Sie sind auf Seite 1von 26

30

Contents Part A: Irrigation and Drainage


30.1 Irrigation fundamental concepts 30.1.1 Introduction 30.1.2 Soil moisture 30.1.3 Crop water requirements 30.1.4 Irrigation efficiency 30.1.5 Effective rainfall 30.1.6 Salinity and leaching requirement Irrigation methods 30.2.1 Introduction 30.2.2 Surface irrigation 30.2.3 Sprinkler irrigation 30.2.4 Trickle irrigation 30.2.5 Sub-irrigation 30.2.6 Irrigation canal design Drainage of agricultural land 30.3.1 Introduction 30.3.2 Sub-surface drainage of irrigated land 30.3.3 Drainable surplus 30.3.4 Drainage of lands subject to excess rainfall 30.3.5 Drain spacing 30.3.6 Drain flow 30.3.7 Drainage layouts 30.3.8 Drainage of heavy soils 30.3.9 Bedding systems 30.3.10 Surface drainage for irrigated land 30/3 30/3 30/3 30/3 30/4 30/5 30/5 30/6 30/6 30/6 30/7 30/9 30/9 30/9 30/9 30/9 30/10 30/10 30/10 30/11 30/11 30/11 30/12 30/12 30/12

Irrigation, Drainage and River Engineering


W Pemberton BSc, FICE
Head of Irrigation and Drainage Department Sir Murdoch MacDonald and Partners Head of River Engineering Department Sir Murdoch MacDonald and Partners
30.4.2 Rivers as natural drains 30.4.3 Economic issues 30.5 Hydrology 30.5.1 Introduction 30.5.2 Measurement 30.5.3 Statistics 30.5.4 Flood flow calculation methods 30.5.5 Hydrographs 30.5.6 Curve number method 30.5.7 The Flood studies report Channel regime 30.6.1 Regime flow 30.6.2 Regime formulae 30.6.3 Practical applications Sediment transport 30.7.1 Basic concepts 30.7.2 Sediment transport estimates 30.7.3 Sediment transport equations 30.7.4 Stable channel design Channel design 30.8.1 Channel flow formulae 30.8.2 Channel stability 30.8.3 Other considerations Channel improvements 30.9.1 Channel clearance 30.9.2 Realignment 30.9.3 Revetments and lining

C E Rickard BSc, CEng, MICE, MIWEM

30/12 30/13 30/13 30/13 30/13 30/13 30/13 30/14 30/14 30/14 30/14 30/14 30/14 30/15 30/15 30/15 30/15 30/15 30/16 30/16 30/16 30/16 30/17 30/17 30/17 30/17 30/17 30/19 30/19 30/20

30.2

30.6

30.7

30.3

30.8

30.9

Part B: Land Drainage and River Engineering


30.4 Land drainage and flood alleviation 30.4.1 Objectives of land drainage 30/12 30/12

30.10 Embankments 30.10.1 Introduction 30.10.2 Design

This page has been reformatted by Knovel to provide easier navigation.

30.10.3 Stability 30.10.4 Construction 30.10.5 Rood walls 30.11 Detention basins, washlands and catchwater drains 30.11.1 Detention basins 30.11.2 Washlands 30.11.3 Catchwater drains 30.12 Structures 30.12.1 Introduction 30.12.2 Retaining walls 30.12.3 Bridges 30.12.4 Weirs

30/20 30/20 30/21 30/21 30/21 30/21 30/21 30/22 30/22 30/22 30/22 30/22

30.12.5 Gated control structures 30.12.6 Tidal outfalls 30.13 Pumping 30.13.1 Single or multiple pumps 30.13.2 Motive power 30.13.3 Pumps 30.13.4 Control 30.13.5 Pump station building 30.13.6 Other types of pumping installation References

30/23 30/24 30/24 30/24 30/24 30/24 30/25 30/25 30/26 30/26

This page has been reformatted by Knovel to provide easier navigation.

PART A: IRRIGATION AND DRAINAGE 30.1 Irrigation - fundamental concepts

Table 30.1
Moisture content (percentage by weight) Soil type Permanent wilting point 4 8 18 30 Available moisture 4 7 10 15

30.1.1 Introduction Irrigation is desirable where natural rainfall does not meet the plant water requirements for all or part of the year. Irrigation is essential for agriculture in the desert but even in areas such as northern Europe it can improve the yield of crops normally grown under rainfall conditions only. 30.1.2 Soil moisture The soil can be considered a moisture reservoir. Soils can be classified under the International Soil Science Association (ISSA) system as follows: Fraction Coarse sand Fine sand Silt Clay Particle size (mm) 2-0.2 0.2-0.02 0.02-0.002 < 0.002 Field capacity Coarse sand Fine sand Silt Clay 8 15 28 45

With knowledge of the crop rooting depth, the available soil moisture and the crop water requirements, it is possible to select a suitable irrigation interval (time between irrigations). Not all water in the root zone is readily available to the crop. It is normal to allow the crop to deplete only 50% of the available moisture before irrigating again. More detailed guidelines are given by the Food and Agricultural Organization.1 30.1.3 Crop water requirements Crop water requirements are defined as the depth of water required to meet the water loss through evapotranspiration CETcrop) of a crop. The effect of climate on crop water requirements is given by the reference crop evapotranspiration CET0) which is defined as the rate of evapotranspiration from an extensive surface of green grass of uniform height (8 to 15cm): ET^-k^ET, (30.1)

Water is held by the soil in the soil pores. The amount of water held can be defined as follows: (1) Saturation: the state of complete soil wetness when no further water may be added to the soil. (2) Field capacity (FC): the condition reached after water has drained from the soil by gravity. (3) Permanent wilting point (PWP): the condition reached after plants have extracted all the moisture they can from the soil. (4) Available water: defined as (FC-PWP), the amount of water held by the soil that plants can use. Plants respond to how tightly the water is held by the soil which is defined as soil moisture tension. Generally, it is assumed that the soil moisture tension at field capacity is 0.3 bar pressure. Soil moisture tension at PWP is assumed to be 15 bar. Typical moisture contents for various soils are shown in Table 30.1.

where kc is the crop coefficient which varies with crop, growth stage, growing period and prevailing weather conditions The most reliable method of estimating T0 is generally considered to be the PENMAN method. This method is best described by Doorenbos and Pruitt1 which also gives details of crop coefficients for a wide range of crops. Values of Tcrop are normally calculated for 10-day periods. A typical crop coefficient curve is shown in Figure 30.1. A simpler method was proposed by Blaney and Criddle2 in

Planting date

Approx. 10% ground cover

Crop coefficient K

Crop Initial development

70-80% ground cover

Mid-season

Late

Figure 30.1 Example of crop coefficient curve. (After J. Doorenbos and W. O. Pruitt (1977) Crop water requirements. Food and Agriculture Organization Irrigation and Drainage Paper No. 24.)

Maturity Harvest

Table 30.2 Monthly percentage of annual daytime hours (p) for different latitudes
Latitude North

South
40 42 44 46 48 50 52 54 56 58 60

Jan. JuL 6.76 6.63 6.49 6.34 6.17 5.98 5.77 5.55 5.30 5.01 4.67

Feb. Aug. 6.72 6.65 6.58 6.50 6.41 6.30 6.19 6.08 5.95 5.81 5.65

Mar. Sep. 8.33 8.31 8.30 8.29 8.27 8.24 8.21 8.18 8.15 8.12 8.08

Apr. Oct. 8.95 9.00 9.06 9.12 9.18 9.24 9.29 9.36 9.45 9.55 9.65

May Nov. 10.02 10.14 10.26 10.39 10.53 10.68 10.85 11.03 11.22 11.46 11.74

Jun. Dec.
10.08 10.22 10.38 10.54 10.71 10.91 11.13 11.38 11.67 12.00 12.39

JuL Jan. 10.22 10.35 10.49 10.64 10.80 10.99 11.20 11.43 11.69 11.98 12.31

Aug. Feb. 9.54 9.62 9.70 9.79 9.89


10.00 10.12 10.26 10.40 10.55 10.70

Sep. Mar. 8.29 8.40 8.41 8.42 8.44 8.46 8.49 8.51 8.53 8.55 8.57

Oct. Apr. 7.75 7.69 7.63 7.57 7.51 7.45 7.39 7.30 7.21 7.10 6.98

Nov. May 6.72 6.62 6.49 6.36 6.23 6.10 5.93 5.74 5.54 5.04 4.31

Dec. Jun. 7.52 6.37 6.21 6.04 5.86 5.65 5.43 5.18 4.89 4.56 4.22

Note: Southern latitudes apply 6-month difference as shown.

which the monthly crop water requirements Tcrop (in millimetres) are found by multiplying the mean monthly temperature Tm (0C) by the monthly percentage of annual daytime hours p and a monthly crop coefficient k\
ETcrop = (OA6Tm + Z)kp

divided into three parts: (1) field application; (2) field canal; and (3) distribution efficiency. 30.1.4.1 Field application efficiency (Ea)

Table 30.2 shows the monthly percentage of p for different latitudes. A sample calculation of water requirements for maize planted mid May near Saskatoon (latitude 520N) is shown in Table 30.3. A simple calculation of gross irrigation requirements (/gross) can be made as follows:

Ea is dependent on soil type and type of irrigation system used. Typical values are given in Table 30.4.

Table 30.4
Irrigation method Application practices a, % water application efficiency Soil texture heavy light Sprinkler Daytime application, moderately strong wind Night application Poorly levelled and shaped - Well levelled and shaped Poorly graded and sized Well graded and sized

i^-(fT^,-W EI

(30.2)

where Tcrop is the crop water requirements, /?e is the effective rainfall and Ea is the field application efficiency

Table 30.3
Mean monthly temp. 0 Water % annual Crop requiredaytime coefficient ments (mm) (V (P)

Trickle Basin

60 70 80 60 75 55 65

60 70 80 45 60 40 50

Period Days

(Q

Furrow Border

30.1.4.2 Field canal efficiency 15/5-31/5 1/6- 3/6 4/6-30/6 1/7- 8/7 9/7-31/7 1/8-17/8 18/8-31/8 1/9-16/9 17 3 27 8 23 17 14 16

(Ef)

11.6 19.7 19.3 21.3


8.7

5.95 1.11 10.02 2.89 8.31 5.55 4.57 4.53

0.35 0.35 0.96 1.05 1.14 1.14 1.02 0.75

27.8 6.6 164.1 51.2 159.9 112.6 83.0 40.8 646.0

Ef is dependent on type of field channel used and area served. Blocks larger than 20 ha Blocks up to 20 ha Unlined canals Lined or piped Unlined canals Lined or piped 0.90 0.95 0.80 0.90

30.1.4 Irrigation efficiency It is necessary to account for losses of water incurred during conveyance and application to the field. Efficiencies can be

30.1.4.3 Distribution efficiency (Ed) Distribution efficiency (Ed) is dependent on area served, the level of water management, and canal seepage which is the main component. Canal seepage can be calculated separately from seepage rates given in Table 30.11 (page 30/9). Typical overall values for EA are given below:

Table 30.5 Average monthly effective rainfall as related to average monthly E7"crop and mean monthly rainfall
Monthly mean rainfall (mm) Average 25 monthly 50 ETCTOP 75 (mm) 100 125 150 175 200 225 250

12.5 25
8 8 9 9 10 10 11 11 12 13 16 17 18 19 20 21 23 24 25 25

37.5 50
24 25 27 28 30 31 32 33 35 38 32 34 35 37 39 42 44 47 50

62.5 75
39 41 43 46 49 52 54 57 61 46 48 52 54 57 61 64 68 72

87.5 100

112.5 125

137.5 150

162.5 175

187.5 200

56 59 62 66 69 73 78 84

62 66 70 74 78 82 87 92

69 73 76 81 86 91 96 102

80 85 89 95 100 106 112

87 92 97 103 109 115 121

94 98 104 111 117 124 132

100 107 112 118 125 132 140

116 119 126 134 141 150

120 127 134 142 150 158

133 141 150 159 167

Where net depth of water that can be stored in the soil at time of irrigation is greater or smaller than 75 mm, the correction factor to be used is: Effective storage Correction factor 20 0.73 25 0.77 37.5 0.86 50 0.93 62.5 0.97 75 1.00 100 1.02 125 1.04 150 1.06 175 1.07 200 1.08

Continuous supply with no substantial change in flow Rotational supply in projects of 3000 to 7000 ha and rotation areas of 70 to 300 ha, with effective management Rotational supply in large schemes (> 10 000 ha) and small schemes (< 1000 ha) with respective problematic communication and less effective management: based on predetermined schedule based on advance request

0.90

salinity level will increase to make it unfit for plant growth. The process of dissolving and transporting soluble salts downwards to below the root zone is known as leaching. The maximum leaching requirements can be calculated from: Leaching requirement (LR) for surface or sprinkler irrigation ,EC9 LR ~5ECe-ECw For drip and high frequency sprinklers (almost daily)
EC

0.80

ro-

2MaxCe

CW = electrical conductivity of irrigation water, mmho/cm 0.70 0.65 ECe = electrical conductivity of the soil saturation extract for a given crop to the tolerable degree of yield reduction (see Table 30.6) Max ECC = maximum tolerable electrical conductivity of the soil saturation extract for a given crop (see Table 30.7) Alkalinity and toxicity may also affect soil permeability and crop growth. For further details, Ayes and Westcott4 can be consulted. Salinity hazard has been classified by the US Department of Agriculture (USDA) as shown in Table 30.6.

The total irrigation requirements at the head of the system (7sys) can be calculated from:
/s =

^ ^ff

(30.3)

30.1.5 Effective rainfall (Re) All rainfall is not effective in providing water for crop use. The calculation of effective rainfall is discussed in detail by Dastane.3 A simple method has been developed by the US Soil Conservation Service which relates effective rainfall with rcrop and mean monthly rainfall (see Table 30.5). For example, with a monthly rainfall of 50 mm, an ETCTOp of 100mm and an effective soil storage of 100mm, the correction factor is 1.02 and the effective rainfall is 1.02 x 35 = 36 mm. 30.1.6 Salinity and leaching requirement All irrigation water contains some dissolved salts. If no effort is made to move salts through and beyond the root zone, the soil

Table 30.6
Salinity of water (mmho/cm) Salinity hazard <0.25
Low

0.25 to 0.75 0.75 to 2.25


>2.25

Medium High Very high

Suitable for most crops and soils Suitable for moderately salt tolerant crops Not suitable for low permeability soils Generally not suitable for irrigation

In many instances, the usual inefficiencies of water application satisfy the leading requirements, but it is sometimes necessary to allow additional irrigation water for leaching. The leaching efficiency varies with soil type and may vary from 100% for sandy soils to perhaps as low as 30% for swelling heavy clay soils.

Siphon pipe Field channel Figure 30.2 Basin irrigation Basin

Bund Drain

30.2 Irrigation methods


30.2.1 Introduction The choice of method of irrigation is dependent on technical feasibility and economics. Normal methods fall into four main categories: (1) surface; (2) sprinkler; (3) trickle; and (4) subirrigation. 30.2.2 Surface irrigation Surface irrigation is still the most common method of irrigation employed and is suitable for the irrigation of most soils with an infiltration rate of less than 150 mm/h and for lands with a flat topography with an overall slope of less than 3%, although these limitations are exceeded in some situations. There are four main types of surface irrigation: (1) basin; (2) border strip; (3) furrow; and (4) corrugation irrigation. 30.2.2.1 Basin irrigation Water is applied from a small canal by gravity to fill a level basin surrounded by earth bunds. In practice, these basins are often small but the most efficient irrigation is obtained by using large basins, at least a hectare in area and accurately levelled. Water should be applied to those basins at a rate of at least 2 to 4 times the infiltration rate of the soil. Basin irrigation is most suitable for very flat and level land and soils with low infiltration rates. When adopted for uneven topography the basin size must be kept small in order to limit the quantity of land levelling required. Land levelling rates in excess of 1000m3/ha should be avoided. The cultivation of paddy rice is normally done using basin irrigation. Basin irrigation is illustrated in Figure 30.2. 30.2.2.2 Border strip irrigation The land is divided into strips separated by earth bunds which run generally down the slope, and water is applied at the head of the strip and allowed to flow down the slope infiltrating the soil as it flows across it (see Figure 30.3). The strip is graded at an even slope along its length in the direction of flow and level across the strip. Efficient irrigation is obtained by choosing the strip width, length and discharge to meet the soil infiltration rate

Field channel Siphons Border Strip Irrigation Furrow irrigation Figure 30.3 Surface irrigation methods
and land slope conditions to give as constant a depth of water as possible infiltrated over the length of the strip. Typical border strip designs from the USDA Yearbook are given in Table 30.8. Border strips are suitable for land with a more pronounced existing slope, thus reducing the amount of land levelling necessary. (For more information on design of sprinklers, see work by Barrs.6) 30.2.2.3 Furrow irrigation Furrow irrigation is used for the irrigation of row crops or crops grown on beds between furrows. Furrow irrigation usually implies sloping land although horizontal furrows can be used for row crops within level basins. Water is applied to the upper end of each furrow and flows down the furrow with water infiltrating into the beds between the furrows on which the crop is grown. Furrow spacings are a function of crop and type of tillage machinery used. Typically furrows are spaced 0.75 to 1.05m apart. Table 30.9 gives recommended maximum furrow lengths in metres for various soil types, furrow slopes and average depth of water applied over the whole field. Furrow slopes should be checked for erodability. The maximum non-erosive flow in furrows (Qn) can be estimated from: Qm = 0.60/5 (30.3)

Surface drain

Table 30.7 Crop salt tolerance for selected crops


Yield potential 100% (ECj (ECJ
8.0 6.0 1.0 4.6 1.7 4.0 5.3 4.0 0.7 3.1 1.1 2.7 90% (ECj 75% (ECj 50% (ECj

Crop Barley Wheat Typical vegetables (beans, carrots, lettuce, onions) Forage, grasses Fruit trees Date palm

(ECJ
6.7 4.9 1.1 3.9 1.6 4.5

(ECJ
8.7 6.4 1.9 5.3 2.2 7.3

(ECJ 12.0 8.7


3.1 7.4 3.2 12.0

Max EC, 28 20 8 18 8 32

10.0 7.4
1.7 5.9 2.3 6.8

13.0 9.5
2.8 7.9 3.3 10.9

18.0 13.0
4.6 11.1 4.8 17.9

Table 30.8 Typical border strip designs


Slope Soil type ((%)
0.25

Depth applied (mm)


50 100 150 50 100 150

Strip width (m)


15 15 15 12 12 12

Strip length (m)


150 250 400 100 150 250

Flow (1/s)
240 210 180 80 70

and is used for close-growing crops such as wheat. The corrugations are some 10cm deep and spaced 40 to 75cm apart. Because the corrugation flows are small, slopes up to 8% have been used. This method of irrigation is not widely used outside the US. For more details of surface irrigation methods, Booher5 can be consulted.

1.00

30.2.3 Sprinkler irrigation 30.2.3.1 Types of sprinkler The application of water by overhead sprinklers takes many forms which include the following. (1) Permanent and solid set - a network of pipes and sprinklers which covers the whole area to be irrigated. No movement of equipment within a season is necessary. This is the most expensive form of sprinkler irrigation. (2) Lateral move sprinklers - sprinklers on a lateral line that is moved by hand after each irrigation application to the next area to be irrigated. This is the most widely used system. (3) Traveller systems - these are motorized methods of moving sprinklers and include: (a) sideroll - lateral pipe and sprinklers on wheels, pushed by hand or small motor from one position to next irrigation position; (b) mobile rain gun - single gun winched across field whilst irrigating and fed from a hose reel; (c) centre pivot - overhead lateral with sprinklers which rotates about centre whilst irrigating; (d) linear move - similar to centre pivot but moves laterally across the field. The most common system used in developing countries, where labour is inexpensive, is the lateral move sprinkler system. In developed countries where labour is expensive, various forms of travellers are more common. Sideroll is suitable for low crops as the lateral is not normally more than 1.8 m above the ground. Rain guns with their high water-pressure requirements and, hence, high energy costs are best used for supplementary irrigation. Centre pivot and linear move are becoming the most popular traveller systems in arid areas. Permanent and solid sets are very expensive and hence used only on high-value crops. 30.2.3.2 Sprinkler design Individual sprinklers provide a cone of precipitation so that

Coarse

70 35 30 30

2.00

50 100 150 50 100 150 50 100 150 50 100 150 50 100 150 50 100 150 50 100 150

10 10 10 15 15 15 12 12 12 10 10 10 15 15 15 12 12 12 10 10 10

60 100 200

0.25

1.00
Medium

250 400 400 150 300 400 100 200 300 400 400 400 400 400 400
320 400 400

210 180 100 70 70 70


30 30 30

2.00

0.25

120 70 40
70 35 20 30 30 20

1.00
Fine
2.00

where Qm is in litres per second and S, the furrow slope, is in per cent. Generally, cross-slopes in furrow irrigation should be less than the major ground slope down the furrows to limit the furrow flows breaking out. 30.2.2.4 Corrugation irrigation Corrugation irrigation is a variant of furrow irrigation in which the furrows are very small. It is suitable for medium soils only

Table 30.9 Maximum recommended furrow lengths (m). (After Booher (1974) Surface irrigation. Food and Agriculture Organization Land and Water Development Series No. 3.)
Soil type Furrow slope (%)
0.05 0.1 0.2 0.3 0.5 1.0 1.5 2.0

Fine average depth of water applied (cm) 7.5 15 22.5 30


300 340 370 400 400 280 250 220 400 440 470 500 500 400 340 270 400 470 530 620 560 500 430 340 400 500 620 800 750 600 500 400

Medium 5
120 180 220 280 280 250 220 180

Coarse 10
270 340 370 400 370 300 280 250

15
400 440 470 500 470 370 340 300

20
400 470 530 600 530 470 400 340

5
60 90 120 150 120 90 80 60

7.5
90 120 190 220 190 150 120 90

10
150 190 250 280 250 220 190 150

12.5
190 220 300 400 300 250 220 190

Sprinkler 2

Sprinkler 3 Combined pattern

Sprinkler 4

Application (mm)

Sprinkler 2 pattern

Sprinkler 3 pattern

Sprinkler 4 pattern

,Sprinkler 5 pattern Sprinkler 1 pattern

Distance (m) Figure 30.4 Sprinkler application patterns


overlap of sprinkler patterns is necessary to give a reasonable uniform application as shown in Figure 30.4. The discharge of a sprinkler Q in cubic metres per second: Q = CFJ2gH (30.4) 0.98, F= the cross-sectional area of the nozzle in square metres, g = 9.81 m/s2 and H= the height of hydraulic head behind the nozzle in metres To give a particular precipitation rate over a field there is a range of solutions of sprinkler spacing, nozzle diameter and pressure as shown in Table 30.10.

where: C= the contraction coefficient varying between 0.79 and

Table 30.10 Spacing and precipitation rates of single-nozzle sprinklers. (After Baars (1973) Design of sprinkler installations. Department of Irrigation, Civil Engineering, Agricultural University, Wageningen)
Details of sprinkler Nozzle Pressure size (mm) (atm)
4.0 5.0 6.0 7.0 8.0 9.0 3.0 3.5 4.0 3.0 3.5 4.0 3.5 4.0 4.5 3.5 4.0 4.5 3.5 4.0 4.5 3.5 4.0 4.5 3.5 4.0 4.5 3.5 4.0 4.5 3.5 4.0 4.5

Discharge (m /h)
.02 .11 .19 1.63 .76 .88 2.56 2.74 2.90 3.48 3.73 3.96 4.44 4.74 5.04 5.67 6.06 6.42 7.12 7.60 8.06 8.63 9.23 9.79 10.18 10.88 11.55
3

Diameter coverage (m)


30 31 32 33 34 35 36 36 37 40 41 42 43 43 44 44 45 46 46 47 48 48 49 50 49 50 52

Square and rectangular spacing of sprinklers (m) 12x12 12x18 18x18 18x24 24x24 Precipitation rate (mm/h)
7.1 7.7 8.3 4.7 5.1 5.5 7.5 8.1 8.7

24x30

30x30

30x36

5.0 5.4 5.8 7.9 8.5 90

5.9 6.3 6.7 8.1 8.6 9.2 10.3 11.0 11.7

10.0 11.0 12.0

6.0 6.5 6.9 7.7 8.2 8.8 9.8 10.5 11.1 12.4 13.2 140

7.9 8.4 8.9 9.9 10.6 11.2 12.0 12.8 13.6 14.1 15.1 16.0

9.6 10.3 10.9 11.3 12.1 12.8

Note: Exceeding the line is not recommended for ideal irrigation.

The variation of head between sprinklers is normally limited to 0.2/f where H is the design head at the sprinkler, taking into account any difference in ground level at sprinklers. It is this criterion which limits the lateral pipe diameter and length. Movable laterals are normally made of aluminium or galvanized steel, the aluminium being lighter and hence easier to handle, and the galvanized steel being cheaper and more easily repaired. The supply pipeline can be designed using the normal pipe friction formulae described in Chapter 5. The calculation of head loss in sprinkler lines having sprinklers at constant spacing can be calculated using the Christiansen formula: *-* (30.5)

required, or a barrier layer beneath the root zone. Water is passed to the crop from open feeder ditches via buried perforated pipes. Control of the water level in the ditches determines the quantity of water available to the crops. A combined system of irrigation and drainage is common with the ditches and pipes doubling for both irrigation and drainage. 30.2.6 Irrigation canal design The basic and most common method of designing a rigid boundary channel is the Manning equation. The design method and values of Manning's n are described in section 30.8. Earth canals which transport significant quantities of sediment can be designed, using a regime method or one of the sediment transport formulae described in sections 30.6 and 30.7. 30.2.6.1 Freeboard Freeboard is defined as the distance between the design water level and the canal bank top level. Minimum freeboard above design water level for earth canals can be defined by: /2> = 0.2 + 0.235el/3 with a minimum value of 0.3 m where Fb is the freeboard in metres and Q is the design discharge in cubic metres per second 30.2.6.2 Canal seepage The quantity of water that will seep from the canal is normally measured in cubic metres per second per million square metres of wetted perimeter. Seepage rates for various materials in which the canals are constructed are given in Table 30.11.8 (30.6)

where hz = head loss in the sprinkler line in metres, h = head loss in 100 m line in metres, through which a quantity of water flows which corresponds to the total discharge of all sprinkl6rs on the line, n = number of sprinklers on the sprinkler line, a = spacing of the sprinklers and/= factor which varies with the number of sprinklers, n, as follows:
n 2 3 4 5 6 7 f 0.625 0.518 0.469 0.440 0.421 0.408 n

8 9 10 11 12 13

f 0.398 0.391 0.385 0.380 0.376 0.373

n 14 15 16 17 18 19

f 0.370 0.367 0.365 0.363 0.361 0.360

n 20 25 30 40

f 0.359 0.354 0.350 0.345

30.2.4 Trickle irrigation The basis of trickle irrigation is to provide irrigation water to individual plants. A plastic pipe is run along the ground at the base of a row of plants and water is carried to each plant through orifices in the pipe or using an emitter. Trickle irrigation is more accurately described as localized irrigation as it includes a wide range of emitters such as micro-sprinklers and bubblers. Trickle irrigation is most suitable for row crops and trees and is generally able to use more saline water supplies than surface irrigation or sprinkler irrigation. The design of localized irrigation systems is described by Vermeirei and Jobling.7 30.2.5 Sub-irrigation Sub-irrigation is only suitable for specialized soil conditions. High horizontal permeability and low vertical permeability are

30.3 Drainage of agricultural land


30.3.1 Introduction Agricultural drainage is necessary to remove excess water from the soil to improve the agricultural potential. The benefits of drainage may include: (1) Seed germination - excess moisture associated with low temperatures impairs germination. Waterlogging may cause seeds to rot and not germinate. (2) Crop growth - most crops require air in the root zone to grow. (3) Control of water table - high water tables will limit depth of root zone. (4) Disease - waterlogged crops are more susceptible to disease. (5) Yield gain - generally higher crop yields are experienced from drained land. (6) Poaching - wet soil that carries stock experiences surface damage by grazing animals. (7) Cultivation - improved drainage will allow easier access for cultivation machinery. (8) Salinity - control of salinity in crop root zone. Drainage systems can be defined as subsurface and surface. Surface drains are designed to remove excess runoff from the land which would otherwise cause localized flooding. Subsurface drainage is designed to remove excess water from the soil mass. It is discussed in the following sections.

Table 30.11 Seepage rates from canals. (After Etcheverry (1915) Irrigation practice and engineering. McGraw-Hill)
Type of soil Impervious clay loam Medium clay loam Clay loam or silty soil Gravelly clay loam or sandy clay or gravel cemented with clay Sandy loam Sandy soil Sandy soil with gravel Pervious gravelly soil Gravel with some earth Seepage losses (m3/s per million m2) 0.8-1.2 1.2-1.7 1.7-2.7 2.7-3.5 3.5-5.2 5.2-6.4 6.4-8.6 8.6-10.4 10.4-20.8

30.3.2 Sub-surface drainage of irrigated land Sub-surface drainage for irrigated lands in arid areas is normally associated with the control of the water table depth. Most crops grow best with the water table below their root depth although crops may not be affected by a higher water table for a short period. Rice is an exception since it grows well in totally waterlogged conditions. Recommended minimum water table depths are shown in Table 30.12. The necessary drainage is frequently achieved by providing perforated drainage pipes below ground at regular intervals. It is necessary to install the drains below the desired design water table depth.

Table 30.13 Estimated recharge to watertable as related to irrigation method and soil type
Average recharge as percentage of irrigation water delivered to the field Application practices Soil texture heavy light Daytime application, moderately strong wind 30 Night application 25 15 Poorly levelled and shaped 30 Well levelled and shaped 20 Poorly graded and sized 30 Well graded and 25 sized

Irrigation method Sprinkler

Table 30.12 Minimum water table depths


Water table depth below ground level (m) Fine textured (permeable soil) Light textured soil 1.2 1.1 1.6 1.0 1.0 1.2

Trickle Basin

30 25 15 40 30 40 35

Crop Field crops Vegetables Tree crops

Furrow, border

The shallowest drain depth for water table control is: //+0.5/z+ 0.Im where H= design water table depth given above and H = rise in water table resulting from the maximum individual recharge from a water application. 30.3.3 Drainable surplus The quantity of water to be removed by a subsurface drainage system can be estimated from a water balance: Qs = Rf+Sc +S-Dn (30.7)

Approximate design drainage rates are likely to be in the following ranges: Less than 1.5 mm/day 1.5 to 3.9 mm/day 3.0 to 4.5 mm/day More than 4.5 mm/day For soils having a low infiltration rate. For most soils, with the higher rate for more permeable soils and where cropping intensity is high. For extreme conditions of climate, crop and salinity management, and under poor irrigation practices. For special conditions, e.g. rice irrigation on lighter textured soils.

where Q5 = water to be removed by drainage, Rf= recharge to the water table from rainfall or irrigation, Sc = seepage from canals or rivers, S{ = groundwater flow into the area and Dn = groundwater flow out of the area. Recharge (7?f) to the water table will vary with soil type, irrigation method and efficiency of water management. Food and Agriculture Organization Paper No. 389 Drainage design factors gives the estimated recharge for various conditions as shown in Table 30.13. Seepage from canals can be estimated using Table 30.11. Groundwater inflow and outflow can be calculated from data on groundwater slope, flow cross-section and soil permeability using Darcy's law, which states that:
K=^ L

30.3.4 Drainage of lands subject to excess rainfall The drainage of irrigated land in arid areas is described above. However, many areas require drainage due to an excess of rainfall. The drain discharge due to rainfall rises to a peak following a rainstorm and then recedes. For the design of a buried pipe-drainage system, the discharge is often based directly on rainfall data. For instance, in the UK field drainage design is based on 5-day rainfall divided by 5 to give the daily drainage rate with return periods as shown in Table 30.14. Typical drainage rates in northwest Europe would be of the order of 7 to 10 mm/day. Drainage systems incorporating mole drainage are normally based on a 1-day rainfall value, because of the shorter response. The design depth to water table is often taken at 0.5m for shallow rooted crops and 0.75 to 1 m for deep-rooted and highvalue crops. Drains in the UK, in practice, are usually laid at

(30.8)

Table 30.14
Crop Specialist high value crops Horticultural Roots Intensive grass, cereals Grassland Design rainfall exceedance 1 yr 1 yr 1 yr 1 yr 1 yr in 25 in 10 in 5 in 2 in 1

where V= flow velocity in metres per day, K= hydraulic conductivity of the soil in metres per day, and h/L = hydraulic gradient.
And Q= VA

where Q = flow in cubic metres per day and A = area of flow in square metres

depths ranging from 0.75 m in low permeability soils to 1.25 to 1.5 m in permeable soils. A more detailed discussion of drainage discharge design is given by Smedema and Rycroft.10 30.3.5 Drain spacing The required spacing of drains can be calculated using the Hooghoudt equation:
T2_^dh.4KJ^

lated. For more details, see van Beers's work.11 Typical values of hydraulic conductivity are given in Table 30.15. A more detailed explanation of the calculation of drain spacings is given in ILRI Bulletin no. 8.12 30.3.6 Drain flow Drain pipe sizes can be calculated using the Darcy-Weisbach equation for smooth pipes and Chezy-Manning for corrugated pipes. For which have a constant discharge along their length: Q = 89^2-711 -57 smooth pipes Q = 38^2-67/ 50 corrugated pipes where Q = discharge in pipe, in cubic metres per second, cp pipe internal diameter, in metres and /=hydraulic gradient, in metres per metre It is common to 'over design' the pipe to allow for some siltation with the drain capacity normally increased by some 30%. It is normal to assume that the hydraulic gradient line coincides with the pipe soffit, i.e. the pipe flows full. If the drains are installed in hydraulically unstable soils they will require to be surrounded by a gravel envelope. Generally, soils with a high clay content will be stable and will not require an envelope. Granular envelopes are normally 50 to 100mm thick. The gradation of the filter should be designed using the US Bureau of Reclamation method.13 30.3.7 Drainage layouts Typical layouts of a buried drainage system, regular and irregular are shown in Figure 30.6.

q~

(30.9)

where Ka = hydraulic conductivity above the drain in metres per day, Kb = hydraulic conductivity below the drain in metres per day, h = height of water table above the drain level midway between the drains in metres, q = drain discharge in metres per day and d= equivalent depth - function of depth to impermeable barrier (D) and drain spacing (L) (see Table 30.15)

Table 30.15 Equivalent depths (d) for 80mm corrugated PVC pipe drains
L(m) 5 D(m) 0.25 0.5 0.75 1.00 1.25 1.50 1.75 2.0
10 15 20 25 30 35 40 45

0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.43 0.46 0.47 0.48 0.48 0.49 0.49 0.49 0.49 0.53 0.62 0.66 0.68 0.69 0.70 0.71 0.71 0.72 0.59 0.74 0.81 0.85 0.88 0.90 0.91 0.92 0.93 0.62 0.83 0.93 1.00 1.04 1.07 1.09 1.11 1.12 0.63 0.89 1.03 1.12 1.18 1.22 1.26 1.28 1.30 0.64 0.94 1.11 1.22 1.30 1.36 1.40 1.44 1.47 0.64 0.97 1.17 1.31 1.41 1.48 1.54 1.58 1.62

The Hooghoudt equation allows two layers of soil with differing hydraulic conductivity (A8, Ab) (see Figure 30.5). Values of hydraulic conductivity can be measured in the field using the auger hole method. Alternatively, the designer can use values measured on similar soils elsewhere. The single auger hole method requires a hole some 80mm in diameter to be bored below the water table. The water in the hole is then pumped or baled out and the rate at which it refills is measured. From these measurements the value of hydraulic conductivity can be calcu-

Open Main Drain

Buried pipe field drains

Open or buried pipe collector drain

Figure 30.5 The Hooghoudt equation (definitions) Table 30.16 Hydraulic conductivity (m/day)
K (m/day)

(a) Regular Open Main Drain Open or buried pipe collector drain

Coarse gravelly sand Medium sand Sandy loam/fine sand Loam/clay loam/clay, well structured Very fine sandy loam Clay loam/clay, poorly structured Dense clay, not cracked and no bio-pores

10-50 1-5 1-3 0.5-2.0 0.2-0.5 0.02-0.2 < 0.002

(b) Irregular Figure 30.6 Typical layouts of buried drainage systems

Collector drains can be open ditches or buried pipes. Buried pipe collectors are to be preferred where sufficient ground slope is available. Pipe drain slopes should not be less than 0.0005 whilst open collector slopes can be as low as 0.0001. To allow drains to be cleaned, they should not exceed 30Om in length without a manhole or outfall into an open channel. 30.3.8 Drainage of heavy soils For soils with very low permeability it becomes uneconomic to install drainage systems with buried field drains at spacings of between 1 and 5 m as are indicated by the use of the Hooghoudt equation. A common solution is the use of mole drainage. Moles are installed by using a mole plough that draws a 75-mm diameter bullet through the soil at a depth between 400 and 600 mm. The mole forms a tunnel in the soil and some fissuring in the upper soil area. Mole drains are normally spaced at 1 to 3 m and drawn across the line of the collector drains which have permeable fill in the pipe drain trench above the drain (see Figure 30.7). Collector drains are normally spaced at 20 to 60 m. Moling is best suited to clay soils with a minimum clay content of 30% and the moles have a relatively long life in stable calcareous clays. However, remoling will be necessary on average every 5 yr or so. Efforts have been made to increase the life of mole drains by filling the tunnels with gravel. However, this is very expensive and for normal field cropping is not economic.

30.3.10 Surface drainage for irrigated land Surface drainage is often provided to irrigated land to collect excess irrigation supplies and runoff from rainfall. For surface irrigation typical surface drain capacities can be based on 24 h, 1 in 5yr rainfall with 24 to 48 h storage on the field. For rice drainage, the drain capacity should be sufficient to allow the drawdown of water in the paddies where this is part of the cultivation pattern. Typical values of surface drainage capacity are in the range of 2 to 41/s per hectare.

PART B: LAND DRAINAGE AND RIVER ENGINEERING 30.4 Land drainage and flood alleviation
30.4.1 Objectives of land drainage The drainage of agricultural lands has already been discussed in the first part of this chapter. To the river engineer the term 'land drainage' has a broader interpretation, encompassing both the removal of excess water and the prevention of flooding of the urban as well as the rural environment. In general terms, the problem of ineffective land drainage occurs when inflow into the system exceeds outflow, so that there is a build-up of water over a period of time. This may occur rapidly over a few hours in response to heavy rainfall, or it may be a gradual rise in water table during wet periods. Flooding occurs when a channel has inadequate capacity to convey the amounts of water flowing into it, or when flood defence works fail. Thus, the solutions to land drainage problems invariably involve either control of inflow into the system or works to improve the capability of the drainage channels to carry flows through the system. The basic objective is to reduce the frequency and/or the intensity of inundation to acceptable levels, appropriate for the situation. 30.4.2 Rivers as natural drains Rivers are the Earth's natural drainage channels, conveying surface flow from the land to the sea or to inland lakes and marshes. Some rivers are essentially ephemeral (wadis) and flow for only very brief periods, often with very high discharges and consequently devastating erosive power. Others are seasonal, being dry for part of the year, but flowing steadily during the wetter months. Others still are perennial, generally flowing throughout the year but with varying intensity. Most rivers in Europe fall into this latter category. No two rivers are the same, but rivers exhibit similarities which, to a certain extent, can be defined mathematically, thus enabling engineers to assess the problems with which they are faced. Perhaps the most fundamental property of a river is its flow or discharge. However, as has been indicated above, this is not a fixed property - the flow varies both spatially and with time. There are ways in which the flow in a river can be controlled or reduced, but often the engineer is faced with the problem of designing a structure or a scheme which is capable of withstanding the flow which passes through a specific point or reach of the river. It is therefore necessary to estimate the river flow for which the scheme or structure must be designed and this involves an exercise in statistics which is described later in this chapter.

Soil fissuring

Mole Plough

Mole channel Permeable fill

Direction of ploughing Pipe drain

Figure 30.7 Mole drainage


30.3.9 Bedding systems Bedding is a common method for the drainage of flat heavy land subject to excess rainfall. Wide beds are most suitable for mechanized agriculture and are up to 30 m in width. Drainage is mainly by surface runoff with some interflow in the topsoil region as shown in Figure 30.8. The shallow drains are normally some 0.5m in depth. The raised crowned beds are normally built up over time by ploughing in such a way to turn the soil towards the centre of the bed.

Rainfall TOPSOIL Overland - flow ' DRAIN

Interflow

IMPERMEABLESOlL Figure 30.8 Wide bedding

30.4.3 Economic issues Since funds are limited and there is always competition from other potential schemes, it is necessary to undertake some form of economic evaluation of proposed drainage improvement works. Such an evaluation requires the estimation of the benefits which might accrue from the scheme and the costs of its implementation. For an urban flood relief scheme, some of the benefits are obvious and can be evaluated in a straightforward manner. Elimination of the physical damage caused by flooding is one such benefit, which can be assessed by counting the cost of replacement or repair of goods and property so damaged. In addition, there are less tangible costs of flooding which must be evaluated, such as loss of production due to flooding of industrial properties and disruption to traffic resulting from flooded roads. These too must be estimated. Finally it is now common practice to evaluate the intangible factors such as the distress caused to the public by flooding, particularly to those people in a high-risk area. From a knowledge of the frequency of flooding the present value of all the 'damage' likely to occur during the lifetime of the proposed works can be estimated. The benefits so derived should then be compared with the estimated costs of the works so that competing schemes can be compared on a similar basis or to determine the most economic level of protection which could be provided. For agricultural lands it is possible to estimate the increased value of production generated by improved drainage, although this can involve some fairly subjective assessments. In general, the agricultural benefit will accrue as a result of either a lowered water table and/or a reduced risk of periodic flooding, both enabling a wider range of crops to be grown and/or better yields to be achieved as well as extending the period for which agricultural operations are possible and improving 'traffickability' of the land. Thus, an estimate of the increased value of annual production is made possible by the drainage works and this figure is capitalized over the life of the scheme to determine the benefits. As with the urban scheme the benefits are then compared with costs as a means of evaluating schemes.

relationship between level upstream of a structure and discharge. Flumes and weirs are commonly used on small rivers whereas velocity-area methods are usually applied to mediumand large-sized channels. In recent years, permanent flow measurement installations using electromagnetic or ultrasonic gauging techniques have been developed as an alternative to weirs and flumes. Essentially, these methods measure velocity at a defined section. Where records of channel flow are not available, rainfall records may be used to estimate likely flows. Within the UK rainfall records are maintained by the Meteorological Office, which operates over 5500 gauges, and flow data are archived by the Institute of Hydrology.

30.5.3 Statistics
Methods of statistics14 are frequently used in hydrology to estimate the return periods of natural events. A flood flow is said to have a return period of, for example, 50 yr if, on average over a long period of time, that flow is equalled or exceeded once in a 50-yr period. Frequency analyses are required so that standards of protection can be met, risk assessments made and economic analyses undertaken. The Fisher-Tippett type 1 extremal distribution (commonly known as the Gumbel distribution) is often used to analyse annual maxima series of discharge and rainfall. For a series of data values, QM, the magnitude of the event of return period Tyr, QT, is given by: CT = <2A + d(0.78>> -0.45) (30.10)

I<2M whereC? A = ,

/^" n 2\ , = ^_ (__- Q^ ,

y loge f loge ( 1 J J and n = number of years of data The probability, PN, of an event of return period T yr being equalled or exceeded during a period of TV yr, is given by: /> N =1-[1-(1/D] N (30.11)

30.5 Hydrology
30.5.1 Introduction The design of river engineering and land drainage works is based on hydrological criteria, predominantly estimates of channel flow and its variation with time. The ideal basis for the calculation of design parameters is a long period of recorded data which can then be analysed using statistical methods. Such data are often not available, but a record from a neighbouring catchment may be, and this can be corrected for use in the area concerned. Even short periods of data are useful, but if no records exist or their reliability is doubtful, empirical techniques of parameter estimation can be employed. 30.5.2 Measurement The measurement of channel flow (discharge) is most commonly undertaken by velocity-area methods or at flow-measuring structures. Flows are measured over a range of stages (water levels) so that a stage-discharge relationship can be developed. Velocity-area methods depend upon the use of a current meter to record velocities and a knowledge of the cross-sectional area to which the velocity measurements can be applied, the product of these two variables being discharge. Flow-measuring structures are operated on the principle that there is a unique

Typical acceptable frequencies of flooding within the UK are once a year for grassland, once in 10 yr for arable land and once in 100 yr for urban areas. Where the risk to life is high, as in coastal situations, standards of protection even higher than those for urban areas may be considered. 30.5.4 Flood flow calculation methods If statistical methods cannot be used, the design flow may be estimated from empirical equations relating rainfall and/or catchment characteristics to runoff. Early formulae to evaluate the 'maximum flood', max, were of the form: Qa** = CA" (30.12)

where C = coefficient depending on the type of climate and catchment, A = catchment area and = an index, usually between 0.5 and 1.2 The rational formula is representative of the many formulae developed to relate rainfall to peak runoff and is of the form:

Q = CiA

(30.13)

where g = peak discharge, C= runoff coefficient, depending on the characteristics of the catchment, /= rainfall intensity and A = catchment area This approach is based on the assumption that maximum flow occurs as the result of the maximum rainfall intensity to be expected within the 'time of concentration' of the catchment. 'Time of concentration' is the time taken for rainfall falling on the most remote part of the catchment to reach the part of the drainage network under consideration. Nowadays the rational formula is generally only used for urban drainage design, for which the assumptions are valid. Rainfall formulae have also been developed relating the intensity, duration and frequency of events. The Bilham formula15 related these for the UK by the following equation: r = 25.4[(1.25777V)0282-0.1] (30.14)

hydrograph plus design storm - the latter is given more weighting in the report. However, for preliminary estimates of peak discharge the simplicity of the former technique is very appealing. Use of the Flood studies report is described in detail by Sutcliffe.18

30.6 Channel regime


30.6.1 Regime flow A channel is said to be in regime when, over a hydrological cycle, the channel shows no appreciable change in its width, depth or gradient. Regime theory postulates that for a stable channel there is a relationship between the channel parameters of width, depth, gradient and flow. Thus if any one of these four parameters is artificially (or naturally) changed, the channel will adjust itself so that regime conditions are re-established. Also fundamental to river regime is sediment transport, which is discussed in more detail later. Sediment is material which is picked up, transported and deposited by the river. It can vary from very fine clay particles, often referred to as wash load, to large cobbles and boulders which can be moved by a river in flood. A regime channel is generally transporting sediment which is similar in size to the material which forms the bed and banks of the channel. Many regime theories have been developed from a study of irrigation canals and the application of regime theory to natural rivers raises the problem of what flow should be taken as significant in determining the dimensions of the stable channel. The low flows which go on for most of the time cause little or no change in the channel section, while the maximum flood flows which occur for a few hours at intervals of several years cause rapid but temporary changes which the river will subsequently tend to restore. The 'bank-full' flow is commonly used in the UK, where it has been quantified as the flow which is exceeded for 0.6% of the time (i.e. about 2 days per year on average). 30.6.2 Regime formulae Of the many regime formulae postulated, that of Lacey is probably one of the most useful. The Lacey formulae can be expressed as:

where r = total rainfall in millimetres, T= duration of storm in hours and TV= probable number of occurrences in 10 yr

30.5.5 Hydrographs A more accurate approach to flood peak and volume prediction, which has been widely accepted and developed since its inception by Sherman, is the unit hydrograph concept. A hydrograph is a plot of discharge versus time (Figure 30.13(b)) and, as such, gives the engineer much more information than just a peak flow estimate. The unit hydrograph of a catchment is defined as the hydrograph of direct runoff resulting from a unit of effective rainfall generated uniformly over the catchment at a uniform rate during a specified period of time. Where runoff and rainfall records are available the catchment's unit hydrograph can be derived from these data and used to produce hydrographs of runoff for any given rainfall profile. Since such data are often not available, synthetic unit hydrograph techniques have been developed which relate catchment characteristics to the parameters of the unit hydrograph. The US Soil Conservation Service (USSCS)16 and the Environmental Research Council17 have presented such techniques. 30.5.6 Curve number method The curve number method developed by the USSCS uses the following methodology: (1) Rainfall is converted to discharge using a curve number graph based on catchment characteristics. (2) Discharge is developed into a basin hydrograph using the USSCS unit hydrograph. (3) The design drainage rate is taken from the peak of the hydrograph. This method is described in detail in the USDA National engineering handbook, section 4. 30.5.7 The Flood studies report In the UK, the Flood studies report provides a comprehensive guide to the estimation of maximum floods, return periods and flood volumes for any site. The methods described apply to both gauged and ungauged catchments. Of the two methods developed for ungauged catchments - the mean annual flood plus growth curve and the synthetic unit

dm

Ws = 4.83eQ]/2 _2A6V2
j-

_WE Ql/6

where Q = design flow in cubic metres per second, 5=channel slope, ^5 = water surface width in metres, dm = mean depth = A/Wt in metres, A = cross sectional area of flow in square metres, V= flow velocity = QjA in metres per second, E= shape factor = P/ W^ P=channel wetted perimeter in metres, and e = width factor and B=bed width (generally taken as IV5). The values of K and x vary with the size of sediment present in the channel bed, as shown in Table 30.17. The shape factor E takes account of the differences between wide shallow channels and narrow channels. For the former, the value of E approaches unity. The width factor e varies with soil type, in response to erodibility. Its value may range between 0.7 for stiff soils (clays) to 1.0 for erodible soils (sands and silts). The silt factor / depends on the size of sediment which will form the channel bed in the long term. Indicative values of/are

Table 30.17
Sediment median grain size, D50 (mm) Z)50 < 0.2 0.2 < D50 < 0.6 0.6 < Z)50 < 2.0 Z)50>2.0 K x

0.000206 0.000274 0.000303 0.000188

$ i f

0.4 to 0.6 for clays and 1.0 for sands. Actual values adopted should be based on local experience of stable channels. In the context of rivers in the UK Nixon19 carried out an interesting study of channel characteristics in 1959. He examined twenty-nine British rivers and attempted to establish regime formulae which were independent of bed material properties. His formulae are thus a simplification of reality, since they imply that the channel cross-section is independent of the material from which it is formed. Nevertheless, they can be used as a guide to channel dimensions for British rivers, although it should be noted that they are not applicable to channels with gravel beds. The Nixon formulae (converted to metric units) are:
^s = 3.02 b05 ^ = 0.552-' F=0.6ie b 17 and A = LMQM*

ing on the mechanism of its transport. The 'wash load' comprises relatively fine material and the rate of wash load transport is mainly determined by its rate of supply from the drainage basin rather than the transport capacity of the stream. This material settles out rather slowly and can be maintained in suspension in large quantities by relatively slow-flowing water. Particles of 0.06 mm or finer are often considered as the wash load fraction. More accurately, wash load can be defined as that fraction of sediment which is finer than the Z)10 size of material (Z)10 = SiZe for which 10% is finer) found on the channel bed. In contrast, the 'bed material load' transported is almost entirely a function of the transporting capacity of the flow. The use of the term 'bed material' indicates that this is what the sediment load mainly comprises. It should not be confused with 'bed load' which has been used in the past to describe the larger particles transported near the bed of the channel. Sediment loads are normally expressed as parts per million (p.p.m.) by weight (i.e. 1 g of sediment in 1 g of water= 1 p.p.m.). Transport rates are generally expressed in tonnes per day.

where W^ A and Fare defined as above, <4 = mean depth = A/ Ws in metres and Qb = bank-full flow in cubic metres per second

30.6.3 Practical applications Regime equations can be used as a guide in the design of new channels or of remodelling works to existing channels. In the absence of any other information the characteristics of a representative reach of existing channel will give a good indication of what is appropriate for new works in terms of channel width and depth, bed slope and bend radius. It is worth remembering that, if it is necessary to improve the conveyance of a river channel (as part of a flood improvement scheme) it is better to increase the depth of flow and/or the slope rather than the channel width, because natural adjustments of the former tend to occur at a much slower rate. A channel may be capable of carrying the required flow with smaller dimensions and a steeper gradient than the regime values but the higher velocities generated will increase the width and depth by scour and will reduce the gradient by deposition of the eroded material at the lower end of the reach. Similarly, many channels have been made with excess width to carry the maximum flood flows, but sediment deposition has subsequently resulted in a narrower meandering deep channel within the main channel. The most common error arising from lack of knowledge of the regime theory is when a river is straightened by cutting across meanders without considering whether the resultant shortening of the course and increase of slope will produce scouring velocities.

30.7.2 Sediment transport estimates Sediment transport estimates are most commonly required so that the rate of scour or deposition in a channel can be predicted. In northern Europe problems of sediment transport are generally modest and concentrations of less than 1000 p.p.m. are common. However, in tropical or arid zones, sediment loads in rivers in flood can exceed 20 000 p.p.m., and much higher loads have been recorded in extreme cases. Clearly, if such highly charged water is diverted from a river into, for example, a slow-flowing irrigation canal, there will inevitably be extensive deposition in the canal. Sediment transport estimates may also be required in the design of river regrading works where it is necessary to check whether the changes imposed (such as steepening the river gradient) are likely to lead to excessive erosion. Of course, wherever possible, attempts should be made to measure sediment transport rates in situ. This is a notoriously difficult operation and even the most carefully controlled sampling can yield widely differing results. This is not only because the sampling technique is prone to error, but also because extreme sediment loads occur in relatively short-lived floods which are unpredictable.

30.7.3 Sediment transport equations There are many sediment transport equations, none of which can claim a high degree of predictive accuracy. An estimate which is in the range 0.5 to 2 times the actual value is all that can be reliably expected. The recent Ackers-White20 equations are amongst the most accurate although these are rather cumbersome. The more simple Engelund and Hansen21 equation yields similar levels of accuracy. The Engelund and Hansen equation for bed material load can be expressed as: 1600OyFWS1-5 (S-D2D50 where A'= sediment concentration in parts per million, 5 = sediment specific gravity (normally 2.65), V average flow velocity in metres per second, d= average channel depth in metres, S= channel slope and Z)50 = median sediment size of bed material in metres Thus, a channel flowing at 1.0 m/s at a depth of 1 m with a slope

30.7 Sediment transport


30.7.1 Basic concepts Sediment load is generally divided into two categories depend-

of 0.5 m/km and a median sediment size of 1 mm would transport: 16 000 x 2.65 x 1.0 x 1.0^(0.0005)'5 , _ . 1.6? x 0.001 =174p.p.m.

30.7.4 Stable channel design Recent research by White, Paris and Bettess20 has resulted in the publication of a set of Tables for the design of stable alluvial channels. These tables have been derived from the results of extensive flume experiments, and list values of sediment size, sediment concentration, channel flow, flow velocity, channel slope, depth of flow, channel width and friction factor. Given a sediment size and any two of the other parameters, it is possible to estimate values for all the other variables.

unchecked, can reduce the capacity of a channel to a fraction of its design capacity. Thus, whereas it is feasible to construct an earth channel which will have an n value of 0.025, experience has shown that even modest vegetative growth will increase this to 0.035. The problem is worse in tropical zones where plant growth is prolific, and irrigation canals can require clearance every few months. It is therefore recommended that, for such channels, a minimum value of n = 0.030 is adopted, with higher values if it is known that maintenance will be infrequent. 30.8.2 Channel stability Manning's equation gives us a simple tool for determining the channel size, but gives us no information as to the long-term stability of the channel. It is most important, therefore, that a check is made on likelihood of sedimentation or scour occurring. Sedimentation has already been discussed. To assess the scouring potential of a stream it is common to determine the tractive force, defined as: -C0 = CyRS (30.15)

30.8 Channel design


30.8.1 Channel flow formulae The most commonly used and universally accepted channel flow equation is that of Manning, which can be expressed in terms of flow as: ARW12 n where Q = channel flow in cubic metres per second, A = channel cross-sectional area below water level in square metres, R = hydraulic radius, A/P in metres, P=wetted perimeter in metres, Schannel slope and n Manning's roughness coefficient The selection of an appropriate value for Manning's roughness coefficient, which in normal engineering practice lies in the range 0.010 to 0.150, is a matter of judgement and experience. Ven Te Chow22 gives comprehensive guidelines, with values for all common situations supported by photographs of typical channels. Table 30.18 gives some illustrative values. In selecting an appropriate value of Manning's n the effect of future changes in the nature of the channel must be considered. Perhaps the most significant factor is vegetation which, if left

where T0 = unit tractive force in newtons per square metre, Y = water specific weight (9810 newtons per cubic metre), S= water surface slope, R = hydraulic radius in metres and C=coefficient depending on the shape of the channel and the part of the channel considered Unless the channel is particularly narrow, values of C of 1.0 for the bed and 0.76 for the banks are usually assumed. For non-cohesive soils there are recommended values of tractive force on the channel bed for a range of soil types, recommended limiting velocities of flow are also given. Guidelines are given in Table 30.19.

Table 30.19 Suggested limiting tractive force (TO) and flow velocity (V) values
Material Clear water Water transporting colloidal silts V T0 (m/s) (N/m2)

V (m/s)
Fine sands and non-colloidal silts Firm loam Stiff clay and colloidal silts Fine gravel Graded colloidal silts to gravel Coarse gravel Cobbles and shingles

\ 2 (N/m )
1.9 3.6

Table 30.18 Values of Manning's n*


Surf ace I channel Concrete lined channel (smooth finish) Brick-lined channel Mortared rubble masonry Earth channel: clean, uniform Earth channel: very overgrown with weeds, etc. Normal range of n (design value) 0.012-0.017 (0.015) 0.012-0.018 (0.015) 0.017-0.030 0.020-0.030 0.050-0.120 (0.025)

0.55 0.75 1.15 0.75 1.20 1.20 1.50

0.85 1.00 1.50 1.50 1.65 1.50 1.65

4.0 7.2

12.5 3.6 20.6 14.4 43.6

22.0 15.3 38.3 32.0 52.7

Source: Etcheverry.(1915) Irrigation practice and engineering. McGraw-Hill.

Minor stream: clean, straight 0.025-0.033 (0.030) Minor stream: sluggish, weedy with deep pools 0.050-0.080 Major stream: regular section Floodplain 0.025-0.060 0.025-0.15Of

Notes: *Assumes channel flowing at or near full stage, lower flows may result in higher n values because of the relative significance of obstructions. fOn floodplains the value of n depends very much on the type of vegetation, its height and the season. A typical value for short grass might be 0.030, whereas dense brush in summer might be 0.100.

The figures in Table 30.18 can be used as a guide to determining limiting slopes for channels in terms of the movement of bed material. For the channel sides, even though the tractive force is lower, the banks may be less stable because of the effect of gravity. In practice it is found that, for fine non-cohesive material, small amounts of sediment in the water tend to cement the particles together and the use of tractive force theory is conservative. However, side slope stability should be considered for coarse non-cohesive materials (medium-sized gravels and above).

30.8.3 Other considerations The channel shape, as well as its size, is also important. For irrigation canals and drainage channels bed width to depth ratios (B :d) of between 3 and 4 are often used. Channels designed using the Lacey regime equations tend to have larger B\d ratios. Such channels have lower tractive force values but this principle cannot be extended too far since drains with beds which are too wide will tend to form sub-channels at lower flows leading to a higher local tractive force and, hence, erosion. Channel side slopes depend mainly on the nature of the ground in which they are cut. Slopes of 1:1.5 (vertical: horizontal) and 1:2 are quite commonly adopted, with flatter slopes of 1:3 or even 1:5 if the bank material is highly erodible. Other considerations may dictate the choice of side slope such as use of the slopes for grazing (where flow is intermittent) and ecological factors (e.g. desire to re-establish reed growth, hence shallow slopes). For bends in channels a rough guide to the appropriate bend radius is I Q x W^ (W^ = water surface width), but this takes no account of the erodibility of the bank material. Lacey proposed design radii of 12SV (? (Q = design flow in cubic metres per second) which is recommended where space is available for channels in fine alluvial soils. If possible, the designer should measure the radii of other stable channels in the area to give a guide to acceptable values. For lined channels (concrete, brick, masonry, etc.) a minimum radius of 3 Ws is recommended, with 5 W^ being used where possible.

be given to forming a bypass channel which would leave a sensitive reach of channel untouched. The bypass would be set at a level where it only operated in flood conditions, thus preserving the main stream for all normal flows and allowing use of the land taken up by the bypass for grazing. 30.9.2 Realignment Natural rivers and streams often follow a meandering course, and lower flood levels in a particular area can be obtained by straightening the course downstream by diversions across meanders or by a completely new channel. Realignment will eliminate sharp bends where erosion takes place and will produce a more stable course, but in rivers of high amenity value long straight reaches will lead to complaints that the river has been converted to a 'canal' and in such cases a course with sweeping curves is preferable. With any realignment exercise it is essential to check that the increased slope which will result will not lead to instability. In extreme cases the shortening of a river reach can result in upstream progressing degradation which could undermine bridge foundations or cause the collapse of river frontages. Where realignment results in a gradient steeper than the stable regime gradient, the provision of weirs may be necessary to limit erosion. In small channels, or in short lengths of larger channels, bed scour may be prevented by the use of some form of flexible bed protection such as dumped stone or gabion mattress. 30.9.3 Revetments and lining 30.93.1 Introduction A revetment is any means of protecting a channel bank from erosion or undermining. Revetments are frequently required on river bends, in the vicinity of structures (where flow may be more turbulent), where the natural bank material is unstable, and where the wash from boats causes progressive collapse of banks. Revetments may be rigid (e.g. sheet piling) or flexible (e.g. dumped stone). Channel lining is used where both bed and bank scour are to be prevented or where it is necessary to streamline the channel. Linings, too, can be rigid (e.g. reinforced concrete) or flexible (e.g. gabion mattress). Flexible linings are often preferable because they will accept some settlement or damage whilst retaining their appearance and integrity. Typical details of rigid and flexible revetment systems are illustrated in Figures 30.10 and 30.11. The most important considerations in revetment design are: (1) Adequate scour protection at the toe to prevent undermining. (2) Flexibility if settlement is likely. (3) Adequate weephole provision if the revetment is impermeable and rapid drawdown of river level is possible. (4) Cost (taking into account locally available materials, especially in developing countries, and availability of inexpensive labour). (5) Environmental acceptability. 30.9.3.2 Traditional methods One of the earliest methods of bed and bank protection was the use of brushwood faggots or fascines and these are still used in certain conditions. For revetment purposes the brushwood is made up into tight bundles laid side by side end-on to the channel, with successive layers stepped back to conform with the bank slope. The bundles are held down by stakes but, after a

30.9 Channel improvements


30.9.1 Channel clearance In northern Europe the clearance of trees, brushwood and weeds offers greater improvement in reduced flood levels in relation to cost than any other form of channel improvement. Such works should be executed with a sympathetic approach since our rivers are frequently areas of great natural beauty and are well used by anglers and for other recreational interests (see Figure 30.9). Some very valuable advice in this respect can be found in a handbook published by the Royal Society for the Protection of Birds and the Royal Society for Nature Conservation.27 The aim should be to retain an appearance which is as natural as possible without prejudicing the aim of improved channel conveyance. As an alternative to channel clearance, consideration should

Unsympathetic treatment Bank established with rye grass Flood level

Tree lost

Sympathetic treatment Bank Pond excavated accessset back to improve and increase marginal] to create fill habitat for bank Flood leve Bank sown with grass and wild flowers and set back to retain tree Figure 30.9 Conservation in river engineering Courtesy. Nature Conservancy Council (1983) Nature Conservation and River Engineering

Maximum water level Precast concrete wall unit Backfill Original bed Mass concrete Steel sheet piling Precast Wall with Sheet Pi Ie Toe

Concrete cope

Maximum water level

Tie rod Anchor block

Steel sheet piling

Anchored Sheet PiIe Wall

Maximum water level Flexible protection Can be formed from a fabric mattress filled with cement grout (placed under water)

Bed level

Concrete Lining

Mortared stone pitch in Concrete backing Concrete toe Mortared Pitching Figure 30.10 Typical rigid revetments Bed level

Maximum water level

Bed level

Reinforced Concrete Wall

few floods or tides the whole mass is impregnated with silt which holds and preserves the brushwood. Brushwood is also used in longer lengths to build up large mattresses which are floated into position and sunk by loading with stone to protect the river bed or the lower bank slopes. Stone is also much used for revetment work. This may be in the form of dumped stone (usually machine placed), dry stone pitching (hand placed), pitching grouted with bitumen, or mortared stone pitching, the latter being rigid. The following

formula may be used to estimate the minimum stone size required for a given flow velocity: 0.01 IV6S (s -I)3sin3(p- a) (30.16)

where W= critical weight of stone in kilograms (two-thirds of stones to be heavier), V= flow velocity in metres per second, s=stone specific gravity (assume 2.5 if no other information),

Box gabion Gabion mattress River bed GABION REVETMENT d - Maximum anticipated scour depth Scoured bed

Table 30.20 Gabion mattress thickness


Clays, heavy cohesive soils: maximum water velocity (m/s) 2 minimum mattress thickness (mm) 170 Silts, fine sands: maximum water velocity (m/s) 2 minimum mattress thickness (mm) 230 Shingle with gravel: maximum water velocity (m/s) 3.5 minimum mattress thickness (mm) 170

3 230

4.5 300

Dumped stone Heavy-duty filter cloth (as alternative to gravel filter) DUMPEDSTONE (RIP-RAP) 0.30 m thick hand-placed .stone pitching DRYSTONEPITCHING Precast Blocks fabricated into Mats Filter cloth INTERLOCKINGCONCRETE BLOCK SYSTEM Figure 30.11 Typical flexible revetments Proprietry geotextile mat

3 300

N/A N/A

5 230

6 300

0.20 m thick gravel backing

Sown witf grass seed

Anchoring pegs GEOTEXTILE

p = a stability factor, 70 for random stone (riprap) and a = slope of the bank (i.e. angle with horizontal, in degrees) Note: Impinging velocity may be assumed to be 1.25 x average velocity on the outside of bends. Thickness of stone should be at least 1.5D where D is the effective diameter of the normal size of rock specified. Bank slopes should not exceed 1:1.5 (vertical to horizontal). A filter of graded gravel or geotextile should be provided under the stone to prevent the leaching of fines from the bank. 30.9.3.3 Gabions Gabions and gabion mattresses have been used for many years for revetment and lining work. The gabions are crates formed from wire or plastic mesh and filled with stone. Common sizes are 2 x 1 x 1 and 2 x 1 x 0.5 m. The gabion mattress is similar but comes in units of 6 x 2 m with a range of thickness (up to 500 mm). The gabion crates/mattress are subdivided by diaphragms and are packed with stones of a size generally just larger than the mesh size. Good-quality control during filling is essential to achieve the desired effect. The end-product is a flexible permeable structure of considerable erosion resistance. It is normal to use a filter beneath the gabions to prevent the washing out of fines. Table 30.20 gives indicative mattress thicknesses for a range of conditions. The mattress can be used alone as a flexible lining, or in conjunction with gabion boxes as illustrated in Figure 30.11. 30.93.4 Concrete, geotextile and other methods In recent years the use of concrete block systems and geotextile

revetments has become popular and there are many proprietary systems on the market. Most of these systems are flexible, permeable and allow the growth of grass and waterside plants through them. Most of the concrete block systems have mechanical interlocks which prevent the lifting of individual units, some are connected together by polypropylene strands or glued to geotextile mats so that they can be placed in large units. The incorporation of a geotextile fabric filter under the armouring is common practice. This prevents fines being washed through the armouring in the same way that a graded gravel filter does under stone protection. Various forms of geotextile reinforcement are also available. These can be placed on the soil surface and sown with an appropriate grass seed to give a natural looking erosion-proof surface (in limited erosion conditions). They can even be provided with grass already established. The use of jute fibres for this purpose is also coming into vogue because of its environmental acceptability (it biodegrades within 1 or 2 yr leaving an established grass cover). Steel sheet piling is commonly used in the neighbourhood of weirs, locks or sluices, where there is wave action or heavy turbulence. In fine bed material such as silts, corrugated asbestos cement sheets have been used as an inexpensive form of sheet piling. This material is only suitable to support up to 1 m faces where damage by boats, etc. is not expected. There are many other proprietary revetment systems, any one of which may be appropriate in certain circumstances. These include: (1) Grout-filled mattress (rigid, permeable, can be placed under water). (2) Fabric with pockets which can be filled with soil (allows rapid planting of appropriate waterside plants). (3) Plastic webbing spanning between vertical supports.

30.10 Embankments
30.10.1 Introduction Embankments are provided along river channels to prevent flooding. They are normally set back from the river so that during floods they provide the necessary increase in the waterway section by providing both extra width and extra depth without overflow. In urban areas the land required to set back the embankments may not be available. In fact, there may not be space for the wide-based embankment at all and a flood wall,

or a 'half-bank' supported by a wall may have to be used (see Figure 30.12). Most flood embankments in the UK and northern Europe are of moderate dimensions, not exceeding 4 m in height and many not exceeding 2 m. They are constructed of the best material available, preferably containing some clay to make the banks watertight. Pure clays are not ideal because they tend to crack on drying. Low flood embankments are often constructed from material dredged from the river channel, frequently sands and gravels with varying silt content. This achieves two objectives simultaneously by enlarging the channel and forming flood banks. Such banks may be designed to overtop every 5 or 10 yr or so when providing protection to agricultural land. Provided that the land-side slopes are relatively flat and a good soil and turf cover is provided, overtopping of long reaches of such embankment will cause little damage. Clay cores are not normally provided in embankments although these can be specified where residential properties are protected and even small amounts of seepage would therefore be unacceptable. Steel sheet piling driven down the middle of an embankment has also been used to cure a local seepage problem. 30.10.2 Design Bank top levels should provide a freeboard above the design flood level to allow for settlement and damage by cattle or pedestrians. The freeboard also allows for any inaccuracies in the estimation of flood level. A minimum freeboard of 0.5 m for embankments is normal, with the lower figure of 0.3 m adopted for walls. The bank top width is often about 2 m but may be increased to permit the passage of a tractor along the top with various maintenance equipment. Side slopes on the river face should not be steeper than 1:2 (vertical: horizontal) and on the landward face may be somewhat flatter. Where space permits and fill is available the bank may be constructed with very flat slopes on the landward side so that the area may be mown as part of the adjoining field. Arable cultivation may also be permitted on such slopes provided that the crest of the embankment is clearly demarcated to prevent gradual lowering by farming operations. The planting of trees or shrubs on the embankment should be discouraged. The ideal surface treatment is good turf grazed by sheep or cut mechanically several times per year. Embankment slopes of 1:2.5 can be mown by tractor, but flatter slopes are preferred. Slopes of 1:3 or flatter may be

grazed by cattle, but cows tend to cause more damage than sheep so careful management will be necessary. The hydraulic gradient through the embankment is the slope of a line from the high-water mark on the river side to the landward toe of the embankment and this should not exceed 1:4. In important cases a 'flow-net' through the embankment and its foundation should be calculated to reveal any risk of 'piping' due to excessive rates of seepage. 30.10.3 Stability The subsoil adjoining many rivers in their lower reaches consists of recent deposits of alluvium and is often waterlogged. Such material may be unable to support the increase of superimposed load due to the construction of the flood bank. Unless the soft material is a thin layer, the solution is to construct an embankment of reduced height with a flood wall on top of it. Alternatively, the use of geotextile layers under the embankment may be considered. Appropriately designed, these effectively reinforce the subsoil and reduce settlement. The most serious condition for bank instability arises when the flood level drops rapidly after prolonged retention at a high level. The increased pore water pressure in the embankment has insufficient time to dissipate, shear strengths are reduced and a classic slip failure may result. In practice, with embankments of no more than 4 m high, stability is rarely a problem given a good fill material adequately compacted. However, if problem soils are encountered, the use of geogrid reinforcement or alternative forms of constuction should be considered. 30.10.4 Construction The area of the base of the embankment should be stripped of turf and topsoil to a depth of 0.25 m. This may be stockpiled for later use on the surfacing of the bank. Where the subsoil is weak, or the embankment high, compaction of the subgrade before placing fill should be considered. This will reduce settlements experienced subsequently. The fill should then be deposited and rolled down in layers, generally by bulldozer or tractor shovel. Finally, the embankment will be cased with topsoil and sown with grass seed. The height of the newly completed bank should make allowance for settlement which is likely during the first year at the rate of about one-tenth of the height of the bank, depending on the type of material and the degree of compaction.

Maximum water level Berm Hydraulic gradient River

Maximum water level

Maximum water level

Sheet pile cutoff Figure 30.12 (a) Flood embankment (b) Half-bank with wall (c) Flood wall (d) Flood wall on embankment

30.10.5 Flood walls Where space or foundation conditions do not permit the construction of embankments, flood walls may be used. To meet amenity requirements, the wall may have to be cased in brick, or natural or artificial stone, or a half bank may be formed behind it which will provide support, seal any leakage and overcome objections to an exposed concrete face. Where it is necessary to allow access through a flood wall in residential or industrial areas and steps over the wall are impracticable, flood gates may be incorporated. These are sidehung hinged steel gates with a rubber seal and stout locking device. Normally left open, these gates are closed by the residents when flood levels get dangerously high. Similar flood gates may be provided on vehicular access ways in which case they may be bottom hinged or housed in a slot below road level and raised when necessary.

Provided that an adequate area of suitable land exists in the river valley, the maximum flow passing downstream may be reduced to any desired extent above the normal flow, but the volume of storage required increases in greater proportion than the reduction of residual flow, and economy may require a compromise between the provision of storage and channel improvements downstream (Figure 30.14). In order to spill a substantial part of the flow into the basin when the safe residual flow is reached without a further increase in the flow passing downstream, a long side weir may be used separating the basin from the normal channel, or an automatic sluice of adequate capacity designed to maintain a constant water level on the main channel side may be used. The long side weir may take the form of a low embankment suitably protected against erosion by stone, gabion, concrete block or geotextile reinforcement.

30.11.1 Detention basins The peak discharge in a river or stream can be reduced by storing some of the flow in a detention basin temporarily. The flow into the basin should be controlled by banks and weirs so that flows up to the bank-full capacity can pass down the stream leaving the basin empty. In flood the excess flow can then be spilled into the basin over a weir or through a sluice so that flow down the channel is still restricted to a safe value (Figure 30.13). When the peak of the flood has passed the basin can be emptied through an escape sluice back into the river. The escape sluice may be manual or automatic and should be designed to allow rapid emptying of the basin so that the storage is available in the event of a second flood in quick succession.

Cost

Residual flow Figure 30.14 Combined channel improvement and storage - economic assessment

Low bank

Sluice High gro

30.11.2 Washlands The use of washlands is common in the Fens of the southeast UK and elsewhere. Flood embankments are constructed well back from the river enclosing an area which may be more than 1 km wide and 30 km long. This area acts both as a detention basin and as a flood channel. Large quantities of water can be temporarily stored and because of the large cross-sectional area of the waterway at flood level, velocities and surface gradients are very small. The area of land involved is considerable and its use is usually restricted to summer grazing or the production of hay or lucerne. If the washland is only required for major floods, lesser floods may be excluded by a lower bank alongside the river and the washland may be cultivated. For small schemes uncontrolled flow on to the washlands may be acceptable but large schemes rely on sluices to regulate the flow into and out of the washlands, thus optimizing their use. 30.11.3 Catchwater drains In many cases the water from areas of high ground runs down into low-lying areas to create or accentuate drainage problems. The upland area may be more extensive than the lowland and also produces a greater runoff per unit area. The upland flow may be diverted away from the lowland area by a catchwater drain. In the case of lowland pumped drainage schemes the diversion of the upland water will reduce both the capital cost of the pumping plant and its subsequent running cost. The catchwater drain is located on the edge of the upland and is designed to intercept the streams running down into the lowland area. Control structures are necessary to allow some flow to follow the original course in dry periods to avoid a shortage of water in the lowland area. In flood periods the whole

F low Q

Safe carrying capacity of channel

Time T Figure 30.13 Flood detention basin, (a) layout, (b) estimation of storage capacity

Cheapest scheme

30.11 Detention basins, washlands and catchwater drains

flow may be diverted down the catchwater to a suitable outlet away from the lowland area.

in New Zealand, may be used to give indicative scour depths. The local scour (d&) due to the piers must be added to the general bed scour (D5).
^ = 0.8 VT^

30.12 Structures
30.12.1 Introduction The design of river structures must involve an understanding of the fundamentals of channel hydraulics, sediment transport and the associated problems of scour and deposition. It must be appreciated that any structure in a river is likely to interfere with the natural regime of the channel, the consequences of which can be far-reaching unless the process is fully understood and is catered for in the design. Some of the structures commonly encountered in river engineering are described below. 30.12.2 Retaining walls The riverside retaining wall is used to support the river bank or a road or building adjacent to it and is used where a sloping revetment is inappropriate. The wall should be designed for the rapid drawdown condition where the water table behind the wall remains high when the river level drops quickly after prolonged flooding. Weepholes and gravel drains behind the wall may be provided to relieve the hydraulic pressure, but these must not be assumed to eliminate the pressure differential, only to reduce it. It is normal to provide a cutoff (see Figure 30.12(d)) on the base of the wall. This member serves three functions: (1) It provides a key to improve the factor of safety against sliding. (2) It reduces seepage under the wall when river levels are high. (3) It provides stability in the event of bed erosion in front of the wall. Various alternative walls have been described in section 30.9.3. Steel sheet piling is very common in river work, particularly for wall heights in excess of 3 m. Sheet piling has the considerable advantages of speed of erection and no need for dewatering. 30.12.3 Bridges 30.12.3.1 General Bridges spanning rivers affect channel regime inasmuch as the presence of abutments and piers changes the flow pattern causing local acceleration of flow. The simplest way to avoid problems is for the bridge to span the entire river channel, although this is often uneconomic or impractical for large rivers. Even if the bridge abutments are outside the main river channel it will be necessary to check that the approaches do not unduly restrict floodplain flow. As a guideline, a bridge should be designed to have an open area of not less than 80% of the channel cross-sectional flow area for bank-full conditions, although smaller values may be appropriate where the natural river is very slow-flowing. 30.12.3.2 Scour at bridges Bridge piers and abutments must be designed such that they are not undermined by any scouring of the river bed or banks. Scour may be caused locally by the presence of the piers, or it may be a feature of the natural channel, occurring during floods. The following equations, put forward by Holmes24 for rivers

where ^8 = local scour depth in metres, V= flow velocity in metres per second and b = projected pier width in metres
D5 = (y VK)U A]W

where Ds = depth of scour measured from flood water level in metres, y = rise in water level to flood level upstream of the bridge measured from normal water level in metres, W= waterway width at bridge, no deduction for piers, in metres, A = waterway area at bridge, no deduction for piers, in square metres, V= mean flow velocity upstream, no allowance for scour, in metres and K=^(W/4.$3Q112), maximum value K= 1, Q = discharge in cubic metres per second

30.12.3.3 Afflux at bridges Unless a bridge is designed with very streamlined approach and exit conditions, head losses are likely to be significant during flood flows. The head loss at a bridge may be estimated from the following equation: AL-^OI-P?). (30.17)

where /zL = head loss (drop in water level through bridge) in metres, K1 = flow velocity upstream in metres per second, K2 = flow velocity through bridge in metres per second, g = acceleration due to gravity (9.8!metres per squared second) and C= coefficient depending on the degree of streamlining (1.2 for normal bridges)

30.12.4 Weirs Weirs are designed for one or more of the following roles: (1) (2) (3) (4) (5) Measurement of flow (as part of a hydrological network). Control of channel depth for navigation. Reducing the slope of steep, erosive channels. Creation of ponds to improve river fisheries. Control of rivel level to allow perennial irrigation abstractions.

The design of weirs is discussed in Chapters 5 and 22. For flow measurement the Crump weir is often adopted. This is a robust structure offering minimal obstruction to the passage of sediment and debris. It is modular up to a drowning ratio of 80% and can be used for measurement in any state of drowning if a crest tapping is provided (i.e. if both upstream and crest water levels are measured). For modular flow the equation is: Q= 1.961Vh15 (30.18)

where Q = discharge in cubic metres per second, W= weir width in metres and h = upstream head over the weir in metres Where it is necessary to record both high and low flows accurately it is normal to provide a compound weir. A typical structure might comprise a Crump weir for high flows with a parallel Vee-notch weir for low flows, the crest of the Crump weir being above that of the Vee-notch.

For controlling navigation depths a wide range of weir types have been used. The weir will, of course, incorporate a lock to permit the passage of vessels and often sluice gates are provided to pass flood flows (these effectively reduce the range of water levels between normal and flood flow conditions and they may be essential to avoid flooding upstream). The typical weir comprises a crest and a stilling basin, the design of which is discussed in Chapter 22. Provision must be made to reduce seepage under the weir and to resist uplift on the downstream apron under all flow conditions. Wherever the head loss across the weir is more than about 1 m, or where the foundations comprise permeable soils, a check on underseepage, uplift and exit gradient should be made. All of these can be determined by plotting a flow net using an electrical analogue device or an appropriate computer program. For larger weirs the use of a physical model is recommended to test the design before the prototype is built. This is particularly important where entry and exit flow conditions are not straightforward or where heavy sediment loads are expected. For rivers which are used by migratory fish, weirs should be provided with a fish pass. The most common type is a series of pools connected by small weirs or submerged orifices. The pools should not be less than 3 m long and 2.5 m wide with not less than 1 m depth below the connecting openings. The difference in level between successive pools should not exceed 500 mm. 30.12.5 Gated control structures Gated control structures are required to maintain a design water level or range of water levels throughout the range of flow conditions experienced. For an unregulated weir the water level upstream rises with increasing flow, but the introduction of a water control gate can enable a relatively constant upstream water level to be maintained. In effect, the gates (which have a sill level lower than the weir crest) allow the waterway area to be increased, thus permitting the flow to increase without raising upstream water level. The most common form of water control gate is the vertical lift sluice gate, but radial gates and hinged tilting gates are also used extensively. Figure 30.15 illustrates the three types diagrammatically. 30.12.5.1 Vertical lift gates For small sizes the vertical lift gate may be a standard penstock gate fabricated from cast iron or steel sliding in cast iron or steel frames and normally operated by screwed spindles and simple reduction gearing. Larger vertical lift gates are fabricated from welded steel plates and sections having wheels and guide rollers running on wheel tracks built into recesses in the piers and abutments. Such gates are usually suspended from the gearing by plate link chains or steel wire ropes passing over sprockets or grooved winding drums mounted on the overhead steel superstructure. To decrease the loading on the gearing and so increase the manual speed of operation, the gates are often counterweighted, the counterweights being connected directly to the chains or ropes. Modern vertical gates have wheels, bushed with selflubricating bearing material, bearing on stainless steel axles fixed to the gate structure and having a fairly low coefficient of friction. Opening and closing of vertical lift gates may be by manual means or electric power. When electric power is used the opening and closing of the gates may be made automatic, controlled by the upstream water level usually by incremental movements of the gate as a result of variations in water level of about 100 mm. Operation of the gate under automatic control is purposely made slow with time delays between each incremental

Weir chamber
Outlet

Controlwater

Figure 30.15 (a) Vertical sluice gate (b) Float operated radial gate (c) Tilting gate
movement in order to prevent 'hunting' which can result if the incremental movement causes large changes in upstream water level. 30.12.5.2 Radial gates Radial gates are constructed so that the resultant reaction of the water loading passes through the centre of rotation of the gate and thus there is no component of water load to be handled by the gate-lifting mechanism. The water loading is transmitted to the pivot bearings by gate arms and thence to the concrete work. Unlike vertical lift gates, there is no need for groove recesses in the piers and abutments and the side members are required simply to provide seal bearing surfaces. With automatic-type radial gates the weight of the gate structure is balanced by counterweights on extensions to the arms which support the gate water load. The lower edge of the gate closes on to a sill and is sealed by a strip of rubber or similar material. The ends of the gate have rubber or leather seals sliding on the metal plates set in the abutments. Radial gates are normally operated by electric motors but there is a form of radial gate operated by floats. The float operated radial gate (see Figure 30.15(b)) is a special type of radial gate designed to maintain a constant upstream water level automatically. It can be very useful in remote locations where the power supply is unreliable or non-existent. The floats are located in chambers in the abutments or piers. Water from the river upstream passes over an adjustable weir and into the float chamber, from which it escapes at a constant rate through a valve. If the upstream water level rises, water flows into the float chamber at a greater rate than it escapes, causing the float to rise in response to the increased level. The float acts on the gate arm causing the gate to open, which in turn increases the flow through the structure and lowers the upstream water level. Conversely, if the level falls, flow over the

control weir does not balance the escape of water from the float chamber, and the gate closes until normal level is restored. Again, the design must be such as to avoid hunting but in practice it has been found possible to control the level of a large river within 12 mm of a set level. In major floods the gate rises completely out of the water and there is a possibility that the downstream water level will take control and prevent the gate from closing when it should. This situation can be avoided by proper design of the control arrangements. 30.12.5.3 Tilting gates Radial and vertical gates have the disadvantage that flow takes place under them and floating debris of all kinds collects against them and has to be removed manually. This is avoided by tilting gates which are hinged at the bottom edge and allow the water to pass over them. The gates are lowered by links or chains and may be operated manually or automatically by electricity. The gate when fully open lies flat in a shallow pit formed in the foundation to give maximum discharge, or in some cases still forms a low weir in the lowered position. In rivers carrying coarse sediment there may be difficulty in lowering the gate to bed level, but in practice the water weiring over the gate usually keeps the area immediately downstream clear. This type of sluice has the advantage that any failure of power supply or other operating failure will not allow water levels to fall below the gate level. As an adjustable weir it is particularly suitable for small installations with manual operation. 30.12.6 Tidal outfalls Tidal outfalls are required where drainage channels discharge through a sea wall or tidal embankment. Their function is to allow discharge at low tide but to prevent the tidal water from flowing back into the drainage system during the high tide. Essentially they consist of a culvert through the tidal embankment with a tidal flap or door, which may be at the outer end of the culvert or in a chamber in the embankment (Figure 30.16). Where beach levels near the sea wall are relatively high the culvert may be extended for some distance out on the foreshore and a door on the outer end could be subject to severe wave action. At the wall itself the door can be sheltered by wing walls and breakwaters, and still greater protection is obtained by the use of a chamber in the embankment. The door is usually circular or square, of cast iron, steel, or plastic, hung from the top by double hinges which allow the door to seat freely and to accommodate small obstructions such as weed or sticks on the seat. There should be sufficient space between the bottom of the door and the apron on to which the water falls to avoid debris from being trapped behind the door and preventing closure. There should also be adequate clearance between the sides of the door and the wing walls or sides of the chamber. When the door is in a chamber it must be mounted on the upstream wall and the chamber itself must be built up above high tide level. It is not advisable for the chamber to be sited inland of the tidal embankment as the intervening culvert will be under pressure at high tides. If the door is some distance out on Stoplogs for water retention and maintenance

the foreshore there is the possibility of tidal water breaking into the culvert behind the door, and it is therefore advisable to provide a sluice or penstock capable of shutting off the culvert at the tidal embankment or at the inland end. For small drainage channels discharging into a river a similar arrangement is used. The culvert through the river bank is fitted with a flap gate, which is a small version of the tidal door, generally circular and of diameter 300 mm to 1.2 m. Larger flap gates may be rectangular and can be counterweighted to minimize the head loss required to open the gate. Counterweighted gates would not be used in tidal situations where waves could cause cyclical movements leading to damage.

30.13 Pumping
30.13.1 Single or multiple pumps Pump capacities for land drainage installations are commonly in the range 500 to 20001/s. It is usual to provide some standby capacity in pumping stations but, in the case of pumps which operate infrequently, a single pump may be appropriate for small stations. For larger stations where two or more pumps are required it may be appropriate to omit a standby unit, but where the design capacity is regularly achieved, one standby pump should be provided. Pumps of variable capacity allow flexibility in running but add to capital costs and complicate automatic running. 30.13.2 Motive power Most modern pumping installations are powered by electricity. This provides the considerable advantages of automatic operation and reduced maintenance. In some rural areas failures of supply are not uncommon but these are not often long enough to cause difficulties. Where electrical power is not readily available, particularly in developing countries, direct diesel-driven pumps are normally used. For long-life applications the diesel is of the straight vertical cylinder, slow speed, nonautomotive type and generally requires continuous attendance. 30.13.3 Pumps Land drainage pumps are required to produce large outputs at low heads, generally between 3 and 7 m but occasionally up to 10m. This flow-head characteristic falls into the axial and mixed flow bowl range. For the lower heads the axial-flow pump is used but typically the mixed-flow bowl pump of the vertical spindle configuration is used. A typical pump station layout is shown in Figure 30.17. With axial- and mixed-flow bowl pumps, small head variations have a relatively large effect on the quantity of water pumped. Consequently, careful selection of pump and prime mover is required to cope with all demands. Also, because of the head-flow characteristics, the discharge pipe usually has a submerged termination to provide syphonic recovery. To reduce

Emergency penstock Tidal f l a p door

Cut-off piles

Figure 30.16 Tidal outfall

Control room Syphon breaker

J Pump and motor /


Automatic screen rake

Water level High Water Level Low Water Level

Stop logs Figure 30.17 Typical mixed-flow bowl pump installation


the risk of reverse flow a syphon breaker is installed. This comprises a paddle-operated butterfly valve which is held shut by forward flow and opened to atmosphere during reverse flow, thus breaking the syphon. 30.13.4 Control Unmanned electrically powered pumps are generally stopped and started automatically by preset level sensors located in the suction sump. The level-sensory equipment normally comprises mercury float switches but there is a move to use more sophisticated sensory equipment such as ultrasonics because of their low maintenance, ease with which preset levels can be changed, and because they are more compatible with remote instrumentation and automation. Fully automatic stations are now common in the developed countries. Details of water levels and pump operation can be conveyed automatically by a telemetry system to a central monitoring station. The same system can carry warning alarms to signal pump failure or other maloperation. 30.13.5 Pump station building Vertical spindle axial- and mixed-flow bowl pumps are suspended in sumps close to the back wall which is curved to reduce the risk of swirl and vortices occurring. To maximize on sump

Control electrodes Removable handrail Control panel Electro-submersible pump Removable bar screen Ground level Maximum water level Stop log grooves Maximum water level Flap Valve with chain for manual operation for gravity outlet Stop log grooves

Figure 30.18 Typical electro-submersible pump installation

configuration with minimum civil substructure, particularly for non-standard applications, sump model tests are often carried out prior to design. The main walls of the sump may be of reinforced concrete or steel sheet piling. Incoming channels are normally screened with sloping steel bars to prevent weeds and large items of debris entering the sump. Screens require raking either manually or automatically. Buildings are generally required to give protection against the elements for plant and operating staff, to avoid damage due to vandalism, and to improve the appearance of the station. For remote locations and in developing countries little or no superstructure is provided and equipment is suitably rated for the outside locations. 30.13.6 Other types of pumping installation For rivers carrying a high sediment load conventional channel off-takes can become quickly blocked, so requiring continual dredging. One method which has been employed to overcome this is the use of floating pontoons to carry the pumps, or to carry the suction pipes from land-based pumps. Such installations are also useful where the water level varies to the extent that the water's edge recedes leaving a conventional station dry. The most recent development towards changing the pump type is the electro-submersible pumpset. This is being favoured because it can be located below ground level requiring little or no ground equipment or superstructure. A typical arrangement is shown in Figure 30.18.

References
1 Doorenbos, J. and Pruitt, W. O. (1977) Crop water requirements. Food and Agriculture Organization Irrigation and Drainage Paper No. 24. 2 Blaney, H. F. and Criddle, W. D. (1950) Determining water requirements in irrigated areas from climatological and irrigation data. US Department of Agriculture - SCS TP 96. 3 Dastane, N. G. (1974) Effective rainfall in irrigated agriculture. Food and Agriculture Organization Irrigation and Drainage Paper No. 25. 4 Ayes, R. S. and Wescot, D. W. (1976) Water quality for agriculture. Food and Agriculture Organization Irrigation and Drainage Paper No. 29. 5 Booher, L. J. (1974) Surface irrigation. Food and Agriculture Organization Land and Water Development Series No. 3.

6 Baars, C. (1973) Design of sprinkler installations. Department of Irrigation, Civil Engineering, Agricultural University, Wageningen. 7 Vermeirei, L. and Jobling, G. A. (1980) Localized irrigation, Food and Agriculture Organization Irrigation and Drainage Paper No. 36. 8 Etcheverry, B. A. (1915) Irrigation practice and engineering. McGraw-Hill. 9 Food and Agriculture Organization (1980) Drainage design factors. Food and Agriculture Organization Irrigation and Drainage Paper No. 38. 10 Smedema, L. K. and Rycroft, D. W. (1983) Land drainage. Batsford Academic. 11 Van Beers, W. J. F. (1963) The Auger hole method, ILRI, Wageningen. 12 Van Beers, W. J. F. (1979) Some nomographs for the calculation of drain spacings, ILRI, Bulletin No. 8, Wageningen. 13 United States Bureau of Reclamation (1978) Drainage manual. US Government Printing Office. 14 Ven Te Chow and Yevjevich, V. M. (1964) Statistical and probability applied hydrology. McGraw-Hill. 15 Bilham, E. G. (1962) The classification of heavy falls of rain in short periods. HMSO. 16 US Department of Agriculture (1968) A method of estimating volume and rate of runoff in small watersheds. US Department of Agriculture, Soil Conservation Service. 17 National Environment Research Council (1975) Flood studies report. NERC, London. 18 Sutcliffe, J. V. (1978) Methods of flood estimation - a guide to the Flood studies report. IOH. 19 Nixon, M. (1959) 'A study of the bank-full discharges of rivers in England and Wales'. Proc. Instn. Civ. Engrs, 12, p. 157. 20 White, W. R., Paris, E. and Bettess, R. (1981) Tables for the design of stable alluvial channels. Hydraulics Research Station. 21 Vanoni, V. A. (ed) (1975) Sedimentation engineering. American Society of Civil Engineers, New York. 22 Ven Te Chow (1959) Open channel hydraulics. McGraw-Hill. 23 Royal Society for the Protection of Birds and the Royal Society for Nature Conservation (1984) Rivers and wildlife handbook. 24 Holmes, P. S. (1974) 'Analysis and prediction of scour at railway bridges in New Zealand.' New Zealand Engineering, (Nov) p. 313.

Suggested further reading


Withers, B. and Vipond, S. (1983) Irrigation: design and practice. United Nations (Economic Commission for Asia and the Far East) (1953) River training and bank protection. Brandon, T. W. (ed.) (1987) River engineering, Part 1: 'Design principles'. Institution of Water Engineers and Scientists, London.

Das könnte Ihnen auch gefallen