 Cherrapunji, India 1860 August to 1861 July

2 yr 4077.0 Cherrapunji, India 1860-1861


Fig. 2-11 Depth-duration data for the world's greatest observed rainfall events.

Storm Intensity and Duration. Storm intensity and duration are inversely
related. From Eq. 2-2, an equation linking storm intensity and duration, can
be obtained by differentiating rainfall depth with respect to duration, to yield:

dh
______
= i = c n t n -1 (2-4)
dt

in which i = storm intensity. Simplifying,

a
______
i= (2-5)
tm

in which a = cn, and m = 1 - n. Since n is less than 1, it follows that m is also


less than 1.

Another intensity-duration model is the following:

a
_________
i= (2-6)
t+b
in which a and b are constants to be determined by regression analysis
(Chapter 7).

A general intensity-duration model combining the features of Eqs. 2-5 and 2-


6 is:

a
_____________
i= (2-7)
(t+b)m

For b = 0, Eq. 2-7 reduces to Eq. 2-5; for m = 1, Eq. 2-7 reduces to Eq. 2-6.

Intensity-Duration-Frequency. For small catchments it is often necessary


to determine several intensity-duration curves, each for a different frequency
or return period. A set of intensity-duration-frequency curves is referred to as
IDF curves, with duration plotted in the abscissas, intensity in the ordinates,
and frequency (or alternatively, return period) as curve parameter. Either
arithmetic (Fig. 2-12 (a)) or logarithmic (Fig. 2-12 (b)) scales are used in the
construction of IDF curves. Such curves are developed by government
agencies for use in urban storm-drainage design and other applications
(Chapter 4).

A formula for IDF curve can be obtained by assuming that the constant a in
Eqs. 2-5 through 2- 7 is related to return period T as follows:

a = k Tn (2-8)

in which k = a coefficient; T = return period; and n = an exponent (not related


to that of Eq. 2-2). This leads to

kTn
_____________
i= (2-9)
(t+b)m

The values of k, b, m, and n are evaluated from measured data or local


experience.
Fig. 2-12 (a) An arithmetic IDF curve
(Davenport, Iowa) [84].

Fig. 2-12 (b) A logarithmic IDF curve


(San Jose, California) [84].

Storm Depth and Catchment Area. Generally, the greater the catchment
area, the smaller the spatially averaged storm depth. This variation of storm
depth with catchment area has led to the concept of point depth, defined as
the storm depth associated with a given point area. A point area is the
smallest area below which the variation of storm depth with catchment area
can be assumed to be negligible. In the United States, the point area is
usually taken as 25 km2 (10 mi2).

The point depth applies for all areas less than the point area. For areas
greater than the point area, a reduction in point depth is necessary to account
for the decrease of storm depth with catchment area. This depth reduction is
accomplished with a depth-area reduction chart, a function relating
catchment area (abscissas) to point depth percentage (ordinates). Storm
duration is usually a curve parameter in a depth-area reduction chart.

Generalized depth-area reduction charts applicable to the contiguous United


States, for areas up to 1000 km2 (400 mi2) and durations from 30 min to 10 d
have been published by the National Weather Service (Figs. 2-13 (a) and
(b)). Regional and locally derived depth-area reduction charts may differ from
these generalized charts (see Section 14.1).
Fig. 2-13 (a) Generalized depth-area reduction Fig. 2-13 (b) Generalized depth-area reduction
chart chart
for 30-min to 24-h duration. for 1-d to 10-d duration.

Depth-Duration-Frequency. For midsize catchments, hydrologic analysis


shifts its focus to rainfall depth. Isopluvial maps depicting storm depths,
applicable for a range of durations, frequencies and catchment areas, are
available for the entire United State [58, 59, 85-88]. These maps show point
depth values and are therefore subject to depth-area reduction by the use of
an appropriate chart.

Depth-Area-Duration. Another way of describing the relation between


storm depth, duration and catchment area is the technique known as depth-
area-duration (DAD) analysis. This technique is basically an alternate way of
portraying the reduction of storm depth with area, with duration as a third
variable.

To construct a DAD chart, a storm having a single major center (storm eye)
is identified. Isohyetal maps showing maximum storm depths for each of
several typical durations (6-h, 12-h, 24-h, etc.) are prepared. For each map,
the isohyets are taken as boundaries circumscribing individual areas. For
each map and each individual area, a spatially averaged rainfall depth is
calculated by dividing the total rainfall volume by the individual area. This
procedure provides DAD data sets used to construct a chart showing depth
versus area, with duration as a curve parameter (Fig. 2-14).
Fig. 2-14 A depth-area duration curve.

DAD analysis can also be used to study regional rainfall characteristics.


Table 2-3 shows maximum DAD data for the United States, based on four
extreme events. The data confirm that storm depth increases with duration
and decreases with catchment area.

Table 2-3 Maximum depth-area-duration data for the United States [76].*

Duration (h)
Area
(km2)
6 12 18 24 36 48 72

25 62.7a 75.7b 92.2c 98.3c 106.2c 109.5c 114.8c

250 49.8b 66.8c 82.5c 89.4c 96.3c 98.8c 103.1c


500 45.5b 65.0c 79.8c 86.9c 93.2c 95.8c 99.6c

1250 39.1b 62.5c 75.4c 83.0c 88.9c 91.4c 94.7c

2500 34.0b 57.4c 69.6c 76.7c 83.6c 85.6c 88.6c

5000 28.4b 45.0c 57.1c 63.0c 69.3c 72.1c 75.4c

12500 20.6b 28.2b 35.8b 39.4c 47.5d 52.6d 62.0d

Storm: a. July 17-18, 1942, Smethport, Pennsylvania;b. September 8-10,


1921, Thrall, Texas; c. September 3-7, 1950, Yankeetown, Florida;d. June
27-July 1, 1899, Hearne, Texas.

* Average depth in cm; storm indicated by superscripted letter.

Probable Maximum Precipitation. For large projects, storm analysis using


depth-duration-frequency data is not sufficient to eliminate the likelihood of
failure. In such cases, the concept of PMP is used instead. In the United
States, PMP estimates are developed following guidelines included in the
HM NOAA Hydrometeorological Reports) serie [33-44] and related
publications [84-87]. These reports contain methodologies and maps for the
estimation of PMP for a given geographic location, range of durations and
catchment sizes, and time of the year (Chapter 14).

Geographic and Seasonal Variations of Precipitation

Precipitation varies not only temporally and spatially but also seasonally,
annually, and with geographic location and climate. Mean annual
precipitation, the total amount of precipitation that accumulates in one year,
on the average, at a given location, is used for classify climates (in terms of
precipitation) into eight classes [11]:

 Superarid: Less than 100 mm


 Hyperarid: 100 - 200 mm
 Arid: 200 - 400 mm
 Semiarid: 400 - 800 mm
 Subhumid: 800 - 1600 mm
 Humid: 1600 - 3200 mm
 Hyperhumid: 3200 - 6400 mm.
 Superhumid: More than 6400 mm.
The seasonality of precipitation is assessed with the precipitation seasonality
index: the ratio of accumulated precipitation for the three wettest consecutive
months to that for the three driest consecutive months, in an average year.
This index is used to classify climates into four classes [11]:

1. Nonseasonal: 1.0 - 1.6


2. Weakly seasonal: 1.6 - 2.5
3. Moderately seasonal: 2.5 - 10
4. Strongly seasonal: Greater than 10.

In general, arid and semiarid climates are associated with moderately


seasonal regimes; conversely, subhumid and humid climates are associated
with weakly seasonal or nonseasonal regimes. However, there are some
exceptions; for instance, the monsoon-type climates which prevail in some
parts of the world, which tend to be both humid and seasonal.

Precipitation Data Sources and Interpretation

Precipitation data are obtained by measurement using rain gages (Chapter


3). The National Climatic Data Center (NCDC), Asheville, North Carolina,
publishes precipitation data for about 8000 stations in the United States. A
large number of additional gages are operated by other federal, state, and
local agencies, and by individuals. U.S. federal agencies collecting
precipitation data on a regular basis include the National Weather Service
(NWS), the Army Corps of Engineers, the Natural Resources Conservation
Service (formerly Soil Conservation Service), the Forest Service, the Bureau
of Reclamation, and the Tennessee Valley Authority.

NCDC assembles precipitation data using hourly, daily, monthly, and yearly
intervals. Hourly precipitation data and maximum 15-minute duration
amounts are found in Hourly Precipitation Data. Daily and monthly
precipitation data are found in Climatological Data. Monthly and annual
precipitation data for about 250 major U.S. cities, including hourly rates, are
found in Local Climatological Data.

Regional precipitation-frequency atlases (U.S. Weather Bureau No. 40 [86],


NOAA Technical Memorandum NWS Hydro-35 [59] and Precipitation
Frequency Atlas of the Western United States [58]) are available. NCDC
Monthly and seasonal precipitation maps are found in Weekly Weather and
Crop Bulletin, available from NOAA/USDA Joint Agricultural Weather
Facility, in the USDA South Building, Washington, D.C. Additional sources
of precipitation data are given in Annotated Bibliography of NOAA
Publications of Hydrometeorological Interest, updated at regular intervals by
NWS, and in Selective Guide to Climatic Data Sources, updated at regular
intervals by NCDC.

Precipitation and other relevant climatological data are now accessible


online through NCDC's website at http://www.ncdc.noaa.gov. Clicking on
On-line Data Access provides access to a host of on-line services, including
U.S. monthly precipitation for NWS and Cooperative sites, and the On-line
Access and Service Information System (OASIS), which includes hourly and
15-minute precipitation data.

Filling In Missing Records. Incomplete records of rainfall are sometimes


possible due to operator error or equipment malfunction. In this case, it is
often necessary to estimate the missing record. Assume that a certain station
X has a missing record. A procedure to fill in the missing record is to identify
three index stations (A, B, and C) having complete records, located as close
to and as evenly spaced around station X as possible. The mean annual
rainfall for each of the stations X, A, B, and C is evaluated. If the mean annual
rainfall at each of the index stations A, B, or C is within 10 percent of that of
station X, a simple arithmetic average of the rainfall values at the index
stations provides the missing value at station X.

If the mean annual rainfall at any of the index stations differs by more than
10% from that of station X, the normal ratio method is used [55]. In this
method, the missing precipitation value at station X is the following:

NX NX NX
PX = (1/3) [ _____ PA + _____
PB + _____
PC ] (2-10)
NA NB NC

in which P = precipitation, N = mean annual rainfall, and the


subscripts X, A, B, and C refer to the respective stations.

An alternate method for filling in missing precipitation data has been


developed by the National Weather Service [49]. The method requires data
for four index stations A, B, C, and D, each located closest to the station X
of interest, and in each of four quadrants delimited by north-south and east-
west lines drawn through station X (Fig. 2-15). The estimated precipitation
value at station X is the weighted average of the values at the four index
stations. For each index station, the applicable weight is the reciprocal of the
square of its distance L to station X.
Fig. 2-15 Position of station X and index stations A, B, C, and D.

The procedure is described by the following formula:

Σ ( Pi / Li 2
)
i=1
PX = _____________________
(2-11)
4

Σ ( 1 / Li 2
)
i=1

in which P = precipitation; L = distance between index stations and station X;


and i refers to each one of the index stations A, B, C, and D.

Double-mass Analysis. Changes in the location or exposure of a rain gage


may have a significant effect on the amount of precipitation it measures,
leading to inconsistent data (data of different nature within the same record).
The consistency of a rainfall record is tested with double-mass analysis. This
method compares the cumulative annual (or, alternatively, seasonal) values
of station Y with those of a reference station X. The reference station is
usually the mean of several neighboring stations. The cumulative pairs
(double-mass values) are plotted in an x-y arithmetic coordinate system, and
the plot is examined for trend changes. If the plot is essentially linear, the
record at station Y is consistent. If the plot shows a break in slope, the record
at station Y is inconsistent and should be corrected. The correction is
performed by adjusting the records prior to the break to reflect the new state
(after the break). To accomplish this, the rainfall records prior to the break
are multiplied by the ratio of slopes after and before the break (Fig. 2-16).

Fig. 2-16 Double-mass analysis.


2.2 PRECIPITATION MEASUREMENTS

Introduction

Engineering hydrology is based on analysis and measurements.


Measurements are necessary in order to complement and verify the
analysis. Hydrologic measurements are usually performed in the field, using
equipment and techniques specifically designed to measure a variable
characterizing a certain phase of the hydrologic cycle. For instance, rainfall
is measured with raingages, evaporation is measured with evaporation pans,
and streamflow is measured using streamgaging techniques.

Measurements are closely related to hydrologic analysis. In some cases they


are an integral part of it; in others, they serve to support it. For instance,
statistical hydrology is not possible without measurements. In flood
frequency analysis, a historical flow record is necessary in order to define
the properties of the predictive equations. With parametric models,
measurements aid in parameter estimation, increasing model reliability.
Deterministic and conceptual models also benefit from hydrologic
measurements.

Precipitation

Precipitation is measured with raingages. A raingage is an instrument that


captures precipitation and measures its accumulated volume during a certain
time period. The precipitation depth for the given period is equal to the
accumulated volume divided by the collection area of the gage. The average
precipitation intensity is equal to the precipitation depth divided by the length
of the period.

Any receptacle that has vertical sides and is open to the air is a defacto
raingage and can provide valuable information on accumulated rainfall
during a storm. Two such measurements, however, are not directly
comparable unless the receptacles are of the same size and shape and
similarly exposed. To increase the utility of the measurements, it is
necessary to use standard equipment and procedures.

Raingages can be of two types:

1. Nonrecording, or
2. Recording.

A nonrecording raingage measures the total rainfall depth accumulated


during one time period, usually 1 day. In the United States, the standard
nonrecording rain gage used by the National Weather Service has a funnel-
shaped collector element or receiver of 8-in. top diameter located inside an
overflow can (Fig. 2-1). Rain is caught by the collector and funneled into a
measuring tube. The cross-sectional area of the measuring tube is one-tenth
that of the collector. Therefore, rainfall depths are amplified ten times as they
pass from the collector into the measuring tube, increasing the accuracy of
the measurement.

Fig. 2-1 Nonrecording raingage

A recording raingage records the time it takes for rainfall depth


accumulation. Therefore, it provides not only a measure of rainfall depth but
also of rainfall intensity. The slope of the curve showing accumulated rainfall
depth versus time is a measure of the instantaneous rainfall intensity.
Recording rain gages rely on one of the following devices:

1. A tipping bucket,
2. A weighing mechanism, or
3. A float chamber.

The tipping-bucket gage features a two-compartment receptacle (i.e., the


bucket) pivoted on a knife edge. The device is calibrated so that when one
of the compartments is full (with a fixed amount of rain) and the other is
empty, the bucket overbalances and tips. At the start, rain is funneled into
one of the compartments, which is positioned for filling. As rainfall continues
to fill this first compartment, the second remains empty. When the first
compartment is full, the bucket tips, emptying its contents into a reservoir
and at the same time placing the second compartment in filling position. The
tipping closes an electric circuit, which drives a pen that records on a strip
chart affixed to a clock-driven revolving drum. Thus, each electrical contact
representing a specific amount of rain is recorded. The alternate filling and
emptying of the two compartments continues until rainfall ceases.

The tipping-bucket gage has a few disadvantages. During periods of intense


rainfall, some of the rain may not be measured while the bucket is tipping. In
addition, the record consists of a series of steps rather than being a smooth
curve, and the gage is not suitable for measuring snow. Nevertheless,
tipping-bucket gages are durable, simple to operate, and of good overall
reliability.

A weighing gage has a device that weighs the rain or snow collected in a
bucket. As it fills with precipitation, the bucket moves downward and its
movement is transmitted to a pen on a strip-chart recorder. This type of gage
is useful in cold climates where it is necessary to record both rainfall and
snowfall. However, weighing gages have some disadvantages. Among them
are wind action on the bucket, which produces erratic traces on the recording
chart, and the overall lack of sensitivity of the measurement.

Float gages are essentially water-level gages. A float located inside a


chamber is connected to a pen on a strip-chart recorder. The float rises as
the collected rainwater enters the chamber, and the rise of the float is
recorded on the chart. Some float gages are limited to the capacity of the
chamber. Others are equipped with a self-starting siphoning device that
empties the chamber when it becomes full and returns the pen to the zero
position on the strip chart. The use of float gages is limited to nonfreezing
ambient temperatures, although heaters and other similar devices have been
used in an attempt to overcome the problem of freezing. Oil and mercury,
which have freezing temperatures below that of water, have also been used
inside the chamber. The siphoning action of the float gage can cause serious
losses of rain during severe storms.

Errors in Measuring Precipitation Data

The water collected by a raingage is only a small sample of the precipitation


that has fallen in a certain area. Whether this sample is representative of the
average precipitation over the area remains to be determined by further
analysis.
A series of rain gages located within a drainage area constitutes a raingage
network. The density of the network is the number of rain gages per square
kilometer (or square mile). The error of rainfall measurements can be
investigated by studying the spatial averages computed from networks of
different densities [1, 13, 23]. In general, sampling errors increase with an
increase in rainfall depth. Conversely, sampling errors decrease with an
increase in network density, storm duration, and catchment area.

An important question in engineering hydrology is whether errors in


precipitation measurement can serve to compound the errors inherent in the
use of rainfall-runoff simulation models. The answer to this question is
elusive. Limited data by Johanson [11] indicates that the error variability in
precipitation measurements is likely to be less than the error variability in
model calibration. Calibration is the process by which model parameters are
adjusted to match measured and simulated flows.

Precipitation Measurements Using Telemetry

Self-reporting rain gages (or rainfall sensors) have automatic data transmittal
capabilities. These rain gages use automatic radio transmitters (telemeters)
to broadcast rainfall measurements from a remote station to a central station
in real time, i.e., during the storm event (Fig. 2-2). The advantage of a
telemetric station is that it shortens the time that would otherwise be required
to gather rainfall data. In certain cases, especially when speed of processing
is of utmost importance, a network of rainfall sensors linked by telemetry may
be the only practical means of collecting rainfall data. Applications of
telemetric rainfall sensors are usually found in connection with operational
hydrology and real-time flood forecasting.
Fig. 2-2 A telemetric weather station.

The link between the remote station and the central station is usually
established by radio, telephone, or a combination of both. Where radio
frequencies are scarce, telephone lines can be used to transmit the data.
The remote rainfall station is interfaced to a telephone line either through a
modem or an acoustic coupler. The latter enables the remote station to be
called from any telephone, whether it is equipped with a modem or not.

Radio transmission may take the form of a very high frequency (VHF) or
ultrahigh frequency (UHF) link for short distances, or high frequency (HF) for
very long distances. VHF and UHF frequencies behave in a manner similar
to light and, therefore, cannot travel far beyond the horizon. The best
reception is obtained when the transmitting and receiving antennas are
within line of sight of each other. With the use of high masts, a span of 40
km or more can be achieved. Transmitter power ranges from 5 W for short
distances to 25 W for longer distances.
Very long distance transmissions require repeating stations at 30- to 60-km
intervals, but these are expensive and difficult to maintain. An alternative is
to use HF radio, by which great distances can be spanned through a series
of reflections between ionosphere and ground. Normal transmissions have
a span of a few hundred kilometers. These HF links, however, are subject to
variations in signal strength and are susceptible to interference with other
transmitters, making their use more difficult than VHF and UHF.

Another way to transmit data through radio waves is by the use of satellites.
A remote station can transmit data to a satellite for retransmirtal to a ground
receiving station. The radio link operates in the UHF range and requires only
a few watts of power. To transmit data via a satellite, the station is linked to
a commercially made data collection platform. This device stores the day's
data in its solid state memory for transmission every 24 h, although hourly
transmissions are also possible.

Precipitation Measurements Using Radar. Weather radar systems are a


potentially powerful tool for measuring the temporal and spatial variability of
rainstorms. A radar system operates by emitting a regular succession of
pulses of electromagnetic radiation from its antenna. The pulses are on the
order of 1 μs, and the system emits approximately 1000 of these pulses
every second. Between pulses, the system's antenna becomes a receiver of
the energy of the emitted pulses scattered by various targets. These returned
signals are transformed into a visual display on the radar scope.

For spherical objects (e.g., raindrops), the power received can be expressed
as follows:

KΣnD6
__________
P= (3-1)
λ 4R 2

in which P = power received, n = number of drops, D = diameter of drops, λ


= wavelength of the radiation, R = distance (range) from the radar, and K =
a factor that depends on the power of the transmitted signal, antenna size
and shape, and properties of the scattering particles.

Weather radars have wavelengths in the range of 3 to 10 cm. Following Eq.


3-1, a 3-cm radar returns about 120 times as much power as that returned
by a 10-cm radar. Hence a 3-cm radar can detect weak targets such as the
small droplets associated with very light rain, whereas a 10-cm radar can be
used to sense much heavier rains.
Attenuation, caused by absorption and scattering by clouds and
precipitation, can also affect the performance of the radar. Attenuation is a
function of radar wavelength, being greater for shorter wavelengths.

The reduction in power received P with distance R (range) is a constant for


a given system and can be adjusted to make distant targets show the same
brightness as closer targets of similar character. Since power P is
proportional to the sixth power of the drop diameter D, radars in the 3- to 10-
cm wavelength size can easily sense rainsize droplets and not sense other
size particles at all. A 10-cm radar is used for detecting highly intensive
storms, which are likely to produce extreme floods. For light rains or snow
sensing, a shorter wavelength radar is preferable.

The radar reflectivity (ΣnD 6) can be empirically related to precipitation


intensity as follows:

Z= AIB (3-1)

in which Z = radar reflectivity; I = precipitation intensity; and A and B are


empirical constants. The values of A and B depend upon the type of
precipitation being observed. Many values have been specified; those most
often used are A = 200 and B = 1.6 [4].

Several types of error are possible when using radar to sense precipitation.
For instance, the radar beam can overshoot shallow precipitation at long
ranges, missing the target. Another source of error is the presence of low
level evaporation beneath the radar beam, as well as several other
meteorological factors [7]. Uncertainties in radar sensing of precipitation can
be resolved by calibrating the system with a raingage. This is usually
accomplished by fixing exponent A (in Eq. 3-2) at a certain value (for
instance B = 1.6), and using raingage data to derive a value of coefficient A.

Vous aimerez peut-être aussi