Hydrology and Sedimentology of Dynamic Rill Networks ...

131
University of Kentucky UKnowledge KWRRI Research Reports Kentucky Water Resources Research Institute 8-1990 Hydrology and Sedimentology of Dynamic Rill Networks Volume I: Erosion Model for Dynamic Rill Networks Digital Object Identifier: hps://doi.org/10.13023/kwrri.rr.178 Daniel E. Storm University of Kentucky Billy J. Barfield University of Kentucky Lindell E. Ormsbee University of Kentucky, [email protected] Right click to open a feedback form in a new tab to let us know how this document benefits you. Follow this and additional works at: hps://uknowledge.uky.edu/kwrri_reports Part of the Hydrology Commons , Sedimentology Commons , and the Soil Science Commons is Report is brought to you for free and open access by the Kentucky Water Resources Research Institute at UKnowledge. It has been accepted for inclusion in KWRRI Research Reports by an authorized administrator of UKnowledge. For more information, please contact [email protected]. Repository Citation Storm, Daniel E.; Barfield, Billy J.; and Ormsbee, Lindell E., "Hydrology and Sedimentology of Dynamic Rill Networks Volume I: Erosion Model for Dynamic Rill Networks" (1990). KWRRI Research Reports. 29. hps://uknowledge.uky.edu/kwrri_reports/29

Transcript of Hydrology and Sedimentology of Dynamic Rill Networks ...

Page 1: Hydrology and Sedimentology of Dynamic Rill Networks ...

University of KentuckyUKnowledge

KWRRI Research Reports Kentucky Water Resources Research Institute

8-1990

Hydrology and Sedimentology of Dynamic RillNetworks Volume I: Erosion Model for DynamicRill NetworksDigital Object Identifier: https://doi.org/10.13023/kwrri.rr.178

Daniel E. StormUniversity of Kentucky

Billy J. BarfieldUniversity of Kentucky

Lindell E. OrmsbeeUniversity of Kentucky, [email protected]

Right click to open a feedback form in a new tab to let us know how this document benefits you.

Follow this and additional works at: https://uknowledge.uky.edu/kwrri_reports

Part of the Hydrology Commons, Sedimentology Commons, and the Soil Science Commons

This Report is brought to you for free and open access by the Kentucky Water Resources Research Institute at UKnowledge. It has been accepted forinclusion in KWRRI Research Reports by an authorized administrator of UKnowledge. For more information, please [email protected].

Repository CitationStorm, Daniel E.; Barfield, Billy J.; and Ormsbee, Lindell E., "Hydrology and Sedimentology of Dynamic Rill Networks Volume I:Erosion Model for Dynamic Rill Networks" (1990). KWRRI Research Reports. 29.https://uknowledge.uky.edu/kwrri_reports/29

Page 2: Hydrology and Sedimentology of Dynamic Rill Networks ...

Research Report No. 178

FINAL REPORT

HYDROLOGY AND SEDIMENTOLOGY OF DYNAMIC RILL NETWORKS

VOLUME I:

EROSION MODEL FOR DYNAMIC RILL NETWORKS

PART A - INTRODUCTION AND OVERVIEW

PART B - EROSION MODEL DEVELOPMENT

Daniel E. Storm, Billy J. Barfield.

Department of Agricultural Engineering University of Kentucky Lexington, KY 40546

and

Lindell E. Ormsbee

Department of Civil Engineering University of Kentucky Lexington, KY 40506

Final Repon from University of Kentucky to U.S. Geological Survey Matching Grant No. 14-08-001-G 1147 entitled "Hydrology and Sedimentology of a Dynamic Rill Network"

Project Period I 0/1/1985 • 2{28/1990

August 1990

Page 3: Hydrology and Sedimentology of Dynamic Rill Networks ...

FORWARD

The contents of this report were developed under a grant from the Department of the Interior, U.S.

Geological Survey. However, those contents do not necessarily represent the policy of the agency, and

should not assume endonement by the Federal Government

Page 4: Hydrology and Sedimentology of Dynamic Rill Networks ...

iii

1.6.2 Rill Growth and Development . . . • . . • . • • • • . • . . . . . . . . . • . . . . . . 33 1.6.2.1 Rill Incisioo . • • • . . • . . . . • . . • • . . . . . . . . • . . . . . . . . . 33

1.6.2.1.1 Boundary Shear Stress Distribution . . . . . . . . . . . . . . . 33 1.6.2.1.2 Rill Geometry • • . . • . . . . . . • • . • . . . . . . . . . . . . 35 1.6.2.1.3 Critical Shear Stress. . . . . . . . • • . . . . . • . . . . . . . . 36

1.6.2.2 Sidewall Sloughing . . . • . . . . . . . . . . . • . . . . . • . . • . • . . . 38 1.6.2.3 Headwall Cuuing . . . . • . • . . . . . . . . . • . . . . . • . . • . • . . . 38

1.6.3 Rill Erosioo Models . . . . . . . . . . • . . . . . . • . . . . . . . . • . . . . . • . . 39 1.6.3.1 Foster and Lane . . . . . . . . . . . . . • . . . • . . • . . . . . . . • . . 39 1.6.3.2 MIRE . . . . • • . . • . . . . . . • . . . . . . . • . • • . . . . . . . . . . 39 1.6.3.3 KYERMO . • . • • . • . . . . . . • . . . . . . . • . . • . . . . . . . . . . 40 1.6.3.4 Mossaacl and Wu • . • . . • . . . • . . . . • . . . . . • . . • • . . . . . . 40 1.6.3.5 CREAMS . • . • • . . . . • . . . . . • . . . . . . . . . . . . • . • . . . . 40 1.6.3.6 WEPP . . • . . • . . . . • . • . . . • . . . . . . . . . • . . • . . . . . . 41 1.6.3. 7 Additional Rill Models . . . • . . . . • . . . • . • . . . . . . . . . . • . 41

1.6.4 Cooclusions . . . . . . . . . . . . . . . . . . . . . . . • . . . . . . . . . . . . . . . 42 1.7 Sediment Transport Models . . . . . . . . . . . . . • . . . . . . . . . . . . . . . . . . . . . . 42

1.7.1 Yalin Bed Load Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 1.7.2 Yang Tola! Load Equatioo . . . . . . . . . . . . . . . . . . . . . . • . . . . . . . . 44 1.7.3 Simplified WEPP Transport Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 1.7.4 Transport Equations Applicability and Evaluation . . . . . . . . . . . . . . . . . . . 46 1. 7 .5 Conclusions . • . . • . • . . • . . . . • . . . . . . . . . . . . . . . . . . . . . . . . 47

1.8 Particle Size Distribution of Eroded Sediment . . . . . • . . . . . . . . . . . . . . . . . . . . . 47 1.8.1 Deposition Effects . . . . . . . . . . . . . . • . . . . . . . . . . . . . . . . . . . . . 49 1.8.2 Methods to Estimate Eroded Particle Size Distribution . . . . . . . . . . . . . . . . . 49

1.8.2.1 Experimenlai Procedure Using a Rainfall Simulator . . . . . . . . . . . . 49 1.8.2.2 Use of Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

1.8.3 Cooclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

CHAPTER 2 DYNAMIC RILL MODEL OVERVIEW 54

CHAPTER 3 RANDOM SURFACE GENERATOR . . • . . . . • . . . . . . . . • . . • . . . . . 54

3.1 Tmning Bands Method • . . . . . . . . . . • . . . . . . . . . . . . . . . . • . . . . . . . . . • 54 3.2 Spectral Line Generation. . . . • . • . . • . • . . . . . . • . . . . . . • . . . . . . . . . . . . . 58 3.3 TUBA . . • • . . • . • . • . . . . . . . • . • • . . . . . . . . . • . . . . . . . . . . . . . . . . 59 3.4 Summary and Conclusions . • • . • . . • . . • . . . . . . . . . . . . . . . . . . . . . . . . . . 60

CHAPTER 4 RILL NE1WORK GENERATION . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.1 Flow Direction Computation . • . . • . . • . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.2 Filling Topographic Dep~ions. . . • . . • . . . . • . . . . . . . . • . . . . . . . . . . . . . . 62 4.3 Flow Accumulation and Rill Network . . • . . • . . . . . . . . . . . . . . . . . . . . . . . . . 62 4.4 Subwatershed Delineation . . . • . • • . • . . . . . . . • . . • . . . . . . . . . . . • . • . . . 64 4.5 Overland Flow Length . . . . . • . . . . . . . . . . . . . . . . . . . . . • . . . . . . . . • . • 64 4.6 Summary and Conclusions • . . • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . • 68

CHAPTER 5 RUNOFF MODEL ................................... CHAPTER 6 SOIL EROSION MODEL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..

69

70

6.1 Sediment Routing . . . . . . . . . • . . . . . . . . . . . . . . . . • . . . . . . . . . . . . . • . 70 6.2 Interrill Erosion . • . • • • . • • . • . . . . . . . . . . . . . . . . • • . . . . . . . . . . . . . . 71 6.3 Rill Erosion . . . • . . . • . • • . . . . . • . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

6.3.1 Foster and Lane Model Development . . • • . . . . • . . . . . . . . . . . . . . . . . 73 6.3.1.1 State 1 Development: Erodible Channel Bottom . . . . . . . . . . . • . . . 75

6.3.1.1.1 Detachment Rate . . • . . . . . . . . . . • . . . . . • . . . . . 75 6.3.1.1.2 Shear Stress Distribution • . . . . . . . . . . . . . . . . . . . . 75 6.3.1.1.3 Equilibrium Rill Characteristics . . . . . . • . . . . . . . • . • 76 6.3.1.1.4 Rill Equilibrium Conditions . . • . . . . . . . . . . . . . . . . 81

Page 5: Hydrology and Sedimentology of Dynamic Rill Networks ...

6.3.1.2 Stage 2 Development: Nonerodible Channel Bottom . . . . . . . . . .• 6.3.2 Transitional Flow . . • . • . . . • • . . . . • . • • . . • • . . • . . . . . . . . . . . 6.3.3 Sediment Load and Transport Capacity Interactions . . • • . . • . . . . . . . . . . •

6.3.3.l Delachment.~tin&~ .... : ..................... . 6.3.3.2 Transport Lumung : DepoSibon Effects • . . . . . . . . . . . . . . .

6.3.4 Transport Capacity . . • • • • . • . • • . . • . • • . . • . . . . . . . . . . . . . .. 6.4 Erosion Model Swnmary . . . . . . . • . • . . . . . . . • . . . . . . . . . . . . . . . . . . . 6.5 Smnmary and Conclusions • . . . . . • . • . . . . . . . • • . . . . . . . . . . . . . . . . ..

iv

85 89 95 95 98

105 109 110

?MODEL VALIDATION • . . . • . . • . . . . . . . . . . . • • . . . . . . . . . . . . . . . . . . 111

7.1 Field Data Collection • . . . . • . . . . . . . . . • . . . . . . . . . . . . • . . . . . . 111 7 .2 Model Testing and Sensitivity Analysis . . . . . . . • . . . . . • . . . . . • . . . . . . 112

REFERENCES • • • • • • • • • • • • • • . • • • . • . • • . • • • • • • • • • • . • . • • . . • . . . 113

Page 6: Hydrology and Sedimentology of Dynamic Rill Networks ...

LIST OF FIGURES

PARTB

Figure 1.1 Interactions between sediment load, rill detachment and deposition, and flow uansport capacity (Foster and Meyer, 1975) ......••..............

Figure 1.2 Con~=:~gf. ~':'~.3'.1~ ~~~~-~i~ ~~I-~:.~~~ ...... . Figure 1.3 Relationship between time of rainfall exposure and soil Joss rate from a

splash erosion cup (Kinnell, 1974 ). . . • . . . . . . . . . . . . • . . . . . . . . . .

Figure 1.4 Channel network development of a sloping land surface, where (a) is the initial channels (b) is lateral channel development, and (c) is the resulting channel network (after Morisawa, 1985) ....................... .

Figure 1.5 Rill development by cross-grading (Horton, 1945). . . . . . . . . . . . . . . . . . . . . .

Figure 1.6 Channel ordering using the Strahler (1952) method (after Morisawa, 1985).

Figure 1.7 Channel boundary shear stress distribution calculations: (a) orthogonals to isovels (dashed lines) and (b) Lundgren and Jonsson method

v

14

17

20

27

29

30

(Lundgren and Jonsson, 1964}. . . . . . . . . . . • . . • . . . . . . . . . . . . . . 34

Figure 1.8 Shields diagram extended by Mantz (1977).. . . . . . . . . . . . . . . . . . . . . . . . . 37

Figure 2.1 Schematic Diagram of the Component Processes for DYRT. . . . . . . . . . . . . . . . 55

Figure 3.1 Discrete Random Field Generation Using the Turning Bands Method.. . . . • . . . . . . 59

Figure 4.1 Example Showing the Flow Direction Computation Used in the GIS Based Routines.. . . 65

Figure 4.2 Rill Network Delineation Example. . . . . • . . . . . . . . . . . . . . . . . . . . . . . . 67

Figure 4.3 Rill Network Delineation Numbering Scheme. . . . . • . . . . . . . . . . . . . . . . . . 68

Figure 6.1 Discretization of Rill Section Flowing Downhill. . . . . . . . . . . . . . . . . . . . . . . 74

Figure 6.2 Soil clay ~t effi:c~ on interrill erosion exponent b for several soils and cropping condiuons.. . . . • . . . . . . . . . . . • . . . . . . . . . . . • . . . . . 76

Figure 6.3 Erosion vectorS along a rill boundary. . . • . . . . • . . . . . . . • . . . . . . . . . . . 80

Figure 6.4 Normalized rill equilibrium characteristics numerical approximation description. . . . . . 81

Figure 6.5 Normalized equilibrium characteristics for an eroding rill. . . . . . . . . . . . . . . . . . 82

Figure 6.6 Conveyance function g(x•c) for an eroding channel at equilibrium. . . . . . . . . . . . . 83

Figure 6.7 Equilibrium rill geometry for stage 1 rill development . . • . . . . . • . . . . . • . . . . 86

Figure 6.8 Equilibrium rill geometry for stage 2 rill development . . . . . . . . . . . . . . . . . . . 87

Figure 6.9 Rill geometry changes due to flow rate changes. . . . . . . . • . . . . . . . . . . . . . . 92

Figure 6.10 Rill geometry representation showing (a) a two tier rill, and (b) the reduction of the three tier geometry to a two tier system. . • . • . . • . • • . . . . . . . . . . 93

Page 7: Hydrology and Sedimentology of Dynamic Rill Networks ...

vi

Figure 6.11 Multiple rill tier reduction resulting from five decreasing flow rates. . . . . . . . . . . . 94

Figure 6.12 Resulting rill geomeuy for calculating shear excess from a two tier system the increasing flow rate. • • . . . . . . . • . . . . . . . . . • . . . . . . . . . . . 96

Figure 6.13 Sediment trapping efficiency for a rectangular reservoir with steady state unifonn flow (Camp, 1946; Dobbins, 1950). . . . • . . • . . . . . • . . . . . . . . 99

Figure 6.14 Discrete pirticle settling in a rill for (a) two particles and (b) for a unifonn sediment concenttation. . . • . . • . • • . • . . . . • . . • . . . . . . . . . . . . 101

Figure 6.15 Fraction o~ ~led particles for fully IUrbulent flow and ideal quiescent flow condibons. • • • • • • • • • • • • • • • • • . • • • • • • • • • • . • . . . • • • • • . 104

Figure 6.16 Comparison of deposition equations. . . . . . . • . • . . . . . . . . . . . . • . . . . . 105

Page 8: Hydrology and Sedimentology of Dynamic Rill Networks ...

vii

LIST OFT ABLES

PARTB

Table 1.1 Slope lenglh exponents for a range of slopes and rill/inlerrill erosion classes. 8

Page 9: Hydrology and Sedimentology of Dynamic Rill Networks ...

PART A

INTRODUCTION AND OVERVIEW

Page 10: Hydrology and Sedimentology of Dynamic Rill Networks ...

1

CHAPTER!

INTRODUCTION

Soil erosion results from the erosive powers of rainfall and surface runoff. Soil is detached by raindrop

impact and by the shearing forces of surface runoff, and may either be transported off-sire or deposiled on

the field Detached soil is first transported to channelized flow areas known as rills, and then to larger

streams and gullies. In the rills, further detachment can result from the shearing forces of flow, sidewall

sloughing, and headwall advancement The efficiency of the sysu:m to transport soil is a function of the

developing rill network, which is rela1ed to the surface micro-relief and soil properties. Since sediment

yield is a function of the efficiency of the rill network to remove and transport soil, it too must depend on

the surface micro-relief and soil properties.

Recent models have at1emp1ed to include the effects of micro-relief into predictions of soil erosion and

sediment yield, primarily by including random roughness of the surface. Examples include the WEPP

model (Fosler, et al., 1989) and the new Revised Universal Soil Loss Equation (RUSLE) (Renard, et. al,

1990). Although this is an improvement over previous models, random roughness cannot define the rill

network. Knowledge of the spatial correlation of roughness elements is also needed to evaluate impacts on

channelized flow and rill network development Unfortuna1ely, such information is not presently available.

When a new tillage system is developed, one can easily deu:rmine the resulting condition of the soil

surface subsequent to tillage. However, in the absence of information on the impact of a specific surface

condition on flow networks, it is not possible to theoretically evaluau: the impact of the new tillage sysu:m

on erosion. To overcome this knowledge deficit, empirical sbldies are required to develop erosion parame­

ters, which are used to predict erosion and sediment yield impacts. However, it is not possible to theoreti­

cally de1ermine the optimum surface micro-relief which minimizes erosion.

In order to efficiently evaluate new tillage sysu:ms, a sound theoretical procedure for predicting the

erosion for any proposed tillage sysu:m prior to implement development is required. Such a model should

not require field plot data, but require only laboratory deu:rmined parameu:rs. A first logical su:p in the

development of this type of model is the prediction of a bare fallow .soil with a random surface. After theo­

ry is developed for these conditions, other factors such as vegetation and preferential tillage directions may

be incorporated.

Page 11: Hydrology and Sedimentology of Dynamic Rill Networks ...

2

The overall objective of the research discussed in Ibis report is ID develop such a model. The approach

described is a combination of deterministic modeling of runoff and erosion applied ID stochastically genec­

aled swfaces with a defined random roughness and spectrum of roughness elements. Flow netwolks are

defined with a digital elewtion model applied ID generaled surfaces. Average erosion and runoff from a

specified surface is found using many surface micro-relief.realizations.

The report is divided into four volumes as described below:

Vol. 1 Erosion Model for Dynamic Rill Networks

Vol. 2 Hydrologic Model for Dynamic Rill Netwolks

Vol. 3. Simulation of Random Rill Network Geometries on Agricultural Plots

Vol. 4. Dynamic Erosion Model Validation

Page 12: Hydrology and Sedimentology of Dynamic Rill Networks ...

CHAPTER!

LITERATURE REVIEW

1.1 EROSION CLASSIFICATION

3

Soil erosion is typically classified as interrill (sheet), rill, and gully erosion. Intenill erosion is concep­

malized as the uniform removal of soil resulting from raindrop impact and overland flow. This conceptual­

i7.ed form of erosion rarely, if ever, occurs over any sizable area. During the initial development of runoff

very small concentrated flow regions develop, which transpott detached soil particles. These concentrated

flow regions are normally not detected by the naked eye. A majority of soil detached in interrill areas

results from raindrops impacting the soil surface. These detached soil particles are then transpotted to fully

developed concentrated flow regions called micro-rills, and then to rills.

For this research, micro-rills are defined as concentrated flow regions that do not contain incised rills.

Rills are typically several centimeters wide, and are completely obliterated by normal tillage practices.

Enonnous amounts of soil can be detached and transported. Therefore, rill erosion may be the predominant

erosion mechanism for short duration high intensity stonn events. Predominant processes in rill erosion are

head-cut advancement, side wall sloughing, and the detachment of soil particles along the rill boundary due

to excessive shear forces from flowing water.

A gully is a rill in an advanced stage, and can not be obliterated with normal tillage practices. Gullies

fonn in sloping topographic depressions, and once established are considered permanent unless filled or

graded with heavy equipmenL Gully erosion occurs by the same processes as rill erosion, however, head­

cut advancement and side wall sloughing may conuibute higher proportions due to the size of the channel

A new classification or erosion, called ephemeral gully erosion, has recently been established (Foster,

1986). Ephemeral gully erosion occurs in a field's major natural waterways, with channels that are typically

larger than rills and smaller than established gullies. Ephemeral gullies, therefore, may be obliterated by

normal tillage operations, but recur in the same location (Foster, 1986). Ephemeral gullies occur as a result

of the macro-topography whereas rills occur as a result of the micro-topography generated by tillage.

1.2 RAINFALL CHARACTERISTICS

Rainfall energy and runoff shear forces are the primary energy sources for the erosion processes, and

thus the importance of rainfall must never be underestimated. Raindrop impact is the principal energy

Page 13: Hydrology and Sedimentology of Dynamic Rill Networks ...

4

source of detached soil for interrill erosion. Runoff energy resulting from overland flow provides the trans­

port mechanism for detached interrill soil This overland flow moves into concentrated flow regions which

inevitably fonn rills and gullies.

Temporal and spatial distribution of rainfall significantly affects erosion. The temporal distribution of

rainfall affects runoff rate (Ward et al., 1980) as well as total runoff volume, both of which have a signifi­

cant impact on sediment yield. The spatial distribution of rainfall affects the location and intensity of sur­

face runoff, which in tum affects sediment yield However, spatial distribution would only be significant

on large plots or watmheds.

Variations in the drop diameter of natural rainfall also bas an impact on erosion processes. Natural

rainfall contains a distribution of drop sizes. The larger the raindrop, the higher its kinetic energy and

momentum; hence, raindrop size affects the quantity of soil particles detached by raindrop impacL Many

studies have found that the average raindrop size increases with increasing rainfall intensity (Laws and Par­

sons, 1943; Caner et al., 1974).

Rainfall kinetic energy is an integral part of many erosion models. Numerous relationships relating

kinetic energy and rainfall intensity have been developed. The most widely used relationship, developed

from Laws and Parson (1943) Washington D.C. data, is given by Wischmeier and Smith (1958) as:

E = 11.9 + 8.73 logu/ [I.I]

where Eis average kinetic energy (J/m2 /mm of rainfall), and i is rainfall intensity (mm/b).

1.3 EMPIRICAL EROSION MODELS

Understanding of the erosion processes bas been significantly enhanced by empirical models. During

the initial concepbl81izalions of the erosion processes, it was necessary to evaluate the impact of a number

of variables on soil erosion. Empirical models allowed this evaluation with minimal difficulty. However,

empirical models are not physically based, and yield an output from a given input without regard to the

physical mechanisms governing the process. Several of the more widely used empirical erosion models are

described in the following sections.

1.3.1 Universal Soil Loss Equation (USLE)

One of the most widely used empirical erosion models is the Universal Soil Loss Equation (USLE)

(WJScbmeier and Smith, 1978). The USLE is typically used for evaluating the effectiveness of conservation

Page 14: Hydrology and Sedimentology of Dynamic Rill Networks ...

5

practices for farm management plans to maintain soil productivity and minimize off-farm impacts on water

quality. The USLE is given as:

A=RKLSCP (1.2]

where A is the average annual gross soil loss (Mg/ha), R is a rainfall and runoff erosivity factor (El uniW

yr), K is a soil erodibility factor (Mg/ha/EI units), LS is a dimensionless length-slope factor, and CP is a

dimensionless cover-practice factor. LS and CP factors are !wed on deviation from a standard plot 22.1 m

long on a 9 percent slope in continuous fallow with up and downhill tillage. By definition, the LS and CP

factors are 1.0 for a standard plot, and are calculated as the ratio of soil loss on a nonstandard plot to that of

a standard plot. Ero.wn from rill and interrill sources are estimated but gully and ephemeral gully erosion

as well as deposition are neglected. The USLE may be used to estimate average annual, monthly, or single

storm erosion.

1.3.1.1 Rainfall-Runoff Erosivity R Factor

The USLE estimales the combined contributions of rill and interrill erosion, hence the R factor must

account for the erosive po1ential of falling raindrops and surface runoff. For an individual storm event,

WJSChmeier (1959) found that the best erosivity indicator of raindrop impact and surface runoff was the

product of rainfall kinetic energy, E, and the maximum thirty-minu1e inlensity, 130 • The R factor may be

defined as:

E/30 R=-

100 [1.3]

Rainfall inlenSity may be used to predict E, as shown in Equation 1.3. Other relationships predicting natural

rainfall energy were summari7.ed by Hirschi et al. (1983).

Average R factors for synthetic storms may also be estimaled for Type I and Type II storms using

equations developed by Hoies et al. (1973), which are given as:

E/30 p22 Type I: Average Storm lOO = 15 Do.fl¥>S (1.4]

El p22 Type ll: Average Storm

0030 = 19.25 -

l DOMm [1.5]

Equations 1.4 and 1.5 are regression equations developed from depth-duration diagrams. However, a sig­

nificant amount of scatter is present in the data, and therefore these equations only estimate average stonn

RfactorS.

Page 15: Hydrology and Sedimentology of Dynamic Rill Networks ...

6

For computing average monthly R values, Wischmeier and Smith (1965) divided the United States into

geographic regions, and tabulated 10-day R factors for each area using actual rainfa1J data. Average annual

R factors may be obtained from isoerodent maps (Wischmeier and Smith, 1978) or calculated for synthetic

Type I and Type II stonns from a specified 2-year 6-hour rainfa1J event (Hates et al., 1973). It should be

noted that these equations give an average R factor. The R factor at a specific location and time may vary

appreciably, causing significant differences in erosion estimates.

l.JJ.2 SoU Erodibility K Factor

The soil erodlbility factor, K, is quantitatively evaluated from experimental data for a specific soil. The

K factor represents a soil's resistance to erosion. varying with a soil's inf"tltration capacity and intrinsic

propeities that resist detachment The K factor, measured with respect to a standard plot described earlier,

is defined as K=(A/R), , where the subscript s indicates a standard tillage plot (Wischmeier and Smith,

1978).

K values reported in the literanue are avemg~ values for a given soil, and thus are the combined effects

of many soil properties. Extreme care must be taken when selecting a K factor for a specific application

since they have been shown to vary with soil depth as well as soil conditions (Meyer and Romkens, 1976).

Many attempts have been made to relate the K factor to measmable soil properties. Wischmeier et al.

(1971) developed a soil erodibility nomogmph for topsoil and low clay subsoils, estimating the K factor

based on soil organic matter content, permeability, soil structure, and particle si7.e distribution. Experimen­

tal data show that the K factor for many soils and conditions de¥iate from the nomograph. The nomograph

is not applicable to aggregated high clay subsoils (Romlcens et al., 1975; Romkens et al., 1977). In addi­

tion, K factors for subsoils and shale spoil material do not follow the nomograph well for reconstructed

soils (Barfield et al., 1984). Sowces for the latest available K factor estimates are local and state Soil Con­

servation Service (SCS) offices where K values for specific soil conditions at local sites are maintained.

1.3.1.3 Length-Slope (LS) Factors

The length and slope factors, L and S, respectively, have been experimentally evaluated separately.

However, when applying the USLE to field conditions, they have been combined into a single LS factor.

The LS factor is determined experimentally for sediment yield for an arbitrary slope and length, divided by

the sediment yield from standard plots in the same soil. For slopes between 3 and 20 percent, and lengths

up to 400 feet, the original LS factor was estimated by Wischmeier and Smith (1965) as:

Page 16: Hydrology and Sedimentology of Dynamic Rill Networks ...

7

LS= [2,_]'" 430sin20 + 30sin0 + 0.43

72.6 6.613 [1.6]

where A. is the slope length (ft), 0 is the slope angle, and m is a slope dependent exponenL Slope length is

defined as the distance from the initiation of overland flow ID a defined channel, or until the slope decreases

resulting in significant deposition (Smith and Wischmeier, 1957).

McCool et al. (1989) developed new relationships for the slope dependent exponent, m, in Equation

1.6 for 3 conditions. For areas where both interrill and rill erosion occur, Table 1.1 gives recommended val­

ues for varying slopes and rill ID interrill ratios. For thawing soils that do not bave significant interrill ero­

sion, an exponent of 0.5 is recommended. For the last condition when significant deposition occurs, an

exponent of zero is proposed. In addition, McCool et al. (1989) warns that if the USLE is applied ID short

slope lengths, a minimum slope length of 4 m is recommended.

It is important ID note that Equation 1.6 is only applicable ID uniform slopes. Foster and Wischmeier

(I 974) found that convex slopes bave the highest average erosion rates due ID steep base slopes, resulting in

high soil detachment and transport. Conversely, concave slopes have the lowest average erosion rate due ID

the potential sediment deposition on the lower base slopes. For nonuniform slopes, Foster and Wischmeier

(1974) present a technique for estimating erosion by subdividing the slope inlD segments.

1.3.1.4 Co-,er-Practice (CP) Factors

The CP faclDr is the ratio of sediment yield from a plot with specified cover and conservation practice

ID that of a standard plot in continuous fallow. The cover and management faclOr, C, accounts for the

effects of tillage, cover, root density, crop rotation, surface roughness, compaction, consolidation and their

interactions. The C factor varies temporally and spatially. Temporal variation is accounted by time weigh­

ing, and spatial variation is typically adjusted by area weighing. C faclOrS affected by tillage practice and

crop rotation are presented in Wischmeier and Smith (1978). Barfield et al. (1988) discuss the effects on C

of stockpiling reconstructed soils, as affected by consolidation and compaction.

The practice fac10r, P, accounts for the effects of conlDUring, conlDur strip cropping, and terracing. P

values estimates may be obtained from Wischmeier and Smith (1978).

1.3.2 Modified Universal Soil Loss Equation (MUSLE)

Page 17: Hydrology and Sedimentology of Dynamic Rill Networks ...

8

Table 1.1. Slope length exponents for a range of slopes and interrill erosion classes (McCool et al., 1989).

---- ------------------------Slope Length Exponent, m

---- ------RiJIIInterrill Ratio, p

Slope Steepness I.owl Moderate2 Higb3

------- ------------0.2 0.02 0.04 0.07

0.5 0.04 0.08 0.16

1.0 0.08 0.15 0.26

2.0 0.14 0.24 0.39

3.0 0.18 0.31 0.47

4.0 0.22 0.36 0.53

5.0 0.25 0.40 0.57

6.0 0.28 0.43 0.60

8.0 0.32 0.48 0.65

10.0 0.35 0.52 0.68

12.0 0.37 0.55 0.71

14.0 0.40 0.57 0.72

16.0 0.41 0.59 0.74

20.0 0.44 0.61 0.76

25.0 0.47 0.64 0.78

30.0 0.49 0.66 0.79

40.0 0.52 0.68 0.81

50.0 0.54 0.70 0.82

60.0 0.55 0.71 0.83

-----------1 Conditions where rill erosion is slight with respect to interrill erosion; generally C factors will be less than 0.15. 2 Conditions where rill and interrill erosion would be about equal on a 22.1 m long slope in seedbed condi­tion on a 9 percent slope. 3 Conditions where rill erosion is great with respect to interrill erosion; generally C factors will be greater than0.70.

Page 18: Hydrology and Sedimentology of Dynamic Rill Networks ...

9

As staled earlier, the USLE estimateS gross erosion and does not account for deposition. Sediment

yield on the other hand, is the amount of sediment reaching a watershed outlet, and, as a result of deposi­

tion, is not typically equal to the sum of gross erosion over the watershed. To account for deposition, a

delivery ratio, D, is defined as D=Y/A, where Y is sediment yield and A is gross erosion. The delivery ratio

is typically less than unity due to sediment deposition, which results when the sediment load exceeds the

transport capacity of the surface runoff. This deficit in iransport capacity can be caused by vegetation,

incorporated residue, increasing surface roughness, decreasing slope, or a lateral sediment influx.

To estimate sediment yield, Williams (1976) proposed modifying the R factor in the USLE to account

for deposition. Using watershed data from Texas and Nebraska, Williams {1976) found that the product

Qq,, gave the best correlation to observed data, where Q is the runoff volume (ml ) and <!p is the peak dis­

charge (ml /s). The term Qq,, is termed the runoff energy, and when substituted into the USLE yields:

(1.7)

where Y is sediment yield {Mg), and K, LS and CP are the area weighted USLE terms. Equation 1.7 is

known as the modified USLE or MUSLE. Based on the development of MUSLE, it should only be applied

on a watershed scale.

1.3.3 Foster, Meyer and Onstad (FMO) Model

Foster et al. {l977a) developed an erosion model that predicts the contribution of rill and interrill ero­

sion separately using pseudo theoretical relationships. Formulated using the steady-state continuity equa­

tion, the Foster, Meyer and Onstad {FMO) model estimateS gross erosion. The model assumes that interrill

detachment is a linear function of slope and rainfall intensity, and rill detachment is proportional to tractive

force raised to the 3/2's power. The FMO model may be given as:

G=G +G. ' '

[I.SJ

or

G=.x2ka(sin0)' FC P +xk.(bsin0+c)ICP. r I r r I I & I

(1.9]

where G is total gross erosion, G, and G, are rill and interrill erosion.respectively, x is overland flow dis­

tance, le, and k; are rill and interrill erodtbilities, C, P, and C; P, are cover-practice factors for rill and inter­

rill areas, respectively, 0 is slope angle, I, is rainfall erosivity, F, is runoff erosivity, and a, b, c and e are

empirical constants. Foster et al. {1977b) suggests:

Page 19: Hydrology and Sedimentology of Dynamic Rill Networks ...

10

I =El ' 30

(1.10]

and

F = 15EI Q ql/3 ' p

[1.ll]

Foster et al. (1977b) estimated the empirical coefficients a. b, c and e by equating the USLE to the FMO

model

1.3.4 CREAMS Field-Size Sediment Model

Foster et al. (1980) proposed separate detachment equations for rill and interrill erosion for the Chemi­

cals, Runoff and Erosion from Agricultural Management Systems (CREAMS) Model. Based on the FMO

model, the CREAMS rill and interrill detachment equations for upland erosion are:

D, = 0.210E/(S + 0.014)KCP[ ~] [l.12]

and

D = 37983 mV a113[~]-'s-KCP[~] ' " P 72.6 V

" [l.13]

where D; and D, are interrill and rill detachment (lb/ft2 /s), respectively, x is down slope distance (ft), sis

the sine of the slope angle, V. is the runoff volume (ft), and a, is the peak runoff rate (ft/s).

Equations 1.12 and 1.13 may be applied under the same conditions as the FMO model. The use of

· these equations give a slight improvement over the use of the USLE (Foster et al., 1977b). However, it is

important to note that these equations estimate the total erosion over an entire storm, and do not represent

detachment rates.

1.3.S Conclusions

Empirical models are limited in that they should only be applied to conditions under which they were

developed. To extrapolale beyond these conditions could result in significant enors. Hence, the use of

empirical model for erosion modeling should be limited to the condition for which the model was devel-

oped.

Page 20: Hydrology and Sedimentology of Dynamic Rill Networks ...

11

The use of empirical erosion models has aided in the identification and understanding of erosion pro­

cesses. As_ specific erosion processes have been identified, empirical equations were developed to describe

the processes and their interaction with other processes. The USLE, MUSLE, FMO and CREAM erosion

models, although empirical, have been a stepping stone for the development of more physically based and

process oriented erosion models.

1.4 FUNDAMENTAL EROSION MODELING

1.4.1 Erosion Principles

The basic processes of soil erosion are detachment, transpon, and deposition. Interrill erosion occurs as

soil particles are detached by raindrop impact, and transported by raindrop splash or in overland flow. In

addition, minor amounts of soil may be detached by overland flow which begins when rainfall intensity

exceeds the infiltration capacity of the soil Hence, an increase in rainfall excess potentially increases trans­

ported sedimenL

Rill erosion occurs as the overland flow converges into concentrated flow regions. The flowing water

result in shear forces on the soil surface, which may detach additional soil. These shear forces increase

with increasing flow rate and slope. Total erosion rate is the sum of the contributions from rill and interrill

erosion.

1.4.1.1 Sediment Continuity Equation

The primary governing equation for fundamental erosion processes is the continuity equation for mass

transporL For upland erosion, Foster (1982) and Bennett (1974) give the continuity equation as: . di/, +p ~=D +D. ilx 'ilt r.

[1.14]

where q. is sediment load (kgjm/s), xis distance down slope (m), p, is mass density of sediment particles

(kg!m3 sediment), C is sediment concentration in the flow (m3 sedimenl/m3 flow), y is flow depth (m), tis

time (s), D, is rill erosion rate (kg sediment/ml /s), and D, is interrill sediment delivery rate to the rills (kg

sediment/ml /s). The iJqjilx term represents the sediment storage rate along the slope, and p,il(cy)/iJI repre­

sents the change in sediment storage over time. Equation 1.14 assumes spatially uniform flow and sediment

concentration, and neglects dispersion and the distribution of sediment storage over the flow depth.

Page 21: Hydrology and Sedimentology of Dynamic Rill Networks ...

12

By assuming relatively shallow and gradually varying flow, the piJ(.cy)/at storage term may be neg­

lected. The resulting quasi-steady state continuity equation has been used extensively (Foster and Huggins,

1977; Foster and Meyer, 1975; Curtis, 1976; Thomas, 1976; Fostec and Meyer, 1972a), and may be written

as:

d</, -=D +D. Ox r '

1.4.1.2 1nreraction Between Sediment Load and Transport Capacity

(1.15]

As stated earlier, Equations 1.14 and 1.15 neglect the interaction between sediment load and transport

capacity of the flow. Meyer and Monke (1965) observed that this asswnption is incorrect and proposed that

the terms q., D, and D, must be considered as dependent variables. Foster and Meyer (1972a, 1975) pro­

posed that rill detachment and deposition are proportional to the difference between transpon capacity and

sediment load. yielding a diffusion type equation:

D=C(T-q) r 1 C I [l.16]

where C1 is a first-order reaction coefficient (m·1 ), and T. is the sediment transpon capacity of the flow

(kg/s/m rill width). Rearranging Equation 1.16 yields (Foster and Meyer, 1972a, 19__?~1=

(1.17]

Next, Fostec and Meyer (1972a) asswned that the maximwn detachment capacity, n .. , was proportional to

the transpon capacity, such thac

D =CT "' I c

Substituting Equation 1.18 into Equation 1.17 yields:

D, q, -+-=1 D T

"' c

(1.18]

[1.19]

Equation 1.19 indicates that when the sediment load. q. , equals zero, rill detachment, D, , equals the

detachment rate capacity, D., Conversely, when D, equals zero, the sediment load equals the transpon

capacity, T •.

These concepts are conceptually justified by the stream power concept (Foster and Meyer, 1972a). Any

specified flow contains a finite amount of energy, which may be used for detaching and/or transporting soil

particles. 1be D, /D., term in Equation 1.19 represents the relative amount of energy expended on soil

Page 22: Hydrology and Sedimentology of Dynamic Rill Networks ...

13

detachment, and the term q. rr. indicates the relative amount of energy expended on sediment transport.

The sum of these two terms equal unity, the total relative available energy.

Equation l.16 may also be applied to deposition. During deposition C varies with sediment size, and to

date there are no validated relationships describing this phenomena. Based on results by Einstein (1968),

Foster and Huggins (1977) and Davis (1978), Foster (1982) proposed that C1 for overland flow may be

approximated by:

v c =-'

I 2(/ [1.20]

where V, is settling velocity (rn/s), and q is flow discharge (m3 /m width). When Equation 1.18 is applied to

channelized flow, Foster (1982) proposed:

v c =_!.. I q

[1.21]

Equation 1.16 may also be explained by Figure 1.1, a schematic illustrating the interactions between sedi-

ment load, rill detachment and deposition, anp flow transport capacity. As shown in Figure 1.1, initially

, sediment free flow enters a region with a nonerodible surface. At the instant the flow encounters the erodi­

,· ble surface, soil is detached at a rate which decrease exponentially with distance until the sediment load

equals the flows transport capacity. When additional sediment is added to the sediment laden flow, deposi­

tion occurs until the sediment load equals the transport capacity of the flow. This simple illustration shows

how the sediment load changes with varying transport capacity. Soil detachment occurred when the trans­

port capacity· exceeded the sediment load, and deposition occurred when the sediment load exceeded the

transport capacity. Hence, detachment capacity is based on localized conditions, but the actual detachment

rate is governed by the available excess transport capacity.

L4.2 Foster and Meyer Closed Form Erosion Equations

Foster and Meyer (1972a, 1975) developed a closed-form erosion model by simultaneously solving

Equations 1.15 and 1.19, the steady-state sediment continuity and a sediment load-transport capacity equa­

tion described in the previous section. The general solution to their erosion model is:

D, -az, [f ( dg. ) ""', J Dec =e dx. -9 exp dx.+c [1.22]

Page 23: Hydrology and Sedimentology of Dynamic Rill Networks ...

SEDIMENT "fRANSPORT RATE

-----· SEDIMENT oETACHMENT RATE

FLOW -~:__ ........ \ EROSION

SEDIMENT ADDITION

14

EROSION DEPOSITION

NONERODIBLE ERODIBLE

_,, _____ ... _ ... _ ... __ _ (FostflC and }d.eytS, 1975).

Page 24: Hydrology and Sedimentology of Dynamic Rill Networks ...

)C )C =-• L

0

LD. 9=-·-·

T..,

15

[l.23]

[l.24]

[l.25]

[l.26]

where x is the distance downslope, L0 is the slope length, D00 and T co are the detachment and transpon

capacities at x=L. , respectively, and c is an integration constant

Equations 1.22 to 1.26 were formulated by assuming uniform shallow flow, and steady-state rainfall

excess, q. , given by:

[l.27]

where q. is rainfall excess at x=L. • In addition, the detachment and transpon rates, D0 and T0 , respective­

ly, were defined as:

and

D =C t3fl c D [l.28]

[l.29]

where t is the average shear stress on the soil surface, and C0 and CT are constants, The remaining vari­

ables, D00 , T .. , and I>; are solved with auxiliary relationships. If uniform slope and constant rainfall excess

are assumed, Equation l.22 may be written as:

qi -az. -1 r=x.-(1-9)(1-e )a [l.30]

"' The general closed form erosion equation is applicable to both detachment and deposition, as well as

several nonunifonn slope conditions (Foster and Meyer, 1975). The erosion equation has been validated

using field data (Foster and Meyer, 1972a) producing acceptable predictions. In addition, the closed fonn

equation was used in CREAMS (Foster et al., 1980) for identifying detachment and deposition of overland

and channelized flow.

Page 25: Hydrology and Sedimentology of Dynamic Rill Networks ...

16

The closed fonn equation may be applied 10 many other applications as well. For example, Foster

(1982) suggests that the efficiency of grass filrer strips in ttapping sediment may be evaluated using the

equation. For many applications the erosion equation requires minimal data, and may be applied 10 relative­

ly simple scenarios 10 evaluate specific erosion processes and interpret field and laboratory experiments. A

disadvantage of the closed form equation is the necessary erodtbility parameters are not available for most

soils. In addition, the model should only be applied 10 conditions that do not violate the underlying assump­

tions. Therefore, the closed fonn erosion equation, although useful for many simple applications, is not

applicable 10 most field conditions.

1.4.3 Meyer and Wischmeier Conceptual Erosion Model

Meyer and WISChmeier (1969) proposed an erosion model with a conceptual framew01k that accounts

for the interactions between the detachment and ttansport of sediment from rainfall and runoff. A schematic

of their framework. is given in Figure 1.2. The model is based on dividing an area in10 segments. Soil is

detached by rainfall impact and surface runoff, and combined with ttansported sediment from an up-slope

segment 10 produce the 10tal detached soil Next. the 10tal ttansport capacity of rainfall and surface runoff is

calculated, and an amount of the 10ta1 detached sediment less than oc equal 10 the 10tal ttansport capacity is

carried downslope.

Meyer and Wischmeier (1969) proposed empirical equations 10 predict each fundamental component.

given by:

D.=C0 AT • • [131)

(132)

(1.33)

(1.34]

where T; and T, and the interrill and rill ttansport capacities (kg soil),respectively, A is slope area (m2 ), I is

rainfall intensity (m/s), S is slope (m/m), Q is flow rate (m3 /s), and Cc; , C0r , Cn and CT, are empirical

constants with appropriate units.

The Meyer and Wischmeier model provides a conceptual framework. for erosion modeling. Each ero­

sion process is accounted for separately and the interaction between total detached sediment and ttansport

capacity for discrete segments is also accounted for.

Page 26: Hydrology and Sedimentology of Dynamic Rill Networks ...

17

l SOIL FROM UPSLOPE

DETACHMENT DETACHMENT TRANSPORT TRANSPORT

BY RAIN BY RUNOFF CAPACITY CAPACITY OF RAIN OF RUNOFF

I

l . TOTAL TOTAL

DETACHED SOIL COMPARE . TRANSPORT CAPACITY

IF DET. < TRANS. IF TRANS. < DET.

,, SOIL CARRIED DOWNSLOPE

Figme 1.2. Conceplual framework of Meyer and Wischmea- erosion model (Meyec and Wiscbrneier,

1969).

Page 27: Hydrology and Sedimentology of Dynamic Rill Networks ...

18

1.4.4 Conclusions

The Foster and Meyer (1975) closed-Conn erosion model was developed using relatiooships that

describe the underlying physical processes. The use of this type of physically based model is a tremendous

advancement over conventional empirical methods. A distinct advantage of physically based models is

that they are much more versatile, and may be applied under conditions where minimal information is

available.

Application of physically based models require the delineation of sub-processes that make up the phys­

ical system. The Meyer and Wischmeier (1969) model was the first attempt to develop a conceptual frame­

work for modeling upland erosion. Although the sub-processes were described by empirical equations,

these relationships may very easily be replaced with physically based models.

l.S INTERRil..L EROSION

Interrill erosion is the detachment and transpon of soil particles to a concentrated flow network which

may consist of micro-rills, rills, ephemeral gullies, or gullies. The sum of splash erosion caused by raindrop

impact, and sheet erosion caused by overland flow is net interrill erosion. These components are discussed

in the following sections, along with the effects of surface sealing and cover effects.

l.S.l Splash Erosion

Nearly all detachment of soil particles in interrill areas (Young and Wiersma, 1973) is a result of the

force of raindrops impacting the soil surface. Detached soil particles are transponed in the splashing rain­

fall as well as by overland flow. Thus, splash erosion from raindrop impact consists of soil splash detach­

ment and transpon as discussed separately in the following sections.

1.5.1.1 Raindrop Splash Detachment

Early wm performed by Free (1960) indicated that the quantity of detached soil in interrill areas was

related to the kinetic energy and momentum of raindrops impacting the soil surface. Quansah (1981) devel­

oped an empirical equation relating splash erosion to total rainfall kinetic energy, but Gilley and Fmknea

(1985) found that kinetic energy times drop circumference gave a better statistical fiL However, Park et al.

(1983) and Rose (1960) found that momentum was a better indicator of soil detachmenL To better evaluate

the actual detachment process, Huang et al. (1982, 1983) used a dynamic equation of linear elasticity to

study the deformation resulting from simulated rainfall impact, finding that a cone shaped depression

Page 28: Hydrology and Sedimentology of Dynamic Rill Networks ...

19

resulted, and that a lateral jet stream was the major mechanism governing detachmenL

Since kinetic energy and momentum measurement of natural rainfall is very difficult, numerous rela­

tionships utilizing more easily measured parameters have been proposed. Al-Dmrah and Bradford (1981,

1982) and Cruse and Larson (1977) found that soil detachment varied with soil shear strength measured by

a ttiaxia1 compression test. Cruse and Francis (1984) found that this soil shear strength was inversely pro­

portional to the soil matrix potential. Therefore, die amount of detached soil from raindrop impact varies

during a storm evenL

Gilley and Finknea (1985) suggested that splash detachment parameterS must be evaluated for individ­

ual soils and rainfall conditions. To estimate these parameters, Ellison (1947) introduced a "standard" cylin­

drical splash-cop approximately 9 cm in diameter, and 5 cm deep. However, !here were many inaccuracies

were associated wilh the splash cop, such as die rim impeding die splashing particles as the soil level

decreased ovec time. Rose (1960) developed a correction for rim effects, and Bisal (1950) constructed an

apparatus to eliminate the rim effects.

As pointed out by Kinnell (1974), in Figure 1.3, !here were other errors as well. Initially, when die

splash-cup was full of soil, die amount of splashed material was die sum actual soil splashed and the soil

washed over the side. This additional.soil being pushed ovec the side caused an initially excessive soil loss.

To separate the wash ovec soil from the splashed soil, Kinnell (1974) constructed a splash cup with 2 con­

centtic cylinders, where the outer cylinder was used to catch die wash off caused by the impacting rain­

drops near the cup rim.

Using similar soil splash-cups, Bubenzer and Jones (1971) proposed an equation for estimating splash

erosion given by:

.bkec d ss=a, pc [1.35]

where ss is the soil splash (g/cm2 ), i is rainfall intensity (cm/br), Ire is die total applied rainfall kinetic ener­

gy (J/cm2 ), pc is die soil percent clay, and a, b, c and dare constants. Quansah (1981) proposed a similar

relalionship, but wilh total applied kinetic energy and slope. Hirschi and Barfield (1988a) suggest including

the slope, s, into Equation 1.35, yielding:

J, c "• ss=a, kepc s [1.36]

where s is slope, and e is a constanL Given sufficient information, Equation 1.36 may be used to estimate

soil splash from areas with various soils, topography; and surface conditions. Othec studies indicated that

splash ecosion may be approximated by rainfall intensity alone (Ellison, 1947; Ekem, 1953, Kinnell, 1976).

Page 29: Hydrology and Sedimentology of Dynamic Rill Networks ...

a. ::::, 0 :!! 0 a: u.. w ~ a:

§ .... i

splash I & I

wash-: over I

uninhibited splash

I I splash : Inter­• ceptlon

RAINFALL EXPOSURE TIME

20

Figure 13. Relationship between time of rainfall cxposwc and soil loss rare from a splash erosion cup

(JCinncD, 1974).

Page 30: Hydrology and Sedimentology of Dynamic Rill Networks ...

21

Another factor significantly affecting splash erosion is surface ponding, which dissipates raindrop

energy. Palmer (1965) found that a maximum soil splash occurred at one drop diameter, and Mutchler

(1967) found that this maximum occurred at 1/3 the drop diameter. Later Mutchler and Larson (1971)

found that the maximum splash occurred at a depth of 1/6 the drop diameter. Shultz et al. (1985) found that

splash rate increased exponentially with time until ponding occurred, and then decreased exponentially

after ponding. Shultz et al. (1985) also found that there was no appreciable splash erosion for their silt

loam soil, after 4 mm of rainfall was ponded. In addition, they found that the concentration of suspended

particles, after 5 minutes of rainfall, was approximately 14 times that found in the raindrop splash, and that

the rate of accumulation of suspended particles decreased as the ponded water depth increased.

Another approach to estimating soil detachment from overland flow is to assume that raindrop impact

pressure is proportional to soil detachmenL Gilley et al. (1985a, 1985b) developed a soil detachment equa­

tion using relationships from Wang and Wc:117.Cl (1970) such that:

[d]l.83

P=0.2p cos20 \r y [l.37]

and

D=Kj' [1.38]

where P is the impact pressure (Pa), p is the density of water (kg/m3 ), 9 is slope angle (rad), V is raindrop

impact velocity (m/s), d is equivalent drop diameter (m), y is overland flow depth (m), D is soil detach­

ment from one raindrop (kg), and K.i is a soil detachment factor (s2 -m). Toe advantages of this approach is

the physical basis of the equation, and the fact that ponding depth is considered.

1.5.1.2 Raindrop Splash Transport

Toe energy of a falling raindrop is distributed between soil detachment and transpon. Detached soil

particles contained in the raindrop splash are transported in all directions. Al-Durrah and Bradford (1982)

found that the splash angle 9, in degrees, is independent of soil material, and could be estimated by:

9 = 40.5 t 4425 [l.39]

' where t is the soil shear strength (kPa). ·

On a level surface, the average net soil transport is zero. However, Ellison (1947) observed that on a

sloping surface more soil was splashed downslope. Ekern (1953) proposed that the downslope fraction of

transported soil could be estimated from:

Page 31: Hydrology and Sedimentology of Dynamic Rill Networks ...

22

% Downslope Movemenl = 50 +%Slope [1.40]

Equation 1.40 is based on slopes between O and 36 percent, and indicates that all splashed soil will be trans­

ported downslope for slopes greater than 50 percent. Fanner and Van Haveren (1971) proposed that the

downhill portion of the splash could be represented by:

SD = sin9 + 0.5cos9 [1.41]

where 9 is the plot slope angle and S0 is the downslope portion. This equation is based on the relative

force directions of a soil particle resting on a sloping surface due to a vertical-falling raindrop. It indicates

that all soil splash will travel downslope for slopes exceeding about 75 percent.

Toe studies mentioned above did not propose ttanspon relationships for splash transport, only transport

distribution. Quansah (1981) proposed transport equations of the same form as the detachment equations

presented in a previous section, such that

Q,=aKE'S [1.42]

where Q, is the net splash transport (kgfm2 ), KE is the total applied kinetic energy (J/m2 ), S is the soil

slope(%) and a, band c are constants. As would be expected from the proposed transport distribution, the

transport slope-exponent is much higher than the detachment slope exponent, indicating that transport is

much more sensitive to slope than detachment.

1.5.2 Sheet Erosion

Sheet erosion is the detachment and transport of soil particles resulting from shear forces of overland

flow. Average shear stress may be estimated from the product of the surface slope and overland flow depth.

Based on the Dan:y-Weisbach equation, overland flow depth may be estimated by a procedure given by

Gilley et al. (1985a). Studies have shown, however, that shallow overland flow produces minimal shear

force, hence, little soil detachment (Foster and Meyer, 1975).

Overland flow is responsible for most of the transport of soil to rills and detachment from raindrop

impact (Young and Wiersma, 1973). The dominant portion of interrill transport capacity is contributed

from raindrop impact (Guy et al., 1986). Raindrops impacting overland flow cause significant amounts of

turbulence, which help keep detached soil particles in suspension for transport.

Most sediment transport equations were developed for stream flow, not overland and rill flow condi­

tions. However, due to the lack of adequate relationships many of these equations have been applied to

Page 32: Hydrology and Sedimentology of Dynamic Rill Networks ...

23

overland flow conditions, with a few producing satisfactory results. The Yalin (1963) equation is the most

widely used transport equation for overland flow (Hirschi and Barfield, 1988a; Hirschi and Barfield,

1988b; Di1laha and Beasley, 1983; Foster et al., 1980). The Yang (1972) unit stream power equation has

also been applied to overland flow (Moore and Burch, 1986; Gilley et al., 1985; Kilinc and Richardson,

1973), however, it yields better results in deep flow conditions. Development and testing of the Yang and

Y alin Equations will be discussed in a subsequent section on sediment transpon models.

1.S.3 Net lnterrill Erosion

Net interrill erosion is the amount of sediment actually delivered to a concentrated flow network,

which is the total detached soil particles less the deposited material. Meyer (l 981) found that interrill ero­

sion delivery, D, may be described by rainfall intensity such that

D=al [1.43]

where I is rainfall intensity, and a and bare constants. Meyer (1981) evaluated the effect of the b exponent

for various soil and cropping conditions, and found tl1at the b exponent increased as the present clay

decreased. For the conditions studied, Meyer (1981) concluded that for soil with clay contents between 20

and 50 percent, the b exponent may be estimated from:

b=2.l-Fcl [1.44]

where F c1 is the fraction of clay. Meyer (l 981) also found that for low<lay soils with clay contents less

than 20 percent, such as silts, silt loams, loams, and sandy loam soils, the net interrill erosion may be found

by:

D=Cf [1.45)

where C is a constant that accounts for varying soils, tillage, and cropping conditions.

The net interrill erosion used in the USDA Water Erosion Prediction Project (WEPP) is based on

Equation 1.45, and given by Nearing et al. as (1989):

R D.=KfCG....!..

I I •• •w [1.46]

where Di is net interrill erosion rate (kg/m2 /s), K;. is baseline interrill erodibility, I. is effective rainfall

intensity, (m/s), c. and o. are dimensionless canopy and ground cover effects on interrill erosion, respec­

tively, R,, is rill spacing (m), and Wis a computed rill width (m). Equation 1.46 is applied only during peri­

ods of rainfall excess. During periods without surface runoff, soil detachment by raindrop impact is

Page 33: Hydrology and Sedimentology of Dynamic Rill Networks ...

24

assumed to remain in the interrill areas. Accounting fa only rainfall excess periods is included in the effec­

tive rainfall intensity, which is given as (Nearing et al., 1989):

r=l.rrdt • t }'

[1.47]

• where I is rainfall intensity (m/s), t is time (s), and t. is the cumulative time during which rainfall rate

exceeds infiltration rate (s). Toe baseline erodibility factor, K; may be estimated from soil properties using

relationships developed by Elliott et al (1988). Toe canopy and ground effects are discussed in the follow­

ing section on cover effects. In order to account for the effects of slope on net interrill erosion, Elliot, et al.

(1989) developed the relationship:

sf= 1.05 - 0.85 e -4, mn(,)

where S1is a fractional slope factor, ands is slope angle (degrees).

[1.48]

Toe use of Equation 1.45 in estimating interrill erosion delivery rate is a reasonable approach for most

conditions. However, actual instantaneous erosion rates during the initiation of surface runoff will not agree

with Equation 1.45. Due to the limited overland flow transport capacity present during that time period.

When compared to the total sediment yield for a specific storm event, this deviation should be minimal.

1.5.4 lnterrill Erosion Cover Effects

"The presence of cover on or above the soil surface has a major impact on erosion. Plant canopies inter­

cept a small amount of rainfall. When close to the soil surface, the plant canopy significantly reduces rain­

drop impact energy. Other soil cover such as mulches and plant residue reduces erosion even further. This

material intercepts the rainfall and prevents soil detachment. In addition, cover on the soil surface reduces

the transport and detachment rates of surface runoff by decreasing overland and concentrated flow veloci­

ties. WEPP uses canopy and ground cover parameters in their interrill erosion equation 1.46 to account for

these effects. Canopy effect is estimated by (Laflen et al., 1985):

--0.34H C =I-Fe • . '

[1.49]

where C.. is a dimensionless canopy parameter affecting interrill erosion, F0 is the fraction of soil protected

by canopy covers, and H. is effective canopy height (m). Ground cover is accounted for by (Nearing et al,

1989):

-2.s, G =e ' • (1.50]

Page 34: Hydrology and Sedimentology of Dynamic Rill Networks ...

25

where G0 is a dimensionless ground cover paramet« affecting inrerrill erosion, and g; is the fraction of

inlerrill surface covered by residue.

The effects of cover on erosion should never be neglected. However, the interactions between plant

canopy, mulches, plant residue and incorporated residue create a very complex system. Many unquantified

interactions exist between the various processes. As a result. simple empirical relationships are used to

account for cover effects.

1.5.S Surface Sealing and Crusting Effects on Erosion

Surface sealing and crusting is the development of a thin compact soil layer on a bare soil surface. This

compacted layer is formed by raindrop compaction and the washing of soil particles into the soil matrix.

Crust formation reduces a soil's infiltration capacity, which in tum increases surface runoff and potential

soil erosion. Edwards and Larson (1969) found that surface sealing reduced cumulative infiltration by 50

percent over a 2 hour period. Additional rainfall excess resulting from reduced infiltration is a potential

energy source for detaching and ttansponing soil panicles.

Moore (1981) showed that infiltration rate is more sensitive to hydraulic conductivity than other

parameters. Brakensiek and Rawls (1983) developed simple equations for estimating the reduced soil satu­

rated hydraulic conductivity for an established crust. based on sand fraction, crust thickness, average wet­

ting front depth, and a cumulative rainfall. Using their results, Rawls et al. (1989) incorporated a similar

model into WEPP.

Epstein and Grant (1967) found that during the initial 10 minutes of simulated rainfall, soil loss rate

increased exponentially. After ten minutes, the soil loss rate decreased sharply and decayed to a uniform

erosion rate. This peaking effect was attributed to crust formation during the initial rainfall. However, Man­

nering and Meyer (1963) found that applying straw mulch at densities greater than one ton per acre pre­

vented surface sealing due to rainfall impacL Surface sealing and crusting have a major impact on infiltra­

tion rate and capacity of unprotected soils. Therefore, surface sealing and crust development should be

considered in erosion modeling when conditions dictate.

1.5.6 Conclusions

The sum of splash erosion caused by rain drop impact. and sheet erosion caused by overland flow is

the net inrerrill erosion. Almost all detached soil panicles in inrerrill areas results from raindrop impacL

Page 35: Hydrology and Sedimentology of Dynamic Rill Networks ...

26

The average net transport of splashed soil particles is zero on a level surface, and increases as the slope

increases. However, splash transport is small and overland flow is responsible f<r most of the transport of

detached soil particles to the rill network.

Due to complex interactions, a high degree of empiricism exists in the description of intemll erosion.

Factors such as plant cover and residue, swface sealing and crusting, and the temporal and spatial distnbu­

tion of soil properties complicate matters further. An increasingly popular method to estimate net intemll

erosion is to use the product of a base line erodibility and the square of the rainfall intensity. This approach

lumps the effects of detachment and transport of both raindrop splash and overland flow, and for most

applications is a reasonable approximation.

L6 RILL EROSION

Rills develop as a result of the concentration of overland flow, serving not only as the vehicle for trans­

port of detached soil particles delivered from interrill areas, but also contributing significant amounts of

eroded material resulting from excessive shear forces produced from concentrated flow. In addition to shear

excess erosion, sidewall sloughing and head cut advancement also contribute detached soil in rills. A com­

bination of these processes are responsible for the initiation and development of rill networks, the density

of which may alS!> have a major impact on sediment yield. The following sections describe the processes

involved in rill growth and rill network development, as well as state-Of-the-art rill erosion models.

1.6.1 Rill Networks

In many instances, development and evolution of rills and rill networks are the most important erosion

processes. In the last several decades, significant progress has been made in the understanding and descrip­

tion of individual rill growth and development However, due to the stochastic nature and complex interac­

tions in the development of rill networks, minimal progress has been made understanding rill networks. In

ordec to more accurately predict erosion, rill networking and erosion processes must be integrated to form

dynamic erosion models.

1.6.1.1 Networlc Development

Figure 1.4 illustrates the initiation and development of a rill network. Leopold et al. (1964) observed

that the initiation of a rill network on a bare soil surface starts with the concentration of overland flow into

a series of parallel rills, governed by the swface micro-relief. Cross-grading and micro-piracy can cause

Page 36: Hydrology and Sedimentology of Dynamic Rill Networks ...

27

the integration of these parallel rills. Horton (1945) describes micro-piracy as the overtopping and oblitera­

tion of intennediary ridges between rills, which destroys the original rill network.

Toe process of cross-grading, illustrated in Figure 1.5, occurs when sufficient flow exists in a rill to

overtop an intermediary ridge, creating a lateral flow component toward a lower rill. Over time this micro­

piracy will erode the rill divide, and may create an increasing resultant component towards a main rill. Toe

development of this resultant component across the general surface gradient is called cross-grading (Hor­

ton, 1945).

Mosley (1974) performed indoor rainfall simulator experiments to study rill development, making the

following observation as the infiltration capacity of the soil was reached.

"Puddles appeared, increased in size, and coalesced until runoff was occurring over the whole surface .. .Runoff collected almost immediately into more or less distant flow lines, with small intervening areas on which sheet flow probably occurred.•

Mosley (1974) also observed that these lines later developed into an integrated dendritic rill network with

occasional coalescence. The resulting equilibrium sediment discharge rate was found to be related to total

rill network length, and the equilibrium sediment concentration was governed by the original soil slope.

1.6.1.2 Geomorphological Channel Properties

The geomorphological description of channel networks has almost exclusively been developed based

' on a WlllerShed scale. Horton (1945) performed the first quantitative channel network study, introducing the

concept of channel ordering. Strahler (1952) proposed minor modifications to the Horton (1945) methodol­

ogy, such that "fingertip" channels are defined as first-order. A second-order channel is formed from the

union of two first-order channels. Ann'" order stream is created by the joining of two (n-1)'" order chan­

nels. However, when channels of unequal order join, the resulting downstream channel remains the highest

order of the previous channels. An example of a channel netwon: ordered using this approach is shown in

Figure 1.6.

Another important contribution from Horton (1945) is the introduction of the "Laws of Drainage Com­

position•, which have been summarized by Morisawa (1985). The laws are developed using the channel

ordering concepts previously described. The two most widely used laws are the law of stream numbers,

and the law of stream lengths. Toe law of stream numbers is given as (Horton, 1945):

N =ft:" • b [1.51]

Page 37: Hydrology and Sedimentology of Dynamic Rill Networks ...

( a )

-···i --· -···i ( b ) ·:::.,

( c )

.. "l ... ..,

. ·1· .. ·1 •(

,··--, ···-j····-j r::::i l·····-. ····-;····-; i:::·-j

,····-, ····-

'. ····-,· i::::=. ;::::=; ; ····-; i··--j . ····-·

28

f·--­vr::::

1····-E::

f. ... _

FiglR 1.4. Cfumncl network development of a sloping land surface. where (a) is the initial channels (b) is

lafaa1 cbanncl development, and (c) is the resulting channel network (after Morisawa, 1985).

Page 38: Hydrology and Sedimentology of Dynamic Rill Networks ...

CROSS - SECTIONS RESULTANT DIRECTION OF OVERLAND FLOW

Figure 1.5. Rill development by cross-grading (Horton, 1945).

29

Page 39: Hydrology and Sedimentology of Dynamic Rill Networks ...

\ ... , \ , .. : ),,-:! : . \ ; ..... .

#,( • .... ' ·: : ... ' : : ... ,,. ·--'{ l .,., J/

• if.·· r .... \. / ~-.. -.,.._ .I\ I /. ..... ••• ; ... . ... - . .( . : .. t .,,,.:'.'._ ... ···, .. .. . ,,.. \ .. ,...._:! I ••• :.C:•.

• •• -... I • ••••• -..· \ ;· .. :\ .. \ r·. ·. ·· .. .. ·.. )-1 ! ""·:,~-~- .· ~ . ., .• i .• •.. ' : . . .......

LEGEND:

............. FIRST ORDER

----- SECOND ORDER ----- THIRD ORDER

FOURTH ORDER

Figure 1.6, Cbaooel ordering using the Strahlrz (1952) method (after Morisawa, 1985).

30

where N,. is the ownber of streams of order u. R.t is the bifurcation ownber, ands is the highest order

Page 40: Hydrology and Sedimentology of Dynamic Rill Networks ...

31

stream in the basin. The bifurcation number is the ratio of the number of streams for a given order u, to the

number of streams of order u+ I. The law of stream lengths is given as (Horton, 1945):

L. = L/;_1 (1.52]

where I.,. is the average stream length of order u, L1 is the average stream length of order I, and RL is the

length ratio. 1be length ratio is defined as the ratio of the average stream length of order u+l, to the aver­

age stream length of order u. On a basin scale, the bifurcation ratio and length ratio range from 3 to 5 and

1.5 to 3.5 (Abrahams, 1984), respectively. Ogunlela et al. (1989) evaluated rill networks from a large-scale

indoor laboratory for two surface roughness conditions and two rainfall intensities. They concluded that for

the conditions studied, the average rill networks may be characterized using the bifurcation ratio, length

ratio, and an area ratio deemed similarly to the length ratio.

1.6.1.3 RUZ Density

Rill density has a major impact on sediment yield, and varies with factors like slope steepness and

length, runoff rate, soil texture, and the presence or absence of rainfall (Meyer and Manke, 1965). Ellison

and Ellison (1947) observed that for highly erodible soils many small rills very close togeth« feed into

larger rills and gullies. These rills and gullies remained approximately the same size from the top to the bot­

tom of the slope, th«efore, indicating transport limiting flow where raindrop detachment and interrill trans­

port are dominant. For relatively low «odibility soils, Ellison and Ellison (1947) observed that thCl'.e was

considerable dislllllCes between rills, and that the rills and gullies increased in cross-section as they pro­

gressed downslope, thus indicating that rill incision and sidewall sloughing are significant soun:es of

detached material. Meyer and Manke (1965) observed that short slope lengths have higher rill densities rel­

ative to longer lengths. In addition, more changes in rill networks occur with increasing surface slope and

runoff rate, and decreasing particle size. These factors may affect sediment yield considerably.

Attempts have been made to model rill density. Li et al. (1980) developed a rill density model for lami­

nar and turbulent flows, expressed on a unit width basis. Their model assumed that all rills were the same

size for a specified distance downslope. In addition, numerous empirical constants are required, which lim­

its its use. Li et al. (1980) also estimated critical ttactive force from Shield's diagram (Shields, 1936) using

a representative particle diameter of 0 114 Foster and Lane (1981) indicated that using 0 114 foc primary parti­

cles making up typical tilled agricultural soils would significantly underestimate critical shear stress for rill

erosion.

Page 41: Hydrology and Sedimentology of Dynamic Rill Networks ...

32

Using the KYERMO model, discussed in a following section, Hirschi and Barfield (1988b) performed

a sensitivity analysis on the number of rills across a plot. Their results showed that sediment yield increased

with increasing rill number Wltil a maximum was reached, after which sediment yield decreased. Hirschi

and Barfield (1988b) proposed this sediment yield decrease was due to lower flow rates in each rill as the

surface runoff was distributed over more rills. Hirschi and Barfield (1988b) also showed that the effects of

rill number on sediment yield is governed by the form of the rill detacbmem and bo1D1dary shear stress

equations.

The WEPP Hillslope Profile Version Erosion Model does DOI represent a rill netwOlk, but rather a

series of parallel rills, and hence rill density is analogous to rill spacing. A sensitivity analysis of the

WEPP model indicated that for most cases rill spacing effects were minimal (Nearing et al., 1989b). How­

ever, the erosion model was sensitive to long slope lengths, and especially when coupled with high slope

gradients. To represent most cases, a default rill spacing of one meter is used in the WEPP model. It should

be pointed out, however; that by assuming parallel rills, the complex interactions of rill networks are neg­

lected. Therefore, there is a need to assess the impact of rill networks on sediment yield in order to substan­

tiate the WEPP model assumption of parallel rills.

1.6.1.4 Rill Network Simulation Models

Numerous stochastic channel network simulation models have been applied on a basin scale. Most of

them are based on random walk models applied to a rectangular grid network (Karlinger and Troutman,

1989; Howard, 1971; Leopold and Langbein, 1962). Although these random walk models yield simulated

networks that match the statistical properties of natural networks, the physical appearance is typically inac­

curate (Abrahams, 1984). Stochastic simulation models rarely represent the Wlderlying physical mecha­

nisms of channel network development, such as topography, consolidation, and soil erodibility. Therefore,

application of these models to rill network development on disturbed soils is inappropriate.

An alternative to random walk type network models is to define channel networks using digital eleva­

tion models. Digital elevation models typically use raster type processing systems to perform neighborhood

operations to extract relevant information. Using a raster based system, Jenson and Dominque (1988)

developed the most comprehensive digital evaluation model to date. The base operations performed on

their model are filling topographic depressions, flow direction delineation, and the spatial computation of

flow accumulation. Using this information rill networks may be defined for a specified rill density. A more

Page 42: Hydrology and Sedimentology of Dynamic Rill Networks ...

33

detailoo description of the techniques for extracting this information from topographic data is described in a

following chapter on rill network generation.

1.6.2 Rill Growth and Development

The development and growth of an individual rill is govemoo by rill detachment potential, transport

capacity, sediment load, and their interactions. The following sections discuss rill detachment potential,

which may be classified as rill incision, headwall cutting, and sidewall sloughing.

1.6.2.I Rill Incision

Rill incision results from concenttated flowing water, which creates shear stresses along the rill bound­

ary. The rate of soil detachment in rills due to rill incision is typically assumed to vary with shear excess,

and may be expressed as (Fostec, 1982):

[l.53]

where D., is the maximum rill detachment rate {kg/m2 /s), t is flow shear slress along the rill boundary

(Pa), t, is the critical shear sttess necessary to detach soil particles (Pa), and a and bare constants. In order

to accurately estimate detachment potential by rill incision, the shear distribution along the boundary must

. be determined. .

1.6.2.1.1 Boundary SMar Stress Distribution

When analyzing rill erosion basrA on shear excess, the knowlooge of the shear sttess distribution is

necessary. The average bed shear stress, t., for uniform flow is given as:

[l.54)

where y is specific weight of watec, R is hydraulic radius, and S is the bed slope of the rill. Using average

shear stress to estimate detachment potential could result in significant errors, since the distribution of the

shear around the rill boundary is nonuniform.

For typical rill cross-sections, shear along the boundary is zero at the water surface and increases to a

maximum along the rill bottom (Raudkivi, 1976). Shear distribution along the rill boundary may be found

indirecdy from the velocity distribution of the flow using the area method. From the velocity distribution,

isovels and orthogonals to the isovels may be obtainoo (Figure l.7a). For channelized flow, on the average,

the orthogonals to the isovels are surfaces of zero shear {Raudkivi, 1976). The normal component of the

mass of water between orthogonals creates a shear SlreSS on the channel boundary.

Page 43: Hydrology and Sedimentology of Dynamic Rill Networks ...

34

(al

( b)

FigiR 1.7. Channel boundary shear SIICSS distribution calculalioos: (a) orthogonals to isovels (dashed liocs)

and (b) Lundgren and Jcmsson method (Lundgren and Ionssm, 1964).

Page 44: Hydrology and Sedimentology of Dynamic Rill Networks ...

35

Using !his concept, Leighly (1932) developed graphical methods to oblain the shear stress distribution

for natural channels. Lane (1953, 1955) proposed an analytical method based on the Leighly (1932) method

using velocity measurements. Lane (1953, 1955) also presented a technique that did not require measured

data, by assuming a power law for the velocity distnbution. Shear stress distributions for various cross­

sections were calculated numerically. Raudkivi (1976) proposed that the bed shear stress between the

ortbogonals to the isovels could be calculated from:

"t = yMS • L

[1.55]

where °'• is average bed shear sttess between ortbogonals. M is the area of water conlained between

ortbogonals, and L is the bed length between ortbogonals. By assuming the lines of zero shear are orthogo­

nal to the channel boundary, and using the power law velocity distribution assumption of Lane (1953),

Rohlf (1981) developed a method to calculate Mand the resulting incremental shear distribution along the

channel boundary.

A simpler method for estimating incremental shear in a shallow symmetrical channel is presented by

Lundgren and Jonsson (1964), based on Prandd's turbulence theory and the normal momentum transfer

across die channel bottom. Lundgren and Jonsson (1964) proposed that ortbogonals could be approximated

by lines perpendicular to the channel boundary extending to the water surface (Figure I.To). They found

that for parabolic channel cross sections, their method gave similar results compared to methods based on a

velocity distribution. Hirschi and Barlield (1988a) confllllled these results, and therefore used the Lundgren

and Jonsson (1964) method in KYERMO.

1.6.2.1.2 Rill geometry

Since rill shape has a significant impact on shear distribution along the boundary (Replogle and Chow,

1966), it is therefore important in rill growth and development. In the absence of a nonerodible layer, Lane

and Foster (1980) observed that rill shape may be approximated as a rectangle with a width given as:

b W=aQ [1.56]

where W is the channel width, Q is discharge in the rill, and a and b are constants. However, Schwab et al.

(1981) states that the equilibrium shape of natural sueams is parabolic rather than rectangular. Rohlf (1981)

found that using a segmented semi-drcular rill shape to estimate rill erosion compared more favorably to

experimental results than a rectangular shape when a nonerodible layer is not presenL Although Rohlf

Page 45: Hydrology and Sedimentology of Dynamic Rill Networks ...

36

(1981) did indicate that a rectangular cross section is developed as a result of side sloughing once the rill

encounters a nonerodible layer. Therefore, rill shape is a function of both rill incision and sidewall slough­

ing. To accurately predict rill detachment the geometry of the rill must be coosidered.

1.6.2.1.3 Critical Shear Stress

An important physical parameter used to classify the erosive resistance of a soil is critical shear stress,

i.e. the shear required to detach and initiate sediment 1raRSpOrt. Shields diagram (Shields, 1936) is the most

widely used method to describe this critical tractive force for individual particles on an open channel bed.

Shields diagram uses a time and 5P&tially averaged dimensionless shear stress and a particle Reynolds num­

ber calcu1ared at the bed. The original Shields diagram does not apply to sediment particles of low specif"ic

gravity and small diameter. Fortunately, Mantz (19TI) extended the Shields diagram for smaller particles.

Shields diagram with the Mantz extension is shown in Figure 1.8.

Studies have been performed that relate critical shear stress to soil properties. Kelly .and Gularte (1981)

found that critical shear stress of cohesive soils is physically rated to soil shear strength. 1n addition, Kelly

and Gularte (1981) found that critical shear stress is highly related to soil salinity and moisture content

Lyle and Smerdon (1965) found that critical shear stress is related to plasticity index, percent clay, mean

particle size, dispelsion ratio, vane shear strength, organic mauer content, cation exchange capacity, and

calcium-sodium ratio. Smerdon and Beasley (1961) developed regression equations relating critical shear

stress to many of these soil properties. Foster (1982) recommended the equation of Smerdon and Beasley

(1961) based on dispersion ratio, however, Hirschi and Barfield (1988a) used the following relationship

from Smerdon and Beasley (1961) based on percent clay:

r. -r = 0.493 IO • [1.57]

where -r, is critical shear stress (Pa), and Pc1 is soil percent clay.

Typical critical shear stresses range from I to 30 Pa. For agricultural soils, Foster and Meyer (1975)

recommended an average value of 2.4 Pa. Alberts et al. (1989) developed regression equations using

extensive field data, and found that critical shear stress of cropland soils may be predicted from:

[C JO.I -r = -2.85-

8·87

o.2

- 16.0C + 3.65S + 3.795°"2 + :;.3

c""' • (IOOS.,,+0.1) ' • ' 1 • 1

[1.58]

Page 46: Hydrology and Sedimentology of Dynamic Rill Networks ...

en (/) laJ er ~ 1.0

er Cl: laJ :c (/)

(/) en laJ .J z 0 (/)

z laJ ~

0.1

,rORIGINAL

' ', ' ~

37

SHIELDS DIAGRAM

MOTION

t NO MOTION -C 0.0, ..... ..1.......11.....~~~ ............ ~~~......1~...L.~~~...L~..L...~~~"---' ,.o 10 100 1000 0.1

PARTICLE REYNOLDS NUMBER

Figure UI. Shields diagram exlellded by Mantz (1977).

Page 47: Hydrology and Sedimentology of Dynamic Rill Networks ...

38

where i, is critical shear stress (Pa), s,., is very fine sand fraction, c. is calcium carbonale fraction, S., is

sodium absorption ratio, S, is soil specific surface area (mg ediyleoe glycol mOIUHldiyl edlec adsorbed per

g soil), S, is sand fraction, C...i is water dispersable clay fraction, and C1 is clay fraction. For cropland soils

with clay fraction greater dlan 0.30, Alberts et al. (1989) found that i, may be predicled from:

i·=-0.5-284 9(9 -0.3) c r r [1.59]

where 9, is die ~olurnetric soil water con1ent (ml !ml ).

1.6.2.2 Sidewall Sloughing

Sidewall sloughing is a significant mechanism in the lateral propagation of rills. Sloughing results

from forces caused by gravity, flow hydraulics and dieir combined effects (Grissingec, 1982). A sidewall's

resistance to failure varies with slope geometry and soil properties such as cohesion, bulk density. void

ratio, moisture conlent and others. For rills, sidewall failure typically result from gravity forces acting on an

ovechang caused by undeccutting (Piest et al., 1975), and translational or rotational slips caused by shear

failure along an in1ernal surface (Bradford et al., 1973). Translational slips occur along a plane typically

parallel to die surface slope, and rotational slips occur along a cin:ular an: (Bradford and Piest, 1980).

Scientific studies of sidewall failure have dealt almost specifically with natural channels such as

ephemeral gullies, gullies, streams and rivers. In addition, very few physically based sidewall failure mod­

els have beeo developed due to the spatial and lemporal variation of die syslem. Mossaail and Wu (1984)

incorporaled rill sidewall stability into their erosion model for a triangular channel cross-section, which is

based on a critical slope concepL Once the sidewall reaches a critical slope, it was assumed to slough to a

stable slope depositing the detached soil mass into the rill Hirschi and Barfield (1988a) also used dlis criti­

cal slope method in the KYERMO model. Bradford et al. (1973) developed a two-dimensional rotational

slip type bank failure model. They found that the factors controlling sidewall stability are water table

height, soil cohesion, and seepage rale.

It should be noled that sidewall failure is a three-dimeosional process, and Bradford et al (1973)

warned that end effects, especially near the headcut, are not negligible. However, until physically based

three-dimensional models are developed, simplistic two-dimensional models are die only alternative.

1.6.2.3 HeadwaUCutting

Page 48: Hydrology and Sedimentology of Dynamic Rill Networks ...

39

The upslope propagation of rills through headcuuing may contribute signif"IC3llt amounts of sediment

The headwall is an abrupt break in the longitudinal channel profile (Schmnm et al., 1984), and is the tran­

sition between wide shallow channelized flo~ and nmrow deeper flow. The headcut migrates upstream,

and several headcuts may exist along the same channel. There bas been liule work on headcut advance­

ment, most of what is available is based on drainage basin morphology (Schumm et al., 1984). Even Jess

work bas been perfonned on headwall initiation, and to date there are no physically based models describ­

ing the process. Therefore, until a physically based model describing headcut initiation and advancement

are developed, the inclusion of headcut processes in erosion modeling is not practical.

1.6.3 Rill Erosion Models

To date, rill erosion models are almost exclusively shear excess based. The following sections intro­

duce and briefly describe the basic components for several state-of-the-art rill erosion models.

1.6.3.1 Foster and Lane

Assuming a rectangular channel cross section, Lane and Foster (1980) and Foster and Lane (1983)

developed a deterministic channel erosion model. Their model was incorporated into CREAMS for

describing ephemeral gully growth in a tilled agricultural field. Although their model was initially devel­

., oped for ephemeral gully erosion, it should also be applicable to rill erosion.

Assuming a constant steady state flow rate, Lane and Foster (1980) separated channel development

into two distinct stages. During the initial stage the channel bouom erodes uniformly downward at a specif­

ic width, and if flow conditions remain constant an equilibrium width is achieved. During the second stage

of development when the channel bottom reaches a nonerodible layer, the channel expands laterally caus­

ing vertical sidewall sloughing. This lateral expansion continues until a fmal width is reached, after which

soil detachment ceases.

Their model is formulated for a rectangular channel under steady state flow conditions. A shear distri­

bution along the channel boundary is assumed, and potential detachment is based on shear excess. The

potential detachment rate is predicted, and therefore an auxiliary equation is required to determine actual

detachment rate based on transport capacity and sediment load. A detailed discussion on the development

and implementation of the model is presented in the soil erosion chapter.

1.6.3.2 MIRE

Page 49: Hydrology and Sedimentology of Dynamic Rill Networks ...

40

Rohlf (1981) developed a Model of Interrill and Rill Erosion Version-I (MIRE I) using simplified one-­

dimensional partial differential equations for gradually varied umteady flow, which consisted of equations

for the continuity and momentum of a water-sediment mixwre and the sediment continuity equation. Rill

detachment was based on shear excess along a semi-circular channel boundary approximated by straight

line segments. Incremental shear was found using the Lane (1953) method and a simple power relationship

was used to define zero shear lines that were assumed perpendicular to the channel boundary. Rohlfs mod­

el accounted for channel geometry variations during formation, and allowed simultaneous detachment and

deposition for a rill cross section. After the rill encountered a nonerodible layer, the semi- circular channel

became rectangular. Validation results indicated that simulated sediment discharges and rill cross sections

were generally less than observed values. A portion of this discrepancy may be due to the model's inability

to account for channel wall sloughing and head cutting.

1.6.3.3 KYERMO

Hirschi and Barfield (1988a) developed a deterministic shear excess erosion model named the Ken­

tuckY ERosion MOdel (KYERMO). Although concepwally similar to MIRE I, KYERMO was developed

with an emphasis on steep slope applications. Incremental shear stress along an arbitrary channel boundary

was found using the area method of Lundgren and Jonsson (1964). Their model accounted for channel wall

sloughing using a similar method used by Wu et al. (1982), and accounted for the interaction of sediment

load on rill detachment rate. Changes in channel geometry were based on shear excess, and deposition and

subsequent detachment in the development rill was also allowed. However, KYERMO does not account for

headwall cutting, and requires prior knowledge of rill density.

1.6.3.4 Mossaad and Wu

An erosion model incorporating rill network development was presented by Mossaad and Wu (1984).

Their model combined a stochastic swface roughness model with a deterministic interrill and rill erosion

model descnl>ing rill network development over time. Rill detachment, based on shear excess, was found

using the Kalinske (1947) bed load equation for triangular channels. Although their model applied the sedi­

ment continuity equation, the transport capacity of the flow was not accounted for. Although their approach

applies some innovative concepts, their model contain significant deficiencies which limit its use.

1.6.3.5 CREAMS

Page 50: Hydrology and Sedimentology of Dynamic Rill Networks ...

41

Several different typeS of erosion may be simulated in CREAMS. Interrill and rill erosion from over­

land flow elements are described using the Foster-Meyer-Onstad (FMO) model for delllChment, and the

modified YAUN equation f<r sediment transport. Erosion from ephemeral gullies uses the Foster and

Lane model described previously, which assumes rectangular channels. Since CREAMS uses the FMO for

eslimaling the combined effects of rill and interrill erosion, the description of rill network development is

not possible. In addition, empiricism inherent with the FMO also limits its use for evaluating erosion sub­

processes on sediment yield.

1.6.3.6 WEPP

A process based soil erosion model was developed for WEPP (Nearing et al., 1989). Toe rill erosion

model component is based on shear excess, and rill detachment depends on rill erodibility, hydraulic shear

stress, surface cover, below ground residue, consolidation, and the ratio of sediment load to IIallsport

capacity. Net deposition is assumed proportional to excess sediment load, and sediment routing is per­

formed by applying the steady state sediment continuity equation.

Average shear stress for a rectangular rill is found using a rill width calculated from rill flow discharge,

average slope gradient, hydraulic radius, and the ratio of friction factorS for the soil and the total rill. Trans­

. port capacity in rills is found using a simplified ttansport equation, calibrated with the Y alin ttansport equa­

tion.

The implementation of WEPP presents a major advancement in erosion prediction using physically

based and process oriented models. WEPP was conceived as a tool to be used for soil and water conserva­

tion and environmental planning and assessment (Foster and Lane, 1987). Although WEPP is an excellent

tool for action agencies, such as the Soil Conservation Service, it does have its limitations. The model does

not account for the development and resulting effects of a rill network. Without this ability, its usefulness in

accurately evaluating new management practices without extensive field data is minimal.

1.6.3.7 Additional RUI Models

There are numerous other rill erosion models, each with their own advantages and deficiencies. An

increasing popular approach is based on energy dissipation rate, i.e. unit stream power concepts (Wilson et

al., 1988). Other notable shear excess rill erosion models include Lu, et al. (1987), Wu and Meyer (1988),

Borah and Bordoloi (1989), and Ewing and Mitchell (1986). An excellent summary of the fundamental

concepts in modeling upland and lowland channel erosion processes is presented by Bennett (1974).

Page 51: Hydrology and Sedimentology of Dynamic Rill Networks ...

42

1.6.4 Conclusions

In recent years significant progress has been made in the understanding and description of individual

rill growth and development Despite these advancements, the physical mechanisms governing soil detach­

ment are still inadequately described. The most common approach to soil detachment in rills is based on

shear excess. However. seepage and gravitational forces cause sidewall sloughing, which is a major source

of detached soil particles. In the past, sidewall sloughing has typically been neglected due to its complex

nature, and when included, oversimplified relationships have been applied. However, until more accurate

quantitative methods are developed, these simple methods are a reasonable alternative.

Another significant source of detached soil in rills is headcut advancement. Little work describing

headcut advancement has been perfonned, and most has been based on the geomorphology of drainage

basins. Until physically based relationships describing this complex process are developed, headcutting will

be neglected.

Another very important aspect (?frill erosion is rill network development. The impact of rill density,

and the temporal distribution of flow has not been adequately addressed. Until the accurate description of

the temporal and spatial development of rills is achieved, the development of an accurate physically based

erosion model is not possible.

1.7 SEDIMENT TRANSPORT MODELS

_Sediment yield from an individual rill, a field, or a basin is controlled by detachment, transport and

deposition. Each of these processes will vary in importance depending on local topographic and soil condi­

tions. Therefore, in order to accurately estimate sediment yield, each of these processes must be adequately

described.

All current sediment transport theory was initially developed for Slrealll flow conditions. However,

significant differences exist between the hydraulics of shallow flows in upland area and deeper channel

flow (Foster and Meyer, 1972b). Current transport models were developed using a limited range of sedi­

ments and flow conditions. In addition, limited transport studies have been performed on aggregates and

fine particles, which are typically encountered in agricultural watersheds.

Transported sediment may be classified as wash load, suspended load, bed load, and total load which is

the sum of the suspended and bed loads. Wash load is made up of particles liner than the bulk of the bed

Page 52: Hydrology and Sedimentology of Dynamic Rill Networks ...

43

malllrial and rarely settle to the channel bottom (Graf, 1971). Bed load is composed of sediment particles

moving along the channel bottom by saltation, rolling, or sliding. Suspended load is bed load size sediment

that is lrllllspOrled in complete suspension for an appreciable length of time. In contrast to bed load, sus­

pended load is much more uniformly dislnbuted throughout the flow depth. When a decrease in transpon

capacity occurs, the suspended load is not immediately deposited on the channel bottom due to the relative­

ly small particle fall velocities.

Due to the complexity of the transport processes, all existing transport models contain a certain degree

of empiricism, and therefore require some degree of calibration. The selection of an appropriate transpon

model for a specific application is difficult at besL Currently, the most common transport models are the

bed_load model of Yalin (1963), and the unit stream power concept developed by Yang (1973) for total

load. Therefore, these two sediment transport models will be discussed in detail in the following sections.

Some other commonly applied total load formulas are given by Ackers and White (1973) and Laursen

(1958), and bed load formulas by Meyer-Peter and Muller (1948) and Einstein (1950)." There are numerous

other models as well, and are summarized in Graf (1971), Simons and Senturk (1977), Alonso et al. (1981)

and Task Committee, ASCE (1971).

L7.1 Yalin Bed Load Model

Yalin (1963) developed a bed load b'ansport model for cobesionless grains of equal size over a mov­

able bed. The model was developed based on dimensional analysis and the mechanics of average grain

motion for uniform turbulent flow with a laminar sub-layer not exceeding the bed roughness. Yalin (1977)

presented the model as a series of equations, given as:

,= q,~ -0.635syll2fl--1 ln(l+as)J (Y/) • L as

y s=--1 y

"'

(1.60]

[1.61)

(1.62]

[1.63)

Page 53: Hydrology and Sedimentology of Dynamic Rill Networks ...

44

where + is a dimensionless transport rate number ( originally defined by Einstein), q, is sediment transport

rate (Nlm-s), p, is lbe particle specific weight in water (N/m1 ), d is mean particle diameter (m), Y is a

mobility number, Y., is the critical mobility number found from Shields diagram, Sa is sediment specific

gravity, and u. is bed shear velocity (m/s). The Yalin equatioo can also be written as a concentratioo by

defining a sediment concentration, C (ppm), as (Alonso et al., 1981):

106

C=~ Y. Vh w

[l.64]

where V is average velocity (m/s) and h is flow deplb (m). Substituting into Equation 1.60 and rearranging

yields:

C = 6.35x105 s,dU. s [1 - ..!..(1 + as>] Vh as

[l.65]

Since the Yalin equation accounts for the transport of uniform grain size, Foster (1982) developed a

methodology to predict the transport of individual particle size classes for a particle size distribution.

Details are presented in a subsequent chapter describing the erosion model.

L7.2 Yang Total Load Equation

Based oo unit stream power concepts, Yang (1972, 1973) introduced a sediment transport equation.

Through regression analysis, Yang found that the most dominant flow characteristic affecting sediment

transport was the product of average flow velocity, V, and the slope of the energy grade line, S. The origi­

nal equatioo, presented by Yang (1972), related total sediment load concenb'ation to unit stream power,

such that

[1.66]

where C. is total sediment concentration (ppm), VS is unit stream power (m/s), V., S is the critical unit

stream power required for incipient motion, and a and pare parameters. Yang (1973) non-<limensionalized

Equation 1.66, yielding:

rod ( u.) . 1og10c, = 5.435 - o.2861og

107- o.4511og1\ 00 +

[L799-0.4091og (rod)-o.3141og (u.],og ( vs - v,..sJ (1.671

10 V I\ (J) 1\ (J) (J)

Page 54: Hydrology and Sedimentology of Dynamic Rill Networks ...

45

where III is panicle fall velocity (m/s), dis media particle diameter (m), v is kinematic viscosity (m2 /s), and

u. is average shear velocity (m/s).

Equation 1.66 was developed using flume and stream data for noncohesive sand wilh a median sieve

diameter between 0.062 and 2 mm, and a specific gravity of 265. The nondimensionaliud form of Yang's

model, Equation 1.67, is applicable to one sediment size, which creates uncertainties when selecting a rep­

resentative panicle size for descnlling a particle size distnllution. This difficulty was reconciled by Hirschi

(1985), which developed a procedure for nonuniform sediment sizes incorporated into KYERMO.

1.7.3 Simplified WEPP Transport Model

Finkner et al. (1989) developed and evaluated a simple quadratic function to estimaie sediment trans­

port along a complex surface profile. Their transport model was incorporated into WEPP (Foster et al.,

1989). The model asswnes a linear increase in overland flow discharge wilh downslope distance, and shear

stress on the soil surface is proportional to the 3(2 power. The profile is represented as a continuous series

of uniform convex and concave slope segments, where for each segment a normalized slope, s •. is defined

· as (rmkner et al., 1989):

S =a.!..+b • L

[l.68]

where x is downslope distance, L is segment length, and a and b are coefficients describing !he degree of

curvature. The transport capacity equation is written as:

T .. = [Ai +Bx+ c]K,.

A= a q .. + l

aq .+b B= • .

q .. + l

q .. b C=-­

q .. + l

(1.69]

(1.70]

(1.71]

(1.72]

where T .. is normalized sediment transport capacity, K.• is a normalized transport coefficient, and q.. is

norma1i7.Cd inflow at the upper end of the profile. Further details defining the nonnalized parameterS are

presented by Finkner et al. (1989). The transport coefficient, K., is calculated from:

Page 55: Hydrology and Sedimentology of Dynamic Rill Networks ...

T K =_!!..

• 't3fl ,.

46

(1.73)

where T00 is sediment transport capacity computed from the Yalin equation, and 't• is shear stress on the

soil at the end of the profile.

Their transport model was incorporated into WEPP (Foster et al., 1989). The model assumes a linear

increase in overland flow discharge with downslope distance, and shear saress on the soil surface is propor­

tional to the 3/2 power.

1.7.4 Transport Equations Applicability and Evaluation

Alonso et al (1981) evaluated several sediment transport equations using extensive laboratory and

field data. The formulas selected were applicable to hydrologic modeling, and covered the range of data

typically encountered in agricultural watersheds. The Y alin equation was recommended for sediment trans­

port capacity computation for overland flow for a range of particle siz.es and densities typically encountered

in field conditions. Transport rates of light materials in stream flow may also be estimated by the Y alin

equation.

In contrast, Alonso et al (1981) found that the Yang equation was best suited for streamflow carrying

fine to coarse sands. However, the transport of smaller sized particles was not adequately predicted. Moore

and Burch (1986) proposed that the Yang equation could be applied to sballow flows if the critical unit

stream power at incipient sediment motion was set to a constanL They indicated that this constant was inde­

pendent of slope. and varied with soil texture. Julien (1987) warned, however, that there is insufficient data

to adequately validate the Yang equation when applied to sheet and rill flow. and therefore should not be

used until several issues are resolved. However, Moore and Burch (1987) argued that their application of

the unit stream power theory to sheet and rill flow is appropriate based on their analysis and suggestions

given by Govers and Rauws (1986).

The simplified transport distribution equation by Finkner et al (1989) was a good approximation when

soil shear stress was ~l Pa somewhere along the profile. The simplified transport model reduces the com­

putational time required for numerical solutions, and the functional form of the equation may be used to

derive closed-form solutions for erosion equations under steady state conditions (Fmkner et al., 1989).

1.7.S Conclusions

Page 56: Hydrology and Sedimentology of Dynamic Rill Networks ...

47

The inherent complexities in describing sediment transpon are reflected in the diversity of transpon

· equations present in the literature. In addition, due to the lack of understanding of basic transport processes,

there is some degree of empiricism in all existing sediment transpon equations. Hence, the app,opdate

selection of a transpon equation depends on the application.

The Yang equation for sediment transpon in rill and interrill areas has great potential. However, fur­

ther studies are required to develop the appropriate fonn of the equation. To date, therefore, the modified

Yalin equation (Foster and Meyer, 1972b) is the transpOn equation of choice for describing rill and interrill

conditions based on the evaluation by Alonso et al. (1981) and the current widespread use. It should be not­

ed, however, that the recently developed simplified transport model incorporated into WEPP (Foster et al,

1989), is an appropriate substitute when specific underlying assumptions are valid.

1.8 PARTICLE SIZE DISTRIBUTION OF ERODED SEDIMENT

To adequately apply physically based erosion processes such as transpon and deposition, particle size

distribution of the eroded sediment must be known. Sediment is detached as either primary particles (sand,

silts, and clays) or as aggregateS. Aggregates are a conglomerate of primary particles cemented together by

organic matter, microorganism excrement. and other inorganic materials such as calcium carbonate, and

iron and aluminum oxides. Primary particles and aggregates have specific weights ranging from 2.6-2.7

tilg/m3 (Hillel, 1982) and 1.64-2.15 mg/m3 (Foster et al., 1985), respectively. Due to their large size, aggre­

gates are primarily transported as bed load even though they are less dense. Primary particles are typically

transported as both suspended load and bed load.

1.8.1 Deposition Effects

For transpon limiting flow, deposition of eroded sediment has a major impact on sediment yield

Deposition occurs when either interrill or rill flow does not contain sufficient energy to transport eroded

sediment. which results from either a decrease in transport capacity or an influx of additional sedimenL A

decrease in transport capacity results from decreased flow or a decrease in slope or flow velocity.

Deposition of eroded sediment is a selective process, which varies with particle size distribution. Coar­

ser particles are more likely to deposit. due to their higher setting velocity (Foster, 1982). Clay particles

may also deposit. although they will typically be contained in the form of aggregates, which due to their

high mass deposit readily as well. For discrete quiescent settling, the settling velocity of an individual parti­

cle may be determined using Stokes law, written as (Barfield et al., 1981):

Page 57: Hydrology and Sedimentology of Dynamic Rill Networks ...

48

V = -1 [!!i. (S - 1)] • 18 v •

[1.74]

where V, is particle settling velocity (mis), dis particle diameter (m), g is acceleration of gravity (mJs2 ), v

is kinematic viscosity {m2 /s), and S1 is particle specific gravity. Equation 1. 74 is valid only up to a Rey­

nolds nwnbec, Re, up to 0.5 (Bartield et al., 1981), which is given by:

Vd R =-'­• v

[1.75]

In addition, Equation 1.74 asswnes spherical particles seuling in turbulent free water. At present, for non­

spherical particles and for particle Reynolds numbers greater than 0.5, experimental estimates are required.

Other factors governing settling velocity of discrete particles are: particle shape, aggregation, turbulence,

and flocculation (Barfield et al., 1981 ).

An appreciable number of studies on deposition processes in reservoirs and detention basins have been

performed. Many of the results obtained from these studies are applicable to interrill and rill erosion. Most

of these studies apply the overflow rate equation, given as (Barfield, et al., 1981):

v F=-'

QIA [1.76]

where Fis fraction of tmpped particles, Q is flow discharge ·(m3 /s), and A is surface area {m2 ). The over­

flow equation asswnes a uniform velocity profile, turbulent free settling, and steady Slate flow. Studies

have found, however, that the presence of turbulence significantly reduces the sediment tmpping efficiency,

i.e. the quantity of deposited sediment {Oten, 1975; Brown, 1950; Camp, 1946; and Dobbins, 1944). For

highly turbulent flow, Oten (1975) proposed that tmpping efficiency may be approximated by:

F=l-exp[- ;A] [1.77]

Using a rectangular reservoir, Camp (1946) developed a graphical solution of the sediment mapping effi­

ciency using the ratios V, A/Q and V, D/2£, where D is flow depth and £ is the turbulent eddy diffusivity

for a specific sediment size. Although these relationships were developed for reservoirs, for many cases

these conceplS may be applied to overland and concentmted flow conditions. However, to utilize their

results the turbulent diffusivity must be known, which is in essence an empirical coefficient and varies con­

siderably with flow conditions. Further details describing their concepts may be found in Simons and Sen­

turlc (1977) and Barfield et al. (1981).

Page 58: Hydrology and Sedimentology of Dynamic Rill Networks ...

49

Experimental and theoretical studies have shown that deposition rate decreases exponentially with dis­

tance downsueam (Einstein, 1942; Foster and Meyer, 1972a, 1975). However, due to the complexity of

implementing these concepts, current erosion modeling efforts neglect the exponentially decaying deposi­

tion rate, and assume excess sediment load is instantaneously deposited (Nearing et al., 1989; Hirschi and

Barfield, 1988a; Rohlf, 1981). Therefore, there is a need to implement this into current erosion models.

1.8.2 Methods to Estimate Eroded Particle Size Distribution

Toe simplest method to estimate the eroded particle size distribution is to approximate it from a sample

of the original parent soil material. However, using this approach yields questionable results, especially if

the analysis disperses the soil aggregates. In order to avoid these inaccuracies, the following sections

describe alternative methods.

1.8.2.1 Experimental Procedure Using a Rainfall Simulator

One method to estimate an eroded particle size distribution may be found by applying simulated rain­

fall IQ a soil sample placed in a pan (Barfield et al., 1979). The rainfall simulator system consists of a Vee­

Jet 80150 nozzle placed 3.1 m (10 ft) above a sloping pan which empties into a large buckeL The system is

run until the rainfall volume equals the desired storm evenL Using this sample, an appropriate particle size

analysis is determined experimentally to fmd the eroded particle size distribution. Barfield et al. (1979),

however, caution that low intensity rainfall applications may not be applicable since the spray nozzle

approximates high energy content rainfall of high intensity. Using measured eroded particle distributions

for three tillage treatments, Holbrook et al. (1986) concluded that the method proposed by Barfield et al.

(1986) worked well for the Cecil soil series.

1.8.2.2 Use of Equations

An alternative to experimental methods for estimating eroded particle size distributions is the use of

equations based on the fraction of primary particles found in the parent material. Foster et al. (1985) devel­

oped such a procedure, dividing the sediment into 5 particle classes: primary sand, silt and clay, and small

and larp;e aggregates. The primary sand, silt and clay are assumed to have average diameters of 0.2, 0.01,

and 0.002 mm, respectively, with specific gravities of 2.65 y/cmJ . For the primary particle classes, the

sizes range from < 0.004 mm, 0.004 to 0.63 mm, and > 0.063 mm for primary clay, sand and silt, respec·

lively. Toe small and large aggregates are assumed to have specific gravities of 1.80 and 1.60 y/cm2 ,

respectively. Toe diameter of the small aggregates, D,. , in mm is calculated by:

Page 59: Hydrology and Sedimentology of Dynamic Rill Networks ...

D., =0.030 Od<0.25

D ,, = 0.2( 0., - 0.25) + 0.030 0.25 S O., S 0.(,()

D = 0.100 0 1 > 0.(,() ., .

50

[1.78]

[1.79]

[1.80]

where Oc1 is the clay fraction in lhe soil matrix. The large aggregate diameter, Di. , in mm is calculated

from:

D,, =0.300 o.,so.15

D11

=20., 0.,>0.15

The fraction of lhe sediment composed of the primary clay class, F c1 , is calculated from:

F .,= 0.21, o., The fraction of sediment composed of small aggregates, F" , is calculated from:

F., = 1.8 o., O., < 0.25

F., = 0.45 -0.6(0.,-0.25) 0.25 :SO., :S 0.50

F,1

= 0.(,() 0., 0., > 0.50

The fraction of primary sand classes is calculated from:

5 F =0 (1-0)

- - d

where 0,. is the sand fraction in lhe matrix soil. The fraction of primary silt class is calculated from:

F.=0.-F "' .a ,,

The fraction of the large aggregate classes in the sediment, Fia , is calculated as:

F = 1-F -F .-F -F ,, c1 ..... ,,i11

[1.81]

[1.82]

[1.83]

[1.84]

[1.85]

[1.86]

[1.87]

[1.88]

[1.89]

If the value of F., <O, then the values of the other fractions are proportionally reduced to give zero for Fi. .

The primary particle composition of the sediment classes are calculated as: Primary clay:

f.~= 1.0 [1.90]

f,.,.,=0.0 [1.91]

f ""'' = 0.0 [1.92]

Primary silc

f.1,.=0.0 [1.93]

Page 60: Hydrology and Sedimentology of Dynamic Rill Networks ...

Primary sand:

Small aggregare:

Large aggregate:

!..-.,.= 1.0

f. .=0.0 -.f,,,._=0.0

f.. = 0.0 .. .,.

o" fci.,.= 0 +0.

cl s

0. !.;.., = 0 +" 0 .

cl .,,

f ""'' = 0.0

0 -F !. "' "' ..,,,= F ,,

51

[1.94]

[1.95]

[1.96]

(1.97]

[1.98]

(1.99]

[1.100]

[1.101]

[1.102]

[1.103]

[1:104]

where f is the fraction of clay, silt or sand for a given particle class, the fllSt index is the primary particle

type within the particle class, the second index is the sediment particle class, and cl=clay, si=Silt, sa=sand,

sg=small aggregate, and lg=large aggregate. To ensure aggregate stability, Foster et al. (1985) assumed that

the clay content of the large aggregate must be at least one-half that of the soil mattix. If the computed clay

content is less than this required fraction, then the fraction of sediment in the small aggregate class is

recomputed as:

= [0.3+0.5(Fc1+F.,.+F14)] (O"+O,.)

I., · 1-0.5(0"+0,.) [1.105]

Page 61: Hydrology and Sedimentology of Dynamic Rill Networks ...

52

These equations describe the sediment at the point of detachment and are only rough estimates for repre­

selllative particle classes and their composition, but are the best estimates until more precise relationships

are developed.

Barfield et al. (1984) compared these regression equations to obsel"ved eroded particle size distribu­

tions from rainfall simulator studies in Western Kentucky. They found that the size distribution of topsoil

sediments were well predicted, shale spoil sediments were moderately well predicted, and subsoil sedi­

ments were poorly predicted. Storm (1986) also found discrepancies when applying the equations to Gro­

seclose silt loam and Suffolk sand loam soils, and presents a procedure to reconcile the discrepancies.

Comparing measured eroded particle size distribution for three tillage treatments, Holbrook et al. (1986)

concluded that Foster's regression equations worked well in estimating Dso values.

Based on rainfall simulator studies, Diezman et al. (1987) developed regression equation describing

the eroded particle size distribution under no-till and conventional tillage systems. Their equations included

effects of soil moisture, slope, rainfall volume, and the composition of the original matrix material. They

found that sediment aggregates were finer than those found in the matrix material, and were enriched with

primary clay particles. Their studies also found that eroded sediment from no-till treatments had coarser

aggregates that contained higher .fractions of primary gravel, sand and clay particles compared to conven­

tional tillage treatments. These equations descnbe the IIansported eroded sediment, and are only applicable

to the special soil and topographic conditions under which they were developed.

1.8.3 Conclusions

Detachment and IIansport processes are highly dependent on sediment particle size distribution. There­

fore, to accurately estimate sediment yield with physically based erosion models, the eroded particle size

distribution must be known. In addition, deposition of eroded sediment is a selective process, and varies

with particle shape, degree of aggregation, turbulence, and flocculation (Barfield et al., 1981). For 1Iansport

limiting flow, the deposition process must be adequately described in order to accurately estimate sediment

yield. Although previous erosion models have assumed instantaneous deposition, it has been shown that

deposition rate decreases exponentially with distance downslope. These fundamental concepts are essential

if accurate estimates of sediment yield are to be obtained for un-gauged areas.

Soil are very heterogeneous materials and their properties varies spatially as well as temporally. There­

fore. estimating an average eroded particle size distribution for a specific soil is questionable at best. How-

Page 62: Hydrology and Sedimentology of Dynamic Rill Networks ...

53

ever, two methods have been shown to produce reasonable results. The Barfield et al. (1979) method uses

controlled rainfall simulator experiments using a soil sample, and Foster et al. (1985) proposed regression

equations based on the fraction of primary particles of the parent soil material.

Page 63: Hydrology and Sedimentology of Dynamic Rill Networks ...

54

CHAPTER2

DYNAMIC RILL MODEL OVERVIEW

The ovemll objective of this project is to develop a physically based model describing the runoff and

erosion processes, and the development of rill networks as affected by soil properties, topography, and sur­

face roughness conditions. Due to the complexity of the erosion processes, the effects of plant residue and

cover will not be accounted for. In addition, the micro-relief features described by the model will be ran­

dom surfaces without any drainage preference.

The comprehensive model describing these processes is called DYnamic Rill NeTwork Model

(DYRT). Containing four major components, DYRT is composed of: 1) a random surface generator, 2) a

rill network delineation routine, 3) a hydrologic model, and 4) an erosion model. In order to efficiently

execute the hydrologic and erosion models, the rill network routine is linked by a separate interfacing rou­

tine that preprocesses the rill network information (see Figure 2.1).

For a specified rectangular area, slope, and random roughness parameters, the surface generator simu­

lates a random micro-relief, which is superimposed onto a macro-relief. Using this generated surface, the

rill network routine defines a preferential flow path network containing a series of branches and their sub­

watershed areas. The rill network and random surface are then used by the preprocessor to combine this

information and create an organized array of scratch files. The hydrologic model uses these scratch files

along with other hydrologic inputs to estimate rainfall excess, and rill and interrill flow depths and veloci­

ties over the entire response area. This information, along with preprocessor scratch files, is used by the

erosion model to define interrill and rill erosion rates, rill development, and sediment yield.

Page 64: Hydrology and Sedimentology of Dynamic Rill Networks ...

55

Random Surface Generator

• Rill Network Delineation

Preprocessor - --

-< •

Hydroloi ic Model

< ..

Erosion Model

Figure 2.1. Schematic diagram of the component processes for DYRT.

Page 65: Hydrology and Sedimentology of Dynamic Rill Networks ...

56

CHAPTERJ

RANDOM SURFACE GENERATOR

For specified random roughness parameters, the surface generation model simulates a tw<Hlimensional

random field representing lhe soil surface. A simulated micro-relief is superimposed onto a macro-relief,

which is a rectangular area oriented at a given scope. The melhod chosen to simulate lhe surface micro­

relief is the turning bands melhod ("fBM), originally introduced by Matheron {1973). By converting a mul­

tidimensional simulation into a series of equivalent one-dimensional simulations, lhe TBM creates a ran­

dom field using a specified covariance function. Although the TBM may be used to simulate

three-dimensional fields, only theory related to twCH!imensional simulations is presented. The following

sections summari7.e the TBM development using infonnation presented by Zimmennan and Wilson {1989),

Mantoglou and Wilson {1982), and Mantoglou and Wilson (1981).

3.1 TURNING BANDS METHOD

For the generation of random microrelief surfaces in twCH!imensions, it is assumed that lhe random

field is second-order stationary and isob'Opic. In order to generate lhe random field, the form of lhe covari­

ance must be specified. Covariance of two variates X and Y, C{X, Y), is the joint second moment about lhe

means,µ, and~· which may be written as {Haan, 1977):

C{X,Y) =a..,= E[(X- µ,)(Y-µ1)] =E[X]E[Y] (3.1]

or

(3.2]

where E[] is lhe expectation operator, and p.., {x,y) is lhe joint probability distribution between X and Y.

C{X, Y) is a measure of lhe interaction between the variates X and Y, and is typically normalized, such that

p = C{X,l') (1(1

• J

(3.3]

where p is a correlation coefficient, and a. and a, are the square root of lhe X and Y variances, respective­

ly. A process is said to be covariance stationary, or second-order stationary, if {Mantoglou and Wilson,

1982):

E[Z{x)]=m (3.4]

Page 66: Hydrology and Sedimentology of Dynamic Rill Networks ...

57

and

E[Z(i + 0 z.(i)] = C(0 [3.5)

where Z( 1 ) is a random field, 1 is a location vectm, m is the mean, and C is a separation vectm between

points 1 and x+C. Equations 3.4 and 3.5 indicate that the mean of a second-order stationary pocess is con­

stant, and the covariance between any two points depends only on the distance between the two points and

not their location in the random field In addition, the isotropic assumption implies that the random field is

invariant with respect to direction.

It will-be shown in a following report on model validation and testing, that the value of each point in

the field is normally distributed with a zero mean. In addition, the two-dimensional covariance function

C.J.r), was found to follow a simple exponential form given by:

[3.6)

where r is the separation distance (m), a' is variance of the process (m2 ), and A is a correlation parameter

(m).

Toe correlation parameter A is related to the integral scale, which is a measure of the average correla­

tion length between field variables. Tennekes and Lumley (1987) define the integral scale, A1, as:

')..1=-\- jC(r) dr a o

For the exponential covariance function in Equation 3.7, the integral scale is given as:

-l lf2-.{l.' . ..,= 2 a e dr=11. a o

Thus indicating the integral scale and correlation parameter are equal.

[3.7)

(3.8)

For a two-dimensional second-order stationary random field, the TBM is implemented by generating

lines from an arbitrary origin at uniform intervals between O and 21t (see Figure 3.1). Along each line i, a

series of discrete deviates are generated having a zero mean and covariance function C1 (0, where ~ is a

distance coordinate along line i. Given a random field represented by discrete points, the value of a random

variable at a point "o" in the random field, Z. (x. ), is calculated from (Mantoglou and Wilson, 1982):

L

z,<i) = _Jr f.'Z.(x •• ii,) (3.9)

Page 67: Hydrology and Sedimentology of Dynamic Rill Networks ...

58

where x, represents the locatioo vector from the turning band origin to the generated point at "o", L is the

number of turning band lines, and u, is a unit vector oo line i. Toe dot product (x,•uj represents the scalar

projection of x, onto Mp i.e. the component of x, in the direction of u, For a given line i, the simulated val-

ue occurs a distance I;,, from turning bands origin, and is given by:

C =1 • U. • • • (3.10]

For each line, an independent realization using the one-dimensional covariance function C1 (C) is per­

formed. Hence, for each simulated value in the random field, L values of Z; (!;,,) are weighted.

As stated previously, a tw~ional random field is generated using a series of one-dimensional

simulations. To generate this field, a two-dimensional covariance function C2 (r) must be specified. To

preserve the statistical properties of the isotropic two-dimensional field, Mantoglou and Wilson (1982)

mathematically derived a relationship between the one-dimensional covariance function C1 (0, and C2 (r),

such that

' c <C) c (r) = 1.. f 1 dC 2 lt • ...; cr2-c2)

[3.11]

However, due to the form of Equation 3.ll,C1 CC) may not be expressed in terms ofC2 (r) in a usable form.

To circumvent this problem, Mantoglou and Wilson (1982) developed relationships between C1 (0 and C2

(r) using spectral density functions. Expressing C1 (0 as a Fourier integral, the Fourier transform pair may

be written as:

[3.12]

(3.13]

where i is Ci, (I) is the spectral frequency, and S1 ((I)) is a one-dimensional spectral density function. For a

real process, the Fourier transform pair may be written as:

C1(0=2 jcos(me)S1((1))dco (3.14]

S1((1)) = ! jcos(mQ C1(C) dC (3.15]

Page 68: Hydrology and Sedimentology of Dynamic Rill Networks ...

59

Y Random Field t

ml•fieU-6-,._. limele Pti1l

\

I, \

\ __ Prtjeclh1 \ \

........__,_ li1ercli1ei Lite Precess [Z.l~l]i

Figure 3.1 Disaele random field gcneilllion using the turning bands method.

Page 69: Hydrology and Sedimentology of Dynamic Rill Networks ...

60

For an isottopic field, Mantoglou and Wilson (1982) show that the one-dimensional spectral density func­

tion S1 (m), is related to the two-dimensional radial spectral density function f(ro) by:

cl S (ro) = --"'m) l 2" [3.16]

where c? is the variance of the two-dimensional process. Substituting Equation 3.16 into Equation 3.14

yields the relationship between C1 (0 and f(ro), such that

C1(0= cr2jcos(rol;)Jl:ro) dm [3.17]

• The next step is to develop the relationship between the two-dimensional radial spectral density func­

tion f(ro) and the two-dimensional covariance function Ci (r). For an isotropic tw<Hlimensiooal field, Man­

toglou and Wilson (1982) developed the following Fourier transform pair:

CiCr) = a2~ro)J

0(mr) dro [3.18]

fl.m)= ;ici<r)J.(mr) rdr •

[3.19]

where J0 ( ) is a Bessel function of the first kind of order zero. Using Equation 3.19, Mantoglou and Wil­

son (1982) derived the two-dimensional radial spectral density function for the exponential covariance

function (Equation 3.6), such that:

2 fl.ro) = OlA.

[1 + (OlA.)2]3(2 [3.20]

They also developed several other radial spectral density functions for a variety of covariance functions

typically encountered in natural systems. Therefore, from Equations 3.16 and 3.20 the one-dimensional

spectral density function S1 (m) may be written as:

[3.21]

3.2 SPECTRAL LINE GENERATION

For each turning band line a one-dimensional line process Z1 (0 may be generated using spectral

methods. A spectral density function is based on the frequency domain, and descnbes how the system vari-

Page 70: Hydrology and Sedimentology of Dynamic Rill Networks ...

61

ance is distributed among the frequencies. Therefore, the system variance is the integral of the spectral

density function over all frequencies. In practice, however, the frequency domain is divided inlO discrete

intervals over a finite frequency range. If the range of frequencies is large enough, the resulting system

will closely approximate the theoretical process.

Based on the Fourier-Stieltjes integral {Walker, 1988), Zimmennan and Wilson (1989) developed the

representation of the one-dimensional line process, which may be approximated by:

[3.22]

where Re indicates the real component, M is the number of hannonics, ro1;=(j+ 1/2)11.ro, 11.ro is an incremental

frequency, U; and V; are normally distributed uncorrelated variables with a zero mean and variance of 1/2.

The incremental frequency 11.ro is defined as:

0 11.ro= -

M

where O is the maximum frequency over which Z1 (0 is being integrated.

[3.23]

Equation 3.22 represents a discrete fourier transform, which is an approximation since it is being eval­

uated over a fmite number of frequencies truncated as 0. In order to reduce the computational requirements

of generating Z1 (0, the Fast Fourier Transform (FFij method is applied {Weaver, 1989). The FFT method

is an efficient computational technique, which is used to generate discrete fourier transforms. Zimmerman

and Wilson (1989) present a discussion on the implementation ofFFT to Equation 3.22.

In order to obtain a reasonable numeric approximation, the appropriate number of hannonics M and

the maximum frequency O must be specified. Zimmerman and Wilson (1989) indicate that if O is too small

high frequencies are neglected, which results in the generated system variance being less than the theoreti­

cal variance. In addition, the frequency spacing, 11.ro, must be small enough to accurately represent the spec­

trum at low frequencies when rapid changes in the covariance function occur.

3.3 TUBA

The computer code for generating two-dimensional isotropic random fields via the turning bands meth­

od is Tuba Version 2.0. Tuba Version 2.0 has been modulariz.ed and vectorized, which increases its compu­

tational efficiency. In order to run Tuba 2.0 on an IBM 3090 mainframe computer, it was necessary to

make minor modifications to the code. A user's manual describing the development and application of the

Page 71: Hydrology and Sedimentology of Dynamic Rill Networks ...

62

medlod is presented by Zimmerman and Wilson (1989). Tuba was first introduced by Mantoglou and Wil­

son (1982), and since then there have been numerous developments and applications by Thompson et al.

(1989), Mantoglou and Wiison (1982), Mantoglou (1987), and othels.

Five covariance functions are available in Tuba 2.0: exponential, Gaussian, Bessel, Telis, and Gener­

alized Covariance models. Tuba can generate either isottopic or anisotropic fields, areal average processes,

and can generate a grided system or at arl>ittary locations in space. In addition, both nonnal and lognor­

mally distributed fields may be generated. Although Tuba 2.0 is a versatile computer program, only the

exponential covariance function is used to generate a normally distributed isotropic field on a uniform grid.

TIJBA is based on a discrete line process, and therefore the simulation parameters must be chosen

appropriately to minimize approximation enors. From a sensitivity analysis, Zimmerman and Wilson

(1989) found the truncation frequency O should be at least 100 for the exponential covariance function,

when the one-dimensional line process is generated with the FFf method. They also found that the number

of harmonics M should be 2048, which indicates from Equation 3.23 that the frequency interval Ll.ro should

be 0.05. Another parameter that must be specified is the discretization distance LI.I; along the one­

dimensional space domain. For the FFf method, LI.I; is related to the maximum frequency by:

[3.24]

Zimmerman and Wilson (1989) recommended !hat in general, the discretization distance not exceed:

LI.I; = 0.06 A. [3.25]

3.4 SUMMARY AND CONCLUSIONS

A random surface generation model is presented that simulates an isotropic two- dimensional random

field, which accounts for the inherent two-dimensional spatial correlation found in disturbed soil surfaces.

A simulated micro-relief is superimposed onto a macro-relief at a specified slope. The micro-relief is gen­

erated using TIJBA version 2.0 (Zimmerman and Wilson, 1989), which applies the turning bands method

(Mantoglou and Wilson, 1982). Converting a multidimensional simulation into a series of equivalent one­

dimensional simulations, the turning bands method efficiently generates the random surface for a specified

covariance function.

Page 72: Hydrology and Sedimentology of Dynamic Rill Networks ...

63

CHAPTER4

RILL NETWORK GENERATION

Given a random surface micro-relief, a digital elevation model (DEM) identifies preferential flow

palbs, and defines the rill network. The DEM contains geographic infonnation system (GIS) based

FORTRAN subroutines developed at the USGS EROS Data Center by Jenson and Dominque (1988),

which extract relevant information from the digital elevation data. A GIS may be defined as a computer

based infonnation management system, whereby spatial data such as land-use, topography, and soils may

be stored, retrieved, and analyzed. Infonnation obtained by the DEM is used to find preferential flow

palbs, define the rill network, subdivide the network: into a series of connecting branches, and define sub­

watersheds for each branch.

Using a raster based processing system, the Jenson and Dominque subroutines perform a series of

neighborhood operations on digital elevation data. A raster based system represents a response area as a

series of points that create an interconnected grid of cells. Neighborhood operations are perfonned on a

center cell, surrounded by eight adjacent cells. Base operations are perfonned using the subroutines to fill

single and multi-cell topographic depressions, define flow directions, and calculate flow accumulation val­

ues for each cell. Using this infonnation, rill networks are defined and subdivided into branches using a

special numbering scheme. Applying another subroutine developed by Jenson and Dominque (1988), sub­

watershed cells for each branch are found. Using this information, subwatershed boundaries are identified.

4.1 FLOW DIRECTION COMPUTATION

After subdividing a response area into a raster based system, flow directions for eacb cell are defined.

Four possible conditions exist when determining flow directions for each cell. Condition 1 occurs when all

eight neighboring cells have greater elevations than the center cell, which results in an undefined flow

direction. This is a single cell depression, which is descnbed in the next section. Condition 2 occurs when

the center cell's distance-weighted drop is higher for one neighboring cell. and thus the flow direction is

assigned towards that cell. Distance-weighted drop is defined as the difference in elevation between the

center cell and a neighboring cell divided by the distance between cells. For diagonally located cells this

distance is -./2, and for non-diagonal cells is one. When 2 to 7 cells have the highest distance-weighted drop

Page 73: Hydrology and Sedimentology of Dynamic Rill Networks ...

64

of equal value, condition 3 exists. When this occurs, the flow direction is assigned based on predefined

rules of logic.

When the center cell is located in a flat area, and all neighboring cells have the same distance-weighted

drop, an iterative procedure is required to assign the flow direction. Before condition 4 cell directions can

be found, condition l, 2, and 3 cell directions must be assigned. Testing one cell at a time, a flow direction

is assigned, and if the flow direction does not result in flow returning back to the original cell, the flow

direction is accepted. This iterative procedure continues until all cells have defined flow directions.

Single cell depressions, or condition 1 cells, are assigned flow directions first, using procedures

described in the following section. Condition 2 and 3 cells require minimal computation, however condi­

tion 2 cell occur more often. Conversely, condition 4 cells may require significant computational time to

assign flow directions. An example of a condition 2 cell is shown in Figure 4.1.

4.2 FILLING TOPOGRAPHIC DEPRESSIONS

In order to defme a series of continuous flow directions, topographic depressions must be filled since

they create a sink, which inhibits flow routing. A depression may be a single cell or composed of multiple

cells, and occur when· all surrounding cells· have higher elevations. The first step in eliminating these

depressions is to fill all single-cell depressions. For each single-cell depression, the cell elevation is artifi­

cially increased to the level of its lowest neighboring cell. Filling these single-cell depressions reduces the

complexity of filling multi-cell depressions.

After first estimate flow directions are determined, multi-cell depressions are filled. To determine the

amount of filling required for each multi-cell depression, known flow directions are used to delineate

uniquely identified watersheds for each spatially connected depression. Next, all remaining watersheds are

delineated, and create a pour point elevation table between all adjoining watersheds. A pour point is the

lowest elevation cell connecting two watersheds. The next step is to determine the lowest pour point eleva­

tion for each watershed, and for duplicate minimum pour points arbitrarily select one. For each depression

watershed, the highest elevation is defined as the minimum pour point path, and the elevation of all depres­

sion watenhed cells with lower elevation are increased to this new pour point elevation. After the single

and multi-cell depressions are filled, the resulting "depressionless" DEM allows continuous flow paths that

eventually reach the edge of the data set.

4.3 FLOW ACCUMULATION AND RILL NETWORK

Page 74: Hydrology and Sedimentology of Dynamic Rill Networks ...

65

Flow Directions

""'t / ~-?-..... •

/ 53 51 49

44 33 42

" 41 35 l~o

FigiR 4.1. Example showing the flow dircclion computation used in the GIS based routines.

Page 75: Hydrology and Sedimentology of Dynamic Rill Networks ...

66

Using flow directions assigned to each cell, flow accwnulation for each cell may be found. Flow accu­

mulation is the nwnber of cells flowing into a specific cell, and is shown in Figure 4.2 for a simple 3 row

by 5 colwnn example. For a given flow accwnulation threshold, a rill network may be defined using the

flow accumulation data seL The rill density may be increased by simply decreasing the threshold value. In

the example in Figure 4.2, the channel network for a threshold value of 2 cells is shown.

Once the network is defined, rills are nwnbered left to right looking uphill, and are divided at junction

nodes into a series of branches (see Figure 4.3). For each junction, there are eight possible flow directions.

Branch nwnbering proceeds in a clockwise direction starting with the flow direction at the "1:30" clock

position (see Figure 4.3). For each rill, the initial junction for branch enwneration is found by following

the maximwn flow accwnulation gradient After the upper most junction is enwnerated the previous junc­

tion is evaluated. From this junction a new unenwnerated path following the maximwn flow accwnulation

gradient is established. This process continues until all branches are nwnbered for each rill (see Figure

4.3).

4.4 SUBWATERSHED DELINEATION

Once the rill network is divided into branches, all cells contained in the subwatershed for each branch

are found. In addition, boundary cells are also identified to allow graphical display. Subwatersbed cells are

found using cell flow directions and a subwatersbed starter data seL The starter data set contains a file with

unique numeric values for each subwatershed which are placed in the raster data file at the last downstream

cell location for each branch. Applying an iterative procedure, all cells in each subwatershed are assigned

the corresponding numeric value from the starter data seL An example illustrating this procedure is shown

in Figure 4.2.

4.5 OVERLAND FLOW LENGTH

Rill network density is determined by the selection of a threshold value, which represent the minimwn

cumulative nwnber of cells that flow into an individual cell. For a uniformly grided system, this threshold

value may also be interpreted as the overland flow length required to develop concenttated flow. In order to

select an appropriate threshold value, it is assumed that a concenttated flow region occurs when a specified

rainfall excess resu!IS in flow shear stresses thai exceed the soil's critical shear stress.

Assuming uniform flow in a wide shallow rill, critical shear stress may be expressed as:

Page 76: Hydrology and Sedimentology of Dynamic Rill Networks ...

67

Flow Directions

"" - i ! / t - i / -/ - i - -Flow Accumulation Values

2 0 0 0 0

0 4 9 3 0

0 0 14 1 0

<:bannelN'etwork

"" - i / t

Subwatershed Starter Data Set

1 2

3

Subwatershed Data Set

1 1 1 2 2

.1 1 1 2 2

1 3 3 3 3

Figun: 4.2. Rill network delineatioo example.

Page 77: Hydrology and Sedimentology of Dynamic Rill Networks ...

Br11ch lulherinc lirectio1 Priority

8 181 Priority

Example Branch Nllllherinc Scheme

3 3

9 (

8 2-.1 2-1

8

10

11 ( 6

l 12

7 7

RILL 1 RILL 2 IILL 3

Figure 43 Rill nelWOdt delinealion numbering scheme.

~ ·­-co .. ... ·-.... -0 ;:::

68

Page 78: Hydrology and Sedimentology of Dynamic Rill Networks ...

69

,: = yd s • • (4.1]

or

,: d=-· • yS

(4.2]

where ,:, is soil critical shear stress (Pa), y is the specific weight to water (N/m3 ), de is the critical flow

depth to initiate soil detachment (m), and S is the rill slope (m/m). Assuming the hydraulic radius R=d,, ,

the Cbezy Equation for open channel flow can be written as:

v =C.../dJ • • [4.3]

or

q =dv =c,il2slll c t: c c

[4.4]

where v. and q. are the flow velocity (m/s) and flow rate (m3 /s-m) required to initiate soil detachment,

respectively, and C is the Chezy fraction coefficient related to channel roughness (m112/s). Under uniform

flow conditions, the Cbezy friction coefficient may also be expressed as:

(4.5]

where g is acceleration of gravity (9.81 m/s2 ), and f is the dimensionless Darcy-Weisbach friction factor.

In order to determine the critical threshold value, a uniform rainfall excess rate, i (m3 /s-m2 ), must be

specified. For a uniform grid network representing the response area, the rainfall excess per cell, er (m3 /s),

is:

b2. Cf= l (4.6]

where bis the cell width (m). The critical flow rate, Q, (m3 /s), required to initiate soil detachment is:

Q,= w ,.q. [4.7]

where W oq is the equilibrium channel width defined by the Foster and Lane channel erosion model {Foster,

1982). This critical flow rate is also related to rainfall excess by:

Q =11 er • • [4.8]

where "I\, is the threshold number of cells required to initiate soil detachment. Equating Equations 4. 7 and

4.4, and substituting Equation 4.8, yields:

Page 79: Hydrology and Sedimentology of Dynamic Rill Networks ...

70

1/2 Q = 11 a=[!.&.] i'1' s112 w

t: t: I c "' [4.9]

Substituting Equation 4.2 into Equation 4.9 yields:

11·a=f [¥J1; r/2 [4.10]

and rearranging for 11. and substituting a=lli gives:

W [!.&.Jl/2 [ 'tc ]3/2 11.=~ f y [4.11]

4.6 SUMMARY AND CONCLUSIONS

A raster based procedure is described which defines a rill network using GIS based computer software

developed by Jenson and Dominque (1988). The rill network is divided inro a series of branches, and the

sub watershed area for each branch is found. The steps used to define the rill network are: 1) fill single cell

depressions, 2) fill multi-<:ell depressions, 3) define flow directions, 4) calculate flow accumulation values,

5) define a rill network for a specified threshold value, 6) divide the rill network into branches, and 7)

determine subwatershed cells and boundary cells for each branch.

Page 80: Hydrology and Sedimentology of Dynamic Rill Networks ...

CHAPTERS

HYDROLOGIC MODEL

71

A deterministic hydro logic model named HYMODRIN (HY drology Model Of Dynamic Riii Net­

works) is used to deiermine rainfall kinetic energy, rainfall excess from interrill areas, inflow interrill

hydrographs into the rill network, and the routing of these hydrographs through the rill netwodr:.

Toe generated random surface is subdivided into intenill elements, where infiltration is estimated

using a dual layer Green-Ampt model based on Chow et al. (1988). Toe interrill flow element hydrographs

are routed to the rill network using a nonlinear reservoir approach, which requires a Chezy friction coeffi­

cient for each inlerrill elemenL Depression storage is a function of random roughness, which is allowed to

decrease with time according to the cumulative rainfall kinetic energy.

Inlerrill flow hydrographs are routed through the rill network using a finite element formulation of the

diffusion wave equations. As discussed in a previous chapter, the rill network is subdivided into a series of

branches, and rainfall excess is routed through each of these branches. A detailed description of the hydro­

logic model and its development is presented by Ormsbee et al. (1990) in Volume ll of this reporL

Page 81: Hydrology and Sedimentology of Dynamic Rill Networks ...

72

CHAPTER6

SOIL EROSION MODEL

The dynamic erosion model used in this project predicts rill and interrill erosion from a topographical­

ly random micro-relief surface without any drainage preference. In addition, model components contain

currently acceptable erosion prediction techniques, and when appropriate contain the most physically based

equations available. The following sections describe the development of prediction equations for interrill

erosion, rill detaclunen1, deposition in rills, rill shape, transport capacity of the flow, and the interaction

between sediment load and transport capacity.

6.1 SEDIMENT ROUTING

Toe primary governing equation for fundamental erosion processes is the continuity equation for mass

transport. Assuming relatively shallow flow, the steady state continuity equation used to describe sediment

routing down-slope is:

'i)q fW -a' =D+D.

X r • [6.1]

or

'i)q -a' =E +E.

X r • (6.2]

where q, is sediment load (kg/s), W is rill width (m), x is distance down slope (m), D, is rill erosion/

deposition rate per unit rill width and length (kgfm2 /s) I}, is int.errill sediment delivery rate to the rills per

unit rill width and length (kg/m2 /s), E,, is rill erosion/deposition rate per unit rill length (kg/m/s), and E; is

interrill sediment delivery rate per unit rill length (kg/m/s). Toe i,q, /ox term represents the change in sedi­

ment storage rate along the slope. Equation 6.1 assumes spatially uniform sediment concentration, neglects

sediment dispersion, and the change in sediment storage over time.

Derived from a Taylor series expansion, the central difference method is used to numerically solve

Equation 6.1. A Taylor series expansion of a function, f(x+lll) and f(x-lll) is:

fi'.x+~) = .f(x) + ~ ~) + ~ ~2 .if(;) + ! ~3 .rj(:) + ..•. dx dx

(6.3]

Page 82: Hydrology and Sedimentology of Dynamic Rill Networks ...

73

,ry_A-) :Ir )- •-!!IJ& + l,_2l&)_ _ 1._3,I/(%) + ,.,. ..... "x ..... dx 2 ... dx2 6 ..... dx3 ..•. [6.4]

Subtracting Equation6.4 from 63 yields lhe central difference equation, such that:

ft.x+Ax) - ~Ax)= 2Ax ~) + O(Ax') [6.5]

where O(Ax3 ) represents the residual emir terms. Asswning third and higher order terms are negligible,

Equation 6.5 may be written as:

!!IJ& = f(x+Ax) - f(x-Ax) dx 2Ax

[6.6]

To solve Equation 6.1 numerically, each rill is discretized into .n increments (see Figure 6.1). Using

Equation 6.6, Equation 6.1 may be written as:

E' +It. r •

[6.7]

or

qx+l>z = ~ • ../ E' + It.) + q,..,,, I ~, r I I [6.8]

The Crank-Nicolson (Gerald, 1980) form of Equation 6.8 is:

[6.9]

which will be used to rout sediment under detachment limiting conditions. However, due to the methodol­

ogy that will be used to predict deposition, during transport limiting conditions Equation 6.9 is rewritten as:

q,...""= ( E""" +E~· ";)+q' [6.10] I Ax\.:, I 6

6.2 INTERRILL EROSION

Interrill erosion is the detachment and transpon of soil particles to a concentrated flow network. Rain­

drop impact and overland flow are the detachment and transpon mechanisms. Many complicating facun,

such as depth of ponding, surface sealing and crusting, soil properties and rainfall intensity prevent accurate

description of the individual process composing interrill erosion. To circumvent describing complex indi­

vidual processes and their interactions, Meyer (1981) proposed a net interrill erosion equation, which pre­

dicts the interrill erosion delivery to the concentrated flow network. lnterrill erosion delivery is based on a

Page 83: Hydrology and Sedimentology of Dynamic Rill Networks ...

74

sz I x+dx x+2dx x+3dx x+4h I I I I I -

Q~ I I I I I I I I I I I I I I I I I I I I

»n»»J»»nJ»»»)nwJnwJ»

Figure 6.1. J?iscrelizalion of rill section flowing downbill

Page 84: Hydrology and Sedimentology of Dynamic Rill Networks ...

similar fonn (Elliot, et al, 1989) given as:

or

b A, D,=KJ SIA

'

b A, E.=K./ S1 -

• • L '

75

[6.11]

[6.12]

where Di is interrill erosion delivery per unit rill length and width (kg/m2 /s), K; is the interrill erodibility

coefficient (kg-s/m4 ), I is rainfall intensity (m/s), S1 is a fractional slope factor, b is a soil dependent expo­

nent, A; is the contributing overland flow area (m2 ), A, is the rill surface area over which the interrill sedi­

ment is distributed (m2 ), and L, is the rill length segment (m). The slope factor, Sr is defined as {Elliot, et

al., 1989):

-4mn(,) s,= 1.05 - 0.85 e [6.13]

where s is slope angle (degrees). S1 is 0.20 for a flat slope, 1.00 for a slope of 45 degrees, and bounded by

I.OS for a slope of90 degrees. As shown in Figure 6.2, Meyer (1981) found the ~nent b varied with the

primary clay fraction of the soil(< 2 µm), and proposed that for clay fractions between 0.20 and 0.50 the b

exponent could be approximated by:

b= 2.1-0., [6.14]

where Od is the clay fraction. For soils with clay fraction less than 0.20, Meyer (1981) proposed a "b"

exponent of 2.0.

6.3 RILL EROSION

Rills develop from the concentration of overland_ flow, and transport interrill sediment as well as erod­

ed material detached from within the rill. Soil detachment in rills results from excessive shear forces pro­

duced by concentrated flow, sidewall sloughing, and head cut advancemenL Head cut advancement, due U>

its complexity and the lack of available physically based models, is neglected. Vertical sidewall sloughing

is assumed U> occur on rectangular rills during lateral expansion, and will be discussed in a subsequent sec­

tion. Rectangular rill cross sections are assumed in order to incorporate the Foster and Lane channel erosion

model (Foster, 1982).

6.3.l Foster and Lane Model Development

Page 85: Hydrology and Sedimentology of Dynamic Rill Networks ...

76

2.4

" .. REPLICATED

• NOT REPLICATED

2.2 " • " " ,Q

" " J " . a u " a 8. .= 2.0 - " -1: " ., a • -

1.8

• "

1.6 ..... ~~~---r~~~~...-~~~---,~~~~...-~~~ ......... ~~~~~ 0.0 0.1 0.2 0.3 0.4 0.5 0.6

Clay Fl'acUou of Soll

Figure 6.2. Soil clay percent effect on interrill erosion ~t b for several soils and aopping conditions

(Meyer, 1981).

Page 86: Hydrology and Sedimentology of Dynamic Rill Networks ...

77

The Foster and Lane channel erosion model (Foster and Lane, 1983; Foster, 1982; Foster et al., 1980;

Lane and Foster, 1980) was inCOl)lOl'llled into CREAMS for descnl>ing ephemeral gully growth in a tilled

agricultural field. Although their model was initially developed for ephemeral gully erosion, it should be

applicable to rill erosion as well

The Foster and Lane model assumes a steady state flow rate, and separates channel development into

two distinct stages. During the initial stage the channel bottom erodes uniformly downward at a width

dependent upon flow rate and soil properties. If flow conditions remain constant an equilibrium width is

achieved. During the second stage of development when the channel bottom reaches a non-erodible layer,

the channel expands laierally causing sidewall sloughing. This laleral expansion continues until a final

width is reached, afier which soil detaehment ceases. Their model estimaies the maximum potential erosion

in a rill, therefore the effects of sediment load and transport capacity must be accounled for. The following

sections descnbe the development and implementation of the Foster and Lane model

63.1.1 Stage 1 Development: Erodible Channel Bottom

During the initial stage of rill erosion, the channel bouom erodes downward at a constant rate. For a

constant flow rate, the rill shape is assumed to reach an equilibrium width if the nonerodible layer is deep

enough. From the initial point of rill detachment, this equilibrium width is assumed.

6.3.1.1.1 Detachme,u Rate

The rale of poiential rill detachment is assumed to be based on shear excess, written as (Foster, 1982):

D =K('t-'t)~ ,. r c [6.15]

where D., is the maximum polential rill detachment rate per unit rill length and width (kg/m2 /s), K, is a soil

erodibility factor for rill flow (s/m), 't is the shear stress acting at a point along the rill boundary (Pa), 't, is

the soil's critical tractive force (Pa), and P is a constant From rainfall simulator studies, Meyer et al

(1975a) performed a regression analysis and estimaled the exponent P to be 1.17 and 0.93 for bare soil and

canopy cover, respectively. However, Foster (1982) indicated that an unpublished rill erosion study with

simulated rainfall of 64 mm/h, found a p exponent of 1.05. 11te value of the exponent p varies widely

depending on the estimated critical tractive force 't,. Therefore, a P exponent of 1.0 is assumed for rill

detachment

6.3.1.12 Shear Stress Distributwn

Page 87: Hydrology and Sedimentology of Dynamic Rill Networks ...

78

In order to apply Equation 6.15, the shear stress distribution along the rill boundary must be known. In

order to develop explicit rill growth relationships, a shear stress distnbution is assumed. The original model

formulation used in CREAMS, applied an equation fitted to observed shear disuibutions in smooth chan­

nels (Graf, 1971), which was given as:

't. =exp[0.127 -0.516 Inx. -0.408 (lnx./-0.0344 (lnx.)J x. ~0.02 (6.16]

0.13x. 't, = 0.02 x, < 0.02 (6.17]

where 't;=t/t, 't is shear stress at x. (Pa), tis the average shear stress for a specific cross-section {Pa}, and

x, is the distance from the water surface to the point of interest along the wetted perimeter divided by the

wetted perimeter, WP. The average shear stress tis defined as:

t=yRS (6.18]

where y is the specific weight of water (N/m3 ), R is the hydraulic radius (m}, and S is the slope of the rill

{m/m). An alternative equation for shear distribution was introduced by Foster and Lane (1983), such that:

(6.19]

Equations 6.16 and 6.17 and Equation 6.19 both produce a maximum shear stress of l.35t at x, =0.5.

Both equations have the five following properties: 1) 't• (x. =O)=O at the water surface, 2) 't,(x. =0.5) is a

maximum at the rill center, 3) the slope of the shear stress distribution is zero at x. =0.5, which produces a

smooth transition for symmetrical boundaries, 4) the shear stress distribution function is symmetrical about

x, =0.5, and 5) the integral of the shear stress distribution around the wetted perimeter equals the average

shear stress t.

These equations yield very similar results. However, the form of Equation 6.19 has two distinct aiivan­

tages. F'lrst, the entire shear stress distribution can be represented by one equation. Second, shear stress dis­

tributions as well as the maximum shear stress vary with width to depth ratio, which can easily be account­

ed for in Equation 6.19 by varying the 1.35 constant and the 2.9 exponenL Since Equations 6.16 and 6.17

cannot be modified as easily, Equation 6.19 is used to define the shear stress distribution.

6.3.1.1.3 Equilibrium Rill Characteristics

As stated previously, for a constant llow rate the rill is assumed to have a rectangular cross section

with an equilibrium width. Prior to reaching a nonerodible layer, the rill bottom is assumed to have a para-

Page 88: Hydrology and Sedimentology of Dynamic Rill Networks ...

79

bolic shape, which is asswned to move vertically downward at a unifonn rate. Thereby implying that the

vertical erosion vector, Dm , is constant along the channel bottom. The vertical erosion vector Dm is

asswned to be the maximum detachment rate, which occurs at the middle of the rill at x. =0.5. However,

the governing erosion vector is normal to the rill boundary. As shown in Figure 6.3, the normal erosion

vector, D0 is given as:

D cos9=-·

D .. Substituting Equation 6.15 and 6.19 into Equation 6.20 yields:

't. - t. cos9= 135-; ...

where t,, is the dimensionless critical shear siress equal to tjt.

(6.20)

(6.21]

In order to develop normalized relationships describing channel equilibriwn characteristics, Equation

6.21 was solved numerically by Foster and Lane (1980). Defining the independent variable x-. as the dis­

tance along the wetted perimeter where t,, occurs, the following steps were used to develop normalized

relationships (Foster, 1989):

STEP 1: For an asswned x •• , half the parabolic rill bottom length was defined as 0.5-x-. . This length

was divided into 50 segments of equal length, i.e. define dx=(0.5-x-. )/50.

STEP 2: Using x.,, t., was calculated from Equation 6.19.

STEP 3: Starting with the first boundary segment between x. =0.5 and x. =0.5-dx, 9 was calculated at the

end points of the segment using equation 6.21.

STEP 4: The segment angle e was defined as the average angle between the 9 obtained from the segment

endpoints.

STEP 5: Given the segment angle, e and the segment length dx, the vertical and horiwnllll coordinates of

the segment endpoint were calculated. The next segment was evaluated by repeating steps 3-5

STEP 6: From the computed channel boundary dimensions found in step 5, R. , W. and D. were calculated

as defined in the following paragraph.

STEP 7: Assuming a new x.0 and steps 1-6 were repeated.

Page 89: Hydrology and Sedimentology of Dynamic Rill Networks ...

cos 9 =

Figure 6.3. Erosion vectllrS along a rill boundary.

I I I I

Rill Center Line --1

Iii I BollDlary

I*= 0.5

I I I

80

Page 90: Hydrology and Sedimentology of Dynamic Rill Networks ...

81

Figure 6.4 illustrllleS the above procedure for defining the norrnaliz.ed parameiers R. , W. and D. along

the channel boundary. The normalized hydraulic radius, R •• is the hydraulic radius divided by the wetted

perimeter, WP, w •. the normalized rill width, is the rill width divided by WP, and the normalized depth, n •.

is the maximum flow depth divided by the wetted perimeter. Using a computer program obtained from Dr.

G. R. Foster, normalized equilibrium rill characteristics are illustrated in rtgUrC 6.5 as a function orx ...

In order to define the equilibrium rill characteristics from Figure 6.6, a x .. value must be specified.

Foster et al (1980) developed a conveyance function g(x-.) to relate x .. to specific flow conditions and rill

properties. In developing the g(x .. ) function, Manning's equation was written as:

[6.22]

where Q is flow rate (m3 /s), n is Manning roughness coefficient (s/m1fl ), R is the hydraulic radius (m), S is

the channel slope (m/m), and A is the cross-sectional area of flow (m2 ). From the expressions R=A/WP

and R. =R/WP, the flow area• A• may be defined as:

Rz A=­

R.

Substituting Equation 6.23 into Equation 6.22 yields:

Using the definitions 't,,=tfi and i=-,Rs, the hydraulic radius R can be written as:

't R=--=--

1 S 't ..

Substituting Equation 6.25 into Equation 6.24 and rearranging yields,

or

[6.23]

[6.24]

[6.25]

[6.26]

[6.27]

The conveyance function, g(x., ) was solved using the left side of Equation 6.26, the shear stress distri­

bution Equation 6.19, and R. found from Figure 6.5, for x .. values ranging from 0.0 to0.5 (see Figure 6.6)_

Using this conveyance function g(x., ) relationship and Figure 6.6, x-. may be found given rill slope, soil

Page 91: Hydrology and Sedimentology of Dynamic Rill Networks ...

1. = 0 852 = goo

Iz: I1 + COS 9 di. y2 = y1 + sin H di. 8 = 82

w. 2

I I I I

I D. I I I

1., = 0.5 0, = o I1: 0 Y,= 0

Figun: 6.4. Normalized rill equilibrium characleristics numerical approximation description.

82

Page 92: Hydrology and Sedimentology of Dynamic Rill Networks ...

83

0.8 0.16 ------0.7 ~ 0.14

0.6 0.12 II: ,;

'\, :I .. 0.10 :;

• 0.5 ill = 0 .,, ' -- -; .. \, 1o.4 • 0.08 ..

.!! \ i -• \ I! .,, = ;! 0.3 \ 0.06 --• I!

\ 0 \ :;c

0.2 \ 0.04 \

\ Width \

\

0.1 ----- Hydraulic Radiu• \ 0.02

\

~ o.o 0.00

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

Normalized CrlUcaJ D'9t.ance, X

Figure 6.S. Normalized equilibrium characll:ristics for an eroding rill.

Page 93: Hydrology and Sedimentology of Dynamic Rill Networks ...

84

critical tractive force, Manning's n, and flow rare. As shown in Figure 6.6, g(x •• ) is a double function, i.e.

two values of x-. exist for a given g(x-. ). Foster (1989) suggested that the x-. value which produces the

widest channel should be used. The widest channel occurs for small x-. values, and therefore the left por­

tion of Figure 6.6 is used to fmd x-. .

63.1.1.4 Rill Equilibrium Conditions

For specified flow and soil parameters, nonnalized rill characteristics are found using Equation 627

and Figures 6.5 and 6.6. Actual rill dimensions are still unattainable because the wetted perimeter is

unknown. In order to define the welled perime1er, Foster et al. (1980) assumed a rectangular channel cross

section. Figure 6. 7 illustrates the equilibrium rill geometry for stage l rill development.

For a rectangular cross section, the welled perimeter WP is defined as:

WP= W ,q + 241 (628]

where W "'! is the equilibrium rill width, and dr is flow depth (m). In order to defme W "'! , an equation

describing the flow depth must be found. Expanding the definition of the hydraulic radius for a rectangular

rill yields:

Solving for dr yields:

RW d= .. I W -2R

,q

Substituting Equation 6.30 into Equation 6.28 yields:

Rearranging Manning's equation (Equation 6.22), and substituting A=R WP yields:

nQ =WPKfJ {s

Next, if R=R• /WP then R"'=WP"'R';', and substituting into Equation 6.32 yields:

nQ = WP813 RSfJ {s •

Rearranging yields:

(629]

[6.30]

[6.31)

[6.32]

[6.33]

Page 94: Hydrology and Sedimentology of Dynamic Rill Networks ...

14

12

" ,"' --------- ,, -----------2

85

I

I I

I J

/ .,,,,

o'r-~"""T~~-r-~--,,--~-.-~~T"'""~--.-~~-.-~~~~~~--J

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

Normailzed Cr!Ucal Distance, X-..c

Figure 6.6. Conveyance function g(x..) for an eroding channel at equilibrium.

Page 95: Hydrology and Sedimentology of Dynamic Rill Networks ...

86

I -·--leq--•-'I 1.: 0 ---t--:lt.....-------1--

-

dr dch

t i t

777777h»»»»»h»»» Nonerodible Layer

Figure 6.7. Equilibrium rill geomeuy fa stage 1 rill developmcnL

Page 96: Hydrology and Sedimentology of Dynamic Rill Networks ...

However, WP=Woq/W•, which yields:

3/1 WP= [ nQ ] R"S/1 {s •

3/1 W = [~] W.,R·S/1

"' {s •

87

[6.34]

[6.35]

Based on Equation 6.19 for shear distribution, the maximum rill detachment rate per unit rill length and

width, D,. (kg/m2 /s) is:

D =K (1.3~-'C) ... ' c

and the maximum rill detachment per unit rill length, E,. (kg/m/s), is

E =W D ... "' ... The maximum downward rate of movement of the rill, M,. (m/s), is:

D M =~

"' p•

where p. is the soil bulk density (kg/m3 ). The actual rate of downward movement is:

D M=-'

' p•

[6.36]

[6.37]

[6.38]

[6.39]

where M, and D, are the acwal detachment rate and downward movement rate. The hydraulic radius R is

found by rearranging Equation 6.24 such that:

R=[~J311 {s . [6.40)

As stated previously, the average shear stress Tis calculated from T=,yRS. Substiwting into Equation 6.40

one obtains:

[6.41]

6.3.1.2 Stage 2·Development: Nonerodible Channel Bottom

The second stage of rill development occurs when the rill reaches a nonerodt"ble layer. As illustrated in

Figure 6.8, once the nonerodible layer is reached, the rill expands laterally causing vertical sidewall slough­

ing. This lateral expansion continues at an exponentially decreasing rate until a final width is reached, after

Page 97: Hydrology and Sedimentology of Dynamic Rill Networks ...

88

Ir

1· f eq -1 • ......_.

1 ......_.

- dne ~ - ......_.

dr j Nonerodible Layer

Figure 6.8. Equilibrium rill geometry for stage 2 rill development.

Page 98: Hydrology and Sedimentology of Dynamic Rill Networks ...

89

which soil detachment ceases.

Due to the parabolic shape of the rill bottom, when the rill center reaches the nonerodible layer x.. is

still above the rill comer. This rill comer is defined as x.,, As shown in Figure 6.8, the area contained

between the parabolic and rectangular rill bottom is neglected. Although this assumption introduces emir,

its impact should be minimal.

After reaching the nonerodible layer, soil detachment is related to the shear excess located at x.., . Soil

detachment continues until 't,• equals 't-,. As the rill widens, shear stress at x.,, decreases as the flow depth

decreases. Although the shear stress 't exceeds 't, only along a small portion of the boundary between x..

and x'b , all soil directly above x,. is assumed to vertically slough as the soil material below erodes. It

should also be noted that this soil detachment is occurring on each side of the rill, therefore rill detachment

per unit rill length is:

[6.42]

where t• is the shear stress at x,b (Pa). The shear stress t• occurs at the lower rill comer located at x. •• giv-

enas:

d d x - I __ ...,__ 'b- WP - W+2d

I

where Wis the current rill width at time t (m). Using Equation 6.19, t• is calculated as:

and

From Equation 6.42, the rate of channel widening after reaching the nonerodible layer is:

[6.43]

[6.44]

[6.45]

[6.46]

The rate of channel widening decreases with time due to the reduction in shear stress as flow depth decreas­

es. In order to describe this widening process, Foster (1982) performed a regression analysis 10 find the

relational form. Erosion rates were calculated for varying values of Q, n, S and 't,, and nonnalized using:

W-W. W

,_ .. • - w

J-Wi,. [6.47]

Page 99: Hydrology and Sedimentology of Dynamic Rill Networks ...

t(dW/dl),. t.= W-W

I ;,.

90

[6.48]

where W., and (dW/dt)., are the initial rill width when the rill first encounters a nonerodible layer, respec­

tively, W1 is the final rill width, and W.' and t. are the nonnalized width and time, respectively. Rearrang­

ing Equation 6.47, the rill width at time tis:

W=W.'(W1-W,.)+W,. [6.49]

Foster (1982) found that the time rate of change ofW.' could be expressed as:

dW.' ·1.<>41. --=e

di [6.50]

Integrating Equation 6.50 with respect to time yields:

1 [ ·l.041,J W'=-- 1-e • 1.04

[6.51]

Foster (1982) pointed out, however, based on the definition of W.' the coefficient 1.04 should theoreti­

cally be 1.00, since·W.'=1 as,-. Therefore, Equation 6.51 reduces to:

[6.52]

At the time when the rill encounters the nonerodible layer, the initial change in rill width, (dW/dt)., , is

found using Equation 6.46. Using Equations 6.52, 6.47 and 6.48, the rill width over time may be found, and

the corresponding potential side wall sloughing rate is:

[6.53]

where & is the time interval (s), and ct.. is rill depth (m). This procedure, however, is only applicable

under steady state now conditions. In order to consider transitional nows, the rill width may be estimated

using Equation 6.46 and a forward difference approximation such that

IV""'= w' + dW & dt

[6.54]

In order to bound Equation 6.54 the final rill width, W 1 , must be known. The final rill width occurs

when t,. equals t ••• and thus x •• would correspond to x ... This value of x.. is defined as x'd. Since Lot is

at the rill bottom:

Page 100: Hydrology and Sedimentology of Dynamic Rill Networks ...

or

For a rectangular channel, the wetted perimeter corresponding to the final width is:

WP=W1+2d

1

Substituting Equation 6.56 into 6.57, and rearranging yields:

w ...:.L=l-2x WP 'ef

For the final width condition, the hydraulic radius, R, may be written as:

or from Equation 6.55,

R= dfWf WP

Since R, =R/WP, Equation 6.60 may be written as:

. 3_= R, WP x,ef

Substituting this into Equation 6.58 and rearranging yields:

R. = x,c/(1 - 2x,c/)

Substituting this R, definition into the conveyance function shown in Equation 6.26 yields:

yS (5LJ311 ,: {s •

91

[6.55]

[6.56]

[6.57]

[6.58]

[6.59]

[6.60]

[6.61]

[6.62]

[6.63]

This implicit relationship can be solved iteratively for x.q- Rearranging Equation 6.63, and using the rela­

tionship ,:., =t fi and ~S yields:

[6.64]

From Equation 6.60, the final width may be written from Equation 6.64 as:

Page 101: Hydrology and Sedimentology of Dynamic Rill Networks ...

92

W = [~ I - x.1]318 I ...fs S/3

x.ef [6.65]

6.3.2 Transitional Flow

The Foster and Lane model, described in the previous section, assumes steady state flow conditions.

Therefore, during flow rate changes their steady state erosion model is not applicable. To reconcile Ibis

problem, a procedure is developed to represent rill geometty and detachment rates during this transitional

phase.

Figure 6.9 illustrates rill shapes resulting from several transitional flow scenarios. At time t1 , a steady

state flow rate of Q1 , results in a rectangular rill with a vertical detachment componenL During time t2 , the

new flow rate Q2 is less than Q1 , which creates a narrower width channel being fonned. As the flow rate

increases during t3 • the flow rate Qs is greater than Q2 but less than Q1 • Since Qs is greater than ~ the rill

expands laterally until a new equilibrium width is reached. Since Q3 is less than Q1 the current rill width is

still less than the width at t1 • After the,new equilibrium width is reached at r.. , the rill continues to propa­

gate vertically. If during ts the new flow rate Qs equals the original flow rate Q1 , the rill expands laterally

until reaching the original rill width. After which during time 1(1 , the downward movement of the rill con­

tinues. If the flow rate Qi at t7 increase further, lateral rill expansion continues until a new equilibrium

width is reached, after which vertical detachment continues.

For the simplistic examples illustrated in Figure 6.9, there were never more than two channel widths.

However, when routing a typical runoff hydrograph through a rill, multiple channel widths occur. When

applying these techniques to a rill networlc, the amount of computer storage required to descnl>e the rill net­

worlc geometty would become prohibitive. Therefore, in order to reduce storage requirements, each rill

segment is represented by a maximum of two widths, W 1 and W 2 and two rill depths 0 1 and 0 2 (see Figure

6.IOa). When a third tier is created, the upper two tiers are combined, resulting in a two tier rill geometry.

Figure 6.IOb illustrates the reduction of a three tier rill geometry into a two tier system. Assuming the rill

bottom of the second tier remains constant, the two tiers are combined by equally dividing the bench of soil

created by the upper two tiers. The area of bench is equated to the new channel width by:

d(W +Wz} 2 1 =11.W(d +d\ 2 I 2' [6.66]

Solving for II. W yields:

Page 102: Hydrology and Sedimentology of Dynamic Rill Networks ...

93

.-_ -. -L ....__ __ -"- .J

l2: Qz< Q. t,: Q, = Q5

I I I I

I ' ' I L.. - ..,

•11111• L. - - - - - ..J

!, : Q,> Q 2 < Q. t 1 : Q? > Q.

I I · -- -· ~- •- -· :;=I I~ ·- - I I I -- -L- _J

t,:Q,=Qi t, : Q, = Q?

I ' I I I ~----' •111111• 1...-------....1

Figure 6.9. Rill geometry changes due to flow rate changes.

Page 103: Hydrology and Sedimentology of Dynamic Rill Networks ...

where AW is the change in rill width (see Figure 6.10). The new rill dimensions are:

d/=d, +d2

w,· = w2 + 2aw

94

(6.67]

(6.68]

[6.69]

where di' and W,' are the new upper tier depth and width, respectively. After conversion, the bottom depth

and width are defined as 11{ and W,', respectively.

To illustrate a more complicated example, Figure 6.11 gives an actual and approximate rill shape

resulting from five decreasing flow rates. The shaded region, shown in the approximated rill shape at '5 ,

illustrates the soil mass that is relocated to allow a two tier system. Since a small change in flow rate causes

a subsequent width change, tier aggregation is only allowed when the current flow rate exceeds a specified

threshold. The threshold is defined by a previous flow rate multiplied by a predefined fraction, AQr This

not only increases model efficiency, but also prevents aberrations that may resulL

An inconsistency exists with a two tier rill geometry when an increase in flow rate results in a flow

depth that exceeds the second tier depth, d2 (see Figure 6.12). The steady state erosion model estimates

shear excess for a rectangular cross section, and developing shear excess equations to represent a two tier

geometry introduces inordinate complexity. Therefore, as illustrated in Figure 6.12, the second tier width,

W2 , is assumed to have infinitely high sides. Although this assumption initially results in an excessively

high shear excess, as the rill width expands laterally shear excess estimates decrease rapidly. During lateral

rill expansion under transitional flow conditions, the detachment rate is found using:

D =2K (t -t) l"C' , b c

dW D,.. Tt=-,;:-

w"'"' = w' + dW 61 dt

6.3.3 Sediment Load and Transport Capacity Interactions

[6.70]

[6.71]

(6.72]

[6.73]

Page 104: Hydrology and Sedimentology of Dynamic Rill Networks ...

i------- 11 ------

,. 12--

, . i,.----li-----i -, I I

I

D1 .

i-~-n' 12

Initial Rill Boundary - - - - Final Rill Boundary

r-I I . . I I

_J I I 1

Al1

AI

Al = Dz l I, - 11 I 2lD1+ Dz l

D; = Da+ Dz Ii = lz+ 2AJ

95

Figure 6.10. Rill geometry repreaentalion showing (a) a lWO tier rill, and (b) the ledncrioo of the three tier

geometry to a two tier system.

Page 105: Hydrology and Sedimentology of Dynamic Rill Networks ...

96

Actual Rill Shape

11 lz

t3 t,

l5

!ppro1imated Rill Shape at 15

Figure 6.11. Multiple rill tier mluction resulting from five decreasing flow lllleS.

Page 106: Hydrology and Sedimentology of Dynamic Rill Networks ...

97

D 1

D2 ~

-=

! ~ l Q = 1 1/1 IN

I I IV I v

I I I I I I I I

Q = 4 I/IIN

Figure 6.12. Resulting rill geometry for calculating shear excess from a two lier sys1e111 with increasing

flow rate.

Page 107: Hydrology and Sedimentology of Dynamic Rill Networks ...

98

Meyer and Monlce (1965) observed that Equation 6.1, the steady state sediment continuity equation, is

inconect since it neglects the interaction between sediment and transport capacity. Foster and Meyec

(1972a, 1975) proposed that rill detachment and deposition are proportional to the difference between

transport capacity of the flow and sediment load, yielding a diffusion type equation:

D=C(T-q/W) r 1 c • [6.74]

where C1 is a first-order reaction coefficient (m·t ), and T. is the transport capacity of the flow (kg/s/m rill

width).

6.3.3.1 Detachment Limiting Case

For the detachment limiting case when transport capacity exceeds sediment load, Foster and Meyer

(1972a, 1975) rearranged Equation 6.74 to gee

[6.75]

Foster and Meyer (1972a) then proposed that maximum potential detachment rate, D,., , is proportional to

the transport capacity, such that:

D =CT re l c [6.76]

where C1 must be less than or equal to zero. Substituting Equation 6. 76 into 6.75 yields:

D /W T qi -+--=! D T

[6.77] re c

and rearranging yields:

[ q,/W]

D =D 1---, TC T

c

[6.78]

or

[ q,/W]

E =E !---, "' T

c

[6.79]

Equation 6. 79 is only used to determine the actual rill detaclunent, and does not alter the rate of side wall

sloughing during lateral rill expansion during stage 2 development and transitional flow conditions when

Equation 6.77 indicates that when the sediment load, q. , equals zero, rill detachment, D,, equals the

detachment rate capacity, D,., . Conversely, when D, equals zero, the sediment load equals the transport

Page 108: Hydrology and Sedimentology of Dynamic Rill Networks ...

99

capacity, T •. It is important to note, however, that the acwal detachment rate E,. is bounded by T,-q/W,

the maximum transport capacity available for detachemenL These concepts are concepwally justified by

the stream power concept (Foster and Meyer, I972a). Any specified flow contains a finite amount of ener­

gy, which may be used for detaching and/or transporting soil particles. The D, ID., term in Equation 6.77

represents the relative amount of energy expended on soil detachment, and the term q, rr. indicates the rel­

ative amount of energy expended on sediment transport The sum of these two terms equal unity, the total

relative available energy.

6.3.3.2 Transport Limiting Case: Deposition Effects

For transport limiting flow, deposition of eroded sediment has a major impact on sediment yield.

Deposition occurs when the flow does not contain sufficient energy to transport eroded sediment, which

results from either a decrease in transport capacity or an influx of additional sediment

Deposition of eroded sediment is a selective process, which varies with particle si7.e distribution. Coar­

ser particles are more likely to deposit, due to their higher setting velocity (Foster, 1982). For discrete

quiescent settling, the settling velocity of an individual particle may be determined using Stokes law, writ­

ten as (Barfield et al., I 981 ):

V = -1 [ tfg (S - l)]

• 18 v • [6.80]

where V, is panic le settling velocity (m/s), dis particle diameter (m), g is acceleration of gravity (m/s2 ), v

is kinematic viscosity (m2 /s), and S8 is particle specific gravity. Equation 6.80 is only valid up to a Rey­

nolds number, Re, up to 0.5 (Barfield et al., 1981), which is given by:

Vd R =-'­• v

[6.81]

In addition, Equation 6.80 assumes spherical particles settling in turbulent free water. Other factaS govern­

ing settling velocity of discrete particles are: particle shape, aggregation, turbulence, and flocculation (Bar­

field et al., 1981). When the Reynolds number exceeds 0.5, Wilson, eL al (1983) developed the following

equation:

log10(100 V,) = --0.34246272(log10(1000 dl]2 + 0.98912185 log10

(1000 d) + 1.146128 [6.82]

To account for specific gravity, the settling velocity, V,. is multiplied by the ratio (S,-1)/1.65.

Page 109: Hydrology and Sedimentology of Dynamic Rill Networks ...

1()()

An appreciable number of studies on deposition processes in reservoirs and detention basins have been

performed and many of the results are applicable to deposition in rills. Camp (1946) formulaled sediment

trapping efficiency theory under quiescent settling conditions for an idealired rectangular continuous flow

basin, by assuming: (a) horizontal flow with a uniform velocity distribution, (b) uniform vertical sediment

concentration at the inlet, (c) discrete particle settling, and d) no resuspension of sedimenL The fraction of

trapped sediment, F, for panicles having a settling velocity V, (m/s) is expressed in the overflow rate equa­

tion as (Camp, 1946):

v F=-'-

QIA (6.83]

where Q is flow rate through the basin (m3 /s), and A is the basin subsurface area (m2 ). Equation 6.83 indi-

cates that panicles having settling velocities greater than or equal to the overflow rate, Q/A, will be

trapped.

Using the overflow rate equation, Dobbins (1944) and Camp (1946) developed and numerically solved

an analytical function relating sediment trapping efficiency, F, and the parameters V, A!Q and V, D/2£.,

where D is flow depth (L) and e is the turbulent eddy diffusivity (L2 fr). In addition, Camp (1946) found

that e was constant for a parabolic velocity distribution, given as:

e=0.015DU. (6.84]

where U. is the boundary shear velocity (Lfr). The numerical solution is shown in Figure 6.13, which was

experimentally validated by Dobbins (1949). Figure 6.13 illustrates turbulence effects on basin trapping

efficiency. For high turbulence conditions when e is large, the fraction of sediment trapped is reduced.

However, when turbulence decreases and e is small, the fraction of trapped sediment increases.

Although these relationships were developed for rectangular basins, the concepts may also be applied

to concentrated flow conditions. The overflow rate Equation 6.83 for quiescent settling conditions can be

rewritten for rill flow as:

vx F=-'- (6.85]

q

where q is flow rate per unit rill width (m2 /s), and X is distance downstream from a vertically unifcxm sed-

iment concentration. Figure 6.14a illustrates the path two panicles take under quiescent settling and a uni­

form velocity profile. The particle located at the water surface travels twice the distance downstream to

deposit compared to the panicle at half the flow depth. Extending this same concept to a vertically uniform

Page 110: Hydrology and Sedimentology of Dynamic Rill Networks ...

101

0.9 2.0

1.5

LL. • 0.7

0 ILi > 0.6 0 ::e 0.8 ILi LL. a:: 0.5 0.7 0

z 0.6: 0 ::::, I- 0.4 0.5 _., (.) < < 0.4 > a:: LL. 0.3

0.3

0.2 0.2

0.1 0.1

O.":-~::-'-::'-:~::'-L..1-~--!~L-J..U.J.JJ..L...__J~ 0.1 0.2 0.4 0.6 I . 2 4 6 10 20

Vs D / 2€

Figurc6.13 Sediment uapping efficiency for a rectangular reservoir with steady stale 1Dliform llow (Camp,

1946; Dobbins, 1950).

Page 111: Hydrology and Sedimentology of Dynamic Rill Networks ...

102

sediment concentration with all particles having a settling velocity V, , all particles are deposited at a dis­

tance of X2 for X2 =g/V, . Half the particles settle at a distance X1 =X2 /2, which indicates 1he fraction of

deposited particles increases linearly downstream.

As stated previously, Foster and Meyer (1972a, 1975) proposed that deposition could be described by

Equation 6.74, which states that D ,=C1(T,-q, /W). During deposition, the coefficient C1 varies with particle

size. To date there are no validated relationships describing the phenomena. However, based on results by

Einstein (1968), Foster and Huggins (1977) and Davis (1978), Foster (1982) proposed that for channelized

flow C1 could be represented by:

Substituting Equation 6.86 into Equation 6.74 yields,

v c =-'

I q

v D =_!.(T-q /W)

r q c •

[6.86]

[6.87]

Recalling that D, is the deposition rate per unit rill width and length (kg/m2 /s), the deposition rate per unit

length, E, (kg/mis) may be written as:

vx w E =-'-(T -q /W)-

r q c , L [6.88]

where Lis the total rill length (m), and V, X/q represents the fraction of deposited sediment at a distance X

downstream.

Equation 6.83 assumes (a) quiescent settling conditions in turbulent free flow (b) a uniform velocity

distribution, (c) no lateral sediment inflow, (d) one particle size with settling velocity, V, , (e) a vertically

uniform sediment concentration at X=O, and (g) the deposition rate decreases linearly downstream. Experi­

mental and theoretical studies do not support the linear decrease, but have shown that deposition rate

decreases exponentially with distance downstream (Einstein, 1942; Foster and Meyer, 1972a, 1975). How­

ever, due to the complexity of implementing these concepts, current erosion modeling efforts neglect the

exponentially decaying deposition rate, and assume excess sediment load is instantaneously deposited

(Nearing et al., 1989; Hirschi and Barfield, 1988a; Rohlf, 1981). In addition, quiescent settling in turbulent

free flow and a uniform velocity disttjbution are never achieved in field conditions. Studies have found

that the presence of turbulence significantly reduces the sediment trapping efficiency, i.e. the quantity of

deposited sediment (Chen, 1975; Brown, 1950; Camp, 1946; and Dobbins, 1944). For highly turbulent

Page 112: Hydrology and Sedimentology of Dynamic Rill Networks ...

q

q

Path of Particle 2 v.

Yz = Y1

r Y1 V,

------ lz = 211

Uniform Verticle Sediment Concentration

F = v. 11 = 0.5 I q

F - v. 11 - 1 0 z - q - •

1-4------ lz = 211

103

Figure 6.14. ~ particle settling in a rill for (a) two particles and (b) for a uniform sediment coocen­

ttation.

Page 113: Hydrology and Sedimentology of Dynamic Rill Networks ...

104

flow, Chen (1975) proposed that trapping efficiency may be approximated by:

F = 1 - exp(-V A/Q) • (6.89)

For fully turbulent rill flow, Equation 6.89 may be written as:

F = 1 -exp(-V,X/q) (6.90]

Figure 6.15 shows the fraction of particles deposited for fully turbulent flow and quiescent settling condi­

tions. As shown in Figure 6.15., applying Equation 6.88 under turbulent conditions would result in signifi-

cant errors.

Why then has Equation 6.88 been successfully applied to field conditions. The answer to this question

may be explained by how the equation was implemented. In most applications the rill was discretized into

discrete sections (see Figure 6.1), and the deposition rate q, (kg/s) was calculated for each section using:

VM q =-'-(T-q /W)W

d q C 6 (6.91]

where~ is the discretization length (m). It is important to note that q, must be bounded by WT,-q,, the

maximum available sediment load available for deposition. One of the underlying assumptions in the origi­

nal version of Equation 6.88 is that initially the upstream cross section has a unifonn sediment concentra­

tion in the vertical direction. Hence, Equation 6.91 also implies a vertically uniform sediment concentration

at the beginning of each discretization length, which is incorrect under quiescent settling conditions.

Since the results from Equation 6.91 are dependent on the selection of '1X, a series of simulations were

conducted for several values of~ by assuming no lateral sediment inflow and a constant transport capaci­

ty. The fraction of deposited sediment was defined as:

T -q (x)/W F= c I

T -q (x=O)/W c •

[6.92]

The sediment load at q. (x) decreases downstream since deposition from the previous upstream section.

The results of the simulations are shown in Figure 6.16 for the rill length subdivided into 50, 25 and 10 sec­

tions. Also shown in Figure 6.16 is the fraction deposited under fully turbulent flow (Equation 6.90) and

quiescent settling conditions (Equation 6.85). As illustrated in Figure 6.16, the fmer the discretization

length the closer Equation 6.91 approaches the fully turbulent conditions. In aclUal rills, settling conditions

are between fully turbulent and quiescent settling, depending on local conditions. Therefore, although

Equation 6.91 does not explicitly account for turbulence effects, its implementation approximates natural

conditions, and thus is an appropriate deposition relationship when applied properly.

Page 114: Hydrology and Sedimentology of Dynamic Rill Networks ...

105

1.or---~===========================i I 0.9

0.8

0.7

I&,

-go.a c:i. c:i.

f E'."0.5 i:: 0 :I 0 .,o.4

·t

0.3

0.2

0.1

I i ~

I I I

I I i

I I

Quiescent Settling F = VsX!q

Fully Turbulent Flow F = 1 - exp ( V5 Xlq)

0.0 .. ---.----.--.---,----r---.-----,---,I 0 100 15) 3ll 400

VsX/q

Figure 6.15. Fraction of deposited particles for fully lUlbulent flow and ideal quiescent flow conditions.

Page 115: Hydrology and Sedimentology of Dynamic Rill Networks ...

106

r-~~~~:::::====================i 1.0

0.9

0.8

0.7

Dr.

'g0.6 c:i. c:i. f

E-- 0.5 d ~··

~0.4

it 0.3

0 12)

Figure 6.16. Comparison of deposition equations.

-<l.liesoErll -10R>int -2>R>int. -BJR>int - R.11 ly TurbJIErll

100 100 210 2iO

VaX/q

Page 116: Hydrology and Sedimentology of Dynamic Rill Networks ...

107

6.3.4 Transport Capacity

Most current sediment transport theory was developed for stream flow conditions. However, Alonso et

al. (1981) found that the Yalin equation (1977) was appropriate to use on the transpOrt of light materials in

stream flow, and a range of particle sizes and densities in shallow flow typically encounten:4 in field condi­

tions. Therefore, the Y alin transport equation was selected to valuate the transpOrt capacity of rill flow

based on the evaluation by Alonso et al. (1981) and its current widespread use.

Yalin (1963) developed a bed load transport model for cohesionless grains of equal sim over a mov­

able bed. The model was developed based on dimensional analysis and the mechanics of average grain

motion for uniform turbulent flow with a laminar sublayer not exceeding the bed roughness. Yalin (1977)

presented the model as a series of equations, given as:

q ~ .:J/2 [ I ] $ = ' • = 0.635 s I " I - - In (I + as) (yd)312 as

' y

s=--1 Yer

a = 2.45 Y112.r°"4

CT I

p cl. Y=-·-

Y,d

(6.93]

[6.94]

(6.95]

[6.96]

where $ is a dimensionless transport rate number (originally defined by Einstein), q. is sediment transport

rate (N/m-s), P. is the density of water (kg/m3 ), Y, is the particle specific weight in watec (N/m3 ), d is

mean particle diameter (m), Y is a mobility number, Y" is the critical mobility number found from Shields

diagram, S1 is sediment specific gravity, and u. is bed shear velocity (m/s).

Redefining several terms, Finkner et al. (1989) presented the Yalin equation as:

~ = 2.45 S°OA f 112 6 I CT

y 6=--1

Yer

(6.97]

[6.98]

(6.99]

Page 117: Hydrology and Sedimentology of Dynamic Rill Networks ...

108

Y= 't P,.(S, - l)gd

[6.100]

where T0 is sediment transpon capacity (kg/s/m rill width}, 't is shear stress acting IO detach soil (Pa}, and 6

and ~ are dimensionless parameters.

The Y alin equation was developed for uniform panic le size. To predict transpon capacity for a sedi­

ment mixture, Foster and Meyer (1972b) modified the Y alin equation by distributing the flow's total trans­

pon capacity among the available sediment based on flow characteristics, and particle size and density. The

Yalin equation (Yalin, 1963) was developed by assuming the number of panicles in transp<lrt was propor­

tional to 6. Hence, Foster and Meyer (1972b) assumed for a sediment mixture that the number of panicles

of type i were proponional to 6,. The total 6 for the sediment mixture, T, was found by summing all 6,'s,

such thac

• T= ~), (6.101]

i-1

where n was the number of panicle types. For sediment panicles of type i in a mixture, the number of trans­

pon particles [Ne]; , is:

6. [Ne],=N,~

where N; is the total number of particles of type i for a uniform material.

[6.102]

The dimensionless transpon rate cjl was assumed proportional to the number of transported particles,

such thac

(6.103]

where [cjl.J; is the effective m for particle type i in the mixture and cjl1 is the cjl for particle type i with uniform

material. The potential transpon capacity for particle type i in a mixture, T q,1. was defined as:

[6.104]

The transpon capacity of each particle class was assumed to be represented by Equation 6.104 only if avail­

able sediment in all particle classes was in excess or deficit.

When there was excess transpon capacity in a particle class and a deficit in another, the transpon

capacity was shifted in order to use all existing transpon capacity. For these cases, Davis (1978) developed

Page 118: Hydrology and Sedimentology of Dynamic Rill Networks ...

109

a methodology for redistributing the transport capacity using the following steps.

1. When the sediment load for particle class i, <lsi. was less than or equal to T .,.(.qp .,), the required tranS­

port,., by the available sediment was found using Equation 6.97, such !hat:

'= q. ri s d.fu

,, • w

[6.105]

Thus, the actual transport capacity for particle class i, T ci , was:

(6.106]

2. For all particle classes where q,p:T .,.. the fraction of total transport capacity actually needed, SPT, was

found using:

[6.107]

where n .. was the number of particle classes where q., >Top; .

3. The excess sediment transport fraction, E,. , that was distributed among the transport deficit classes is:

E =1-SPT ,,, 4. For the transport deficit particle classes where q.. >Tcpi, the sum of the li,'s were calculaled as:

·-SDLT= L,li;

... ,

[6.108)

[6.109)

5. The excess transport capacity was distributed among then .. particle classes for q,; >Top; in proportion to

their individual Ii,, such that:

[6.110]

6. If after repeating steps 1-5 all n particle classes had q,; >To1, it was assumed that the proper transport

capacities were found.

7. If after repeating steps 1-5 all n .. particle classes had Q,; <T.;, the excess transport capacity was equally

redisbibuted among all n particle classes by:

• t SMUS= L ...!i..

... , 41; [6.111)

Page 119: Hydrology and Sedimentology of Dynamic Rill Networks ...

110

q.

T"= SM"us [6.112]

8. An additional step was added to the Davis (1978) method when qp"' for all particle siz.e classes. The

excess transport capacity was equally redistributed among all n particle classes using Equations

6.111 and 6.112. This resulted in a larger total potential ttansport capacity, which varied with

sediment load.

6.3 SUMMARY AND CONCLUSIONS

The development of a dynamic erosion model is presented that accounts for in11lrrill erosion, potential

rill detachment, deposition in rills, rill geometry, transpon capacity of the flow, and the interaction between

sediment load and transpon capacity. The structure of the erosion model is based on the steady state conti­

nuity equation, which is numerically approximated using finite difference methods.

Interrill erosion is approximated using a relationship introduced by Meyer (1981), and rill erosion and

geometry are based on the steady state erosion model of Foster and Lane (1983). In addition, a procedure

is presented that accounts for transitional flow rates. For detachment limiting conditions the interaction

between transpon capacity and sediment load is accounted for using the Foster and Meyer (1972a)

approach. Transpon limiting flow conditions with deposition are based on overflow rate concepts, and is

approximated using an expression presented by Foster (1981). Finally, ttansport capacity for rill flow is

approximated using the modified Yalin equation (Foster and Meyer, 1972b).

Page 120: Hydrology and Sedimentology of Dynamic Rill Networks ...

CHAPTER7

MODEL VALIDATION

111

The ultimate test of a mathematical model is the comparison of simulated and observed system respon­

ses to a variety of input conditions. To evaluate the dynamic erosion model, model predictions will be com­

pared to detailed rainfall simulator field studies. In addition, an extensive sensitivity analysis will be per­

formed to evaluate the entire range of conditions typically encountered in natmal systems. A detailed

description of the validation and sensitivity analysis is given in an additional report Volume IV, entitled

"Dynamic Erosion Model Validation" (Storm, et al., 1990).

7.1 FIELD DAT A COLLECTION

The rainfall simulator field studies used to validate the dynamic erosion model were performed at the

·"University of Kentucky's Coldstream Farm in Lexington, Kentucky. Two soil types were used; a McAfee

silty clay loam topsoil and a Maury silty clay loam subsoil. Three test plots per soil type were used, with

dimensions 4.57 m (15 ft) by 22.1 m (72.6 ft). The test plots were graded to a wtiform 8.7 peicent slope

·· with no appreciable cross-slope. Sample pits were constructed at the slope bottom to permit smface runoff

measurement and sampling.

Rainfall was applied at an intensity of 78 mm/hr using the Kentucky Rainfall Simularor (Moore et al.,

1983). Surface runoff was measured using a tipping bucket apparatus (Barfield and Hirschi, 1986). In addi­

tion, initial and final surface profile measurements were performed at 0.610 m (2 ft) intervals up-slope and

1.27 cm (0.5 in) cross slope using a modification of a surface profile meter described by Hirschi et al.

(1987). To document rill development and verify the surface profile measurements, the teSt plots were pho­

tographed at two minute intervals. Initial and final soil bulk density measurements were also taken using a

neutron probe on a 1.22 m (4 ft) grid.

Numerous laboratory tests were also run on bulk soil samples obtained prior to applying rainfall, and

on surface runoff samples taken at the plot bottom. Plasticity index, and aggregate and primary particle

size distributions were found for the bulk soil samples. In addition, eroded particle size distributions of the

matrix soil were found using the method by Barfield et al. (1981), as well as primary and aggregate particle

size distributions of the surface runoff samples.

Page 121: Hydrology and Sedimentology of Dynamic Rill Networks ...

112

To investigate surface roughness effects, two surface conditions per soil type were used. For the sub­

soil test plots, two roughness conditions were obtained by 1) rototilling the soil, and 2) by rototilling and

dragging the surface smooth. For the topsoil plots, the soil was 1) rototilled. and 2) disked up and down

hill with cross slope raking. For all surface roughness conditions the test plots contained no surface cover

and were fallow for at least one year.

7.2 MODEL TESTING AND SENSITIVITY ANALYSIS

In order to evaluate the acclD'acy of the erosion model, the runoff bydrographs will be visually opti­

mized with the observed runoff hydrographs. From the field studies, observed sediment yields, "steady

state" erosion rates and concentrations, and aggregate particle size distributions will be compared to the

response of the dynamic erosion model.

To evaluate the response of the dynamic erosion model under field conditions, a detailed sensitivity

analysis will be performed. By evaluating the effects of changing various input parameters. component pro­

cesses and parameter values having the greatest impact on model response will be identified. The first

series of sensitivity simulations will be performed for one top soil and one subsoil plot using observed field

conditions.

In order to evaluate the random surface generator and the impact of surface roughness, a second series

of simulations will be performed using stochastically generated random surfaces. The first set of simula­

tions will use the rill network found using the digital elevation model. For the same surfaces, a second set

of simulations will be performed without a rill network. Instead, a group of parallel equal-distant rills with

an equivalent rill density will be used.

Page 122: Hydrology and Sedimentology of Dynamic Rill Networks ...

113

REFERENCES Abrahams, A. D. 1984. Channel networks: A geomorphological perspective. WatJ:r Resources Research,

20(2):161-168.

Ackers, P. and W. R. White. 1973. Sediment transport: New approach and analysis. Joor. of Hydr. Div., ASCE 99 (HYl 1):2041-2060.

Alberts, E. E., J. M. Laflen, W. J. Rawls, J. R. Simanton, and M. A. Nearing. 1989. USDA·wala erosion prediction project: Hillslope profile model documentation; Chapter 6, Soil ComponenL NSERL Report No. 2. National Soil Erosion Laboratory. USDA-ARS, W. Lafayette, IN.

Al-Durrah, M. and J. M. Bradford. 1981. New methods of studying soil detachment due to watcnlrop impacL Soil Sci. Soc. Am. J. 45:949-953.

Al-Durrah, M. and J.M. Bradford. 1982. The mechanism ofraindrop splash on soil smfaces. Soil ScL Soc. Am. J. 46:1086-1090.

· Alonso, C. V., W. H. Neibling and G. R. Foster. 1981. Estimating sediment transport capacity in watershed modeling. Transaction of the ASAE, 24(5):1211-1220, 1226.

Barfield, B. J., R. I. Bamhisel, M. C. Hirschi and I. D. Moore. 1988. Compaction effects on erosion of mine spoil and reconstructed topsoil. Transactions of the ASAE. 31(2):447-452.

Barfield, B. J., R. I. Barnhisel, J. L. Powell, M. C. Hirschi and I. D. Moore. 1984. Erodibilities and eroded size distribution of Western Kentucky mine spoil and reconstructed topsoil. Symposium on Surface Mining, Hydrology, Sedimentology, and Reclamation. University ofKentucky,Lexington.

Barfield, B. J. and M. C. Hirschi. 1986. Tipping bucket flow measurement on erosion plots. Transactions of the ASAE, 29(6): 1600-1604.

Barfield, B. J., I. D. Moore and R. G. Williams. 1979. Sediment yield in surface mined wat.enhecls. Pro­ceeds, Symposium on Surface Mine Hydrology, Sedimentology and Reclamation. University of Ken­tucky, Lexington, KY. pp. 83-92. December.

Barfield, B. J., R. C. Warner and C. T. Haan. 1981. Applied Hydrology and Sedimentology for Disturbed Areas. Oklahoma Technical Press, Stillwater, OK, 74074. 603 p.

Bennett, J. P. 1974. Concepts of mathematical modeling of sediment yield. Water Resources Research. 10(3):485-492.

Bisal, F. 1950. Calibration of splash cups for soil studies. Agricultural Engineering. 31:621-622.

Borah, D. K. and P. K. Bordoloi. 1989. Nonuniform sediment transporL Transactions of the ASAE. 32(5): 1631-1636.

Bradford, J.M., D. A. Farrell and W. E. Larson. 1973. Mathematical evaluation of factors affecting gully stability. Soil Sci. Soc. Amer. Proc., 37( !): 103-107.

Bradford, J. M. and R. F. Piest. 1980. Erosional development of valley-bottom gullies in the upper mid­westem United States. IN: Thresholds in Geomorphology, D. R. Coates, and J. D. Vital<, eds. pp. 75-101.

Brakensiek, D. L. and W. J. Rawls. 1983. Agricultural management effects on soil water processes. Part II: Green and Ampt parameters for crusting soils. Transactions of the ASAE, 26(6):1753-1757.

Brown, C. B. 1950. Sediment Transportation. In Engineering Hydraulics, ed. by Hunter Rouse, John Wiley, New York.

Page 123: Hydrology and Sedimentology of Dynamic Rill Networks ...

114

Bubenzer, G.D. and B. A. Jones, Jr. 1971. Drop size and impact velocity effects on the detachment of soils under simulated rainfall. Transactions of the ASAE. 14(4):625-628.

Camp, T. R. 1946. Sedimentation and the design of settling tanks. Transactions of the ASCE, 111:895-958.

Carter, C. E., J. D. Greer, H.J. Brand and J.M. Floyd 1974. Raindrop characteristics in South Central United States. Transactions of the ASAE, 17(6):1033-1037.

Chen, C. 1975. Design of sediment retention basins. Proceedings National Symposiwn on Urban Hydrolo­gy and Sediment Control, UK BU 109, College of Engineering, Univenity of Kentucky, Lexington, KY.

Cruse, R. M. and P. B. Francis. 1984. Shallow-layer soil water potential changes due to waterdrop impact. Soil Sci. Soc. Am. J. 48:498-500.

Cruse, R. M. and W. E. Larson. 1977. Effect of soil shear strength on soil detachment due to raindrop impact. Soil Sci. Soc. Am. J. 41:777-781.

Curtis, D. C. 1976. A deterministic urban storm water and sediment discharge model In: Proc. of National Symposium on Urban Hydrology, Hydraulics, and Sediment Conservation.

Davis, S. S. 1978. Deposition of nonuniform sediment by overland flow on concave slopes. M.S. Thesis, Purdue University, West Lafayette, IN. 137 pp.

Diezman, M. M., S. Mostaghimi and T. A. Dillaha. 1987. Size distribution of eroded sediment from tillage systems. ASAE Paper No. 87-2068.

Dillaha, T. A. and D. B. Beasley. 1983. Distributed parameter modeling of sediment movement and particle size distributions. Transactions of the ASAE. 26(6):1766-1772, 1777.

Dobbins, W. E. 1944. Effect of turbulence on sedimentation. TRANSAcnONS of the ASCE, 109:619-678.

Edwards, W.R. and W. E. Larson. 1969. Infiltration of water into soils as influenced by surface seal devel­opment. Transactions of the ASAE. 12(4):463-465, 470.

Einstein, H. A. 1942. Formulas for the transportation of bedload. TRANSAcnONS of the ASCE, Vol. 107.

Einstein, H. A. 1950. The bed-load function for sediment transportation in open channel flows. USDA-SCS Technical Bulletin No. 1026, 78 pp.

Einstein, H. A. 1968. Deposition of suspended particles in a gravel bed. Journal of the Hydraulics Division, Proc. of the ASCE. 94(HYS):1197-1205.

Ekem, P. C. 1953. Problems of raindrop impact erosion. Agricultural Engineering. 34(1):23-25, 28.

Elliot, W. J., K. D. Kohl and J.M. Laflen. 1988. Methods of collecting WEPP soil erodibility data. ASAE Paper No. MCR 88-138, ASAE, St Joseph, Ml.

Elliot, WJ., A.M. Liebensow, J.M. Laflen, and K.D. Kohl. 1989. A compendiwn of soil erodibility data from WEPP cropland soil field erodibility experiments 1987 & 88. NSERL Report No. 3, Ohio Slate University & USDA-ARS.

Ellison, W. D. 1944. Studies of raindrop erosion. Agricultural Engineering. 25(4):131-136, 25(5):181-182.

Ellison, W. D. 1947. Soil erosion studies - Pan II: Soil detachment hazard by raindrop splash. Agricultural Engineering. 28(5): 197-201.

Page 124: Hydrology and Sedimentology of Dynamic Rill Networks ...

115

Ellison, W. D. and O. T. Ellison. 1947. Soil erosion studies - Part N: Soil detachment by surface flow. Agricultural Engineering. 28(9):402-405, 408.

Epstein, E. and W. J. Grant. 1967. Soil losses and crust formation as related to some soil physical proper­ties. Soil Sci. Soc. Am. Proc. 31:547-550.

Ewing, L. K. and J. K. Mitchell. 1986. Overland flow and sediment llllllSpOrt simulation on small plots. Transactions of the AS AE. 29( 6): 1572-1581.

Farmer, E. E. ad B. P. Van Haveren. 1971. Soil erosion by overland flow and raindrop splash on lhree mountain soils. USDA Forest Service Research Paper INT-100. Intermt. For. and Range Exp. Stat., Ogden, Utah.

Finkner, S. C., M.A. Nearing, G. R. Foster and J.E. Gilley. 1989. A simplified equation for modeling sed­iment transport capacity. Transactions of the ASAE, 32(5):1545-1550.

Foster, G. R. 1982. Modeling the erosion process. In: C. T. Haan (ed.). Hydrologic Modeling of Small Wat­ersheds. Monograph No. 5. American Society of Agricultural Engineers, St. Joseph, MI.

Foster, G. R. 1986. Understanding ephemeral gully erosion. Soil Conservation, Assessing the National Resources Inventory, Volume 2. National Academic Press, Washington, D.C.

Foster, G. R. and L. F. Huggins. 1977. Deposition of sediment by overland flow on concave slopes. In: Soil Erosion Prediction and Control. Special Publication No. 21. Soil Conservation Society of America, Ankeny, IA. ·

Foster, G. R. and L. J. Lane. 1981. Modeling rill density - discussion. Journal of the Irr. and Drain. Div., ASCE. 107(IRl):109-112.

Foster, G. R. and L. J. Lane. 1983. Erosion by concentrated flow in Cann fields. pp. 9.65-9.82. IN: Pro­ceedings of the D. B. Simons Symposium on Erosion and Sedimentation. Colorado State University. Ft. Collins, CO.

-Foster, G. R. and L. J. Lane. 1987. User requirements. USDA-water erosion prediction project (WEPP). NSERL Report No. I, National Soil Erosion Research Laboratory, West Lafayette, IN.

Foster, G. R., L. J. Lane, M.A. Nearing, S. C. Finkner, J. J. Stone and L. T. West. 1989. USDA- Water ero­sion prediction project: Hill slope profile version. Chapter 10. Erosion Component. NSERL Report No. 2, National Soil Erosion Research Laboratory, USDA-ARS, W. Lafayette, IN.

Foster, G. R., L. J. Lane, J. D. Nowlin, L. M. Laflen and R. A. Young. 1980. A model to estimate sediment yield from field-sized areas: development of model. In CREAMS - A field scale model for Chemi­cals, Runoff, and Erosion from Agricultural Management Systems. Volume I: Model Documentation, Chapter 3. USDA-SEA, Conservation Report No. 26, pp. 36-64.

Foster, G. R. and L. D. Meyer. 1972a. A closed-fonn soil erosion equation for upland areas. IN: Sedimen­tation (Einstein). H. W. Shen (ed.). Colorado State University, Fort Collins, CO. Chapter 12.

Foster, G. R. and L. D. Meyer. !972b. Transport of soil particles by shallow flow. Transactions of the ASAE, 15(1):99-102.

Foster, G. R. and L. D. Meyer. 1975. Mathematical simulation of upland erosion by fundamental erosion mechanics. In: Present and Prospective Technology for Predicting Sediment Yields and Sources. ARS-S-40. USDA-Agricultural Research Services. 285 p.

Foster, G. T., L. D. Meyer and C. A. Onstad. 1977a. An erosion equation derived from basic erosion princi­ples. Transactions of the ASAE 20(4):678-682.

Page 125: Hydrology and Sedimentology of Dynamic Rill Networks ...

116

Foster, G. T., L. D. Meyer and C. A. Onstad. 1977b. A runoff erosivity factor and variable slope length exponents for soil loss estimates. Transactions of the ASAE 20(4):683-687.

Foster, G. T. and W. H. Wischmeier. 1974. Evaluating irregular slopes for soil loss prediction. TRANSACTIONS of the ASAE 17(2):305-309.

Foster, G. R., R. A. Young and W. H. Neibling. 1985. Sediment composition for oonpoint soun:e pollution analyses. TRANSACTIONS of the ASAE 28(1):133-139, 146.

Free, G. R. 1960. Erosion characteristics of rainfall. Agricultural Engineering. 41(7):447-449, 455.

Gerald, C.F .. 1980. Applied Numerical Analysis. Second Edition, Addison-Wesley Publishing Company, Reading, MA.

Gilley, J.E. and S. C. Finkner. 1985. Estimating soil detachment caused by raindrop impacL Transactions of the ASAE. 28(1):140-146.

Gilley, J.E., D. A. Woolhiser and D. B. McWhorter. 1985a. Interrill soil erosion - Part I: Development of model equations. Transactions of the ASAE. 28(1):147-153, 159.

Gilley, J.E., D. A. Woolhiser and D. B. McWhorter. 1985b. Interrill soil erosion - Part Il: Testing and use of model equations. Transactions of the ASAE. 28(1):154-159.

Govers, G. and G. Rauws. 1986. Transporting capacity of overland flow on plane and on irregular beds. Earth Surface Processes Landforms. 11:515-524.

Graf, W. H. 1971. Hydraulic of Sediment Transport. McGraw-Hill Book Co,~ New York,NY. 544 pp.

Grissinger, E. H. 1982. Bank erosion of cohesive materials. IN: Gravel-Bed Rivers, R. D. Hey, J. C. Bathurst, and C.R. Thome, eds. John Wiley and Sons. pp. 273-287.

Guy, B. T., W. T. Dickinson and R. P. Rudra. 1986. The roles of rainfall and runoff in the sediment trans-port capacity of interrill flow. ASAE Paper No. 86-2537. ASAE, SL Joseph, MI.

Haan, C. T. 1977. Statistical Methods In Hydrology. Iowa State University Press, Ames, IA.

Hillel, D. 1982. Introduction to Soil Physics. Academic Press, New York.

Hirschi, M. C. 1985. Modeling soil erosion with emphasis on steep slopes and the rilling process. Ph.D. dissertation, University of Kenrucky, Lexington, KY. 497 p.

Hirschi, M. C. and B. J. Barfield. 1988a. KYERMO - A physically based research erosion model. Part I. Model development Transactions of the ASAE. 31(3):804-813.

Hirschi, M. C. and B. J. Barfield. 1988b. A physically based research erosion model. Part II. Model sensi­tivity analysis and testing. Transactions of the ASAE. 31(3):814-820.

Hirschi, M. C., B. J. Barfield, I. D. Moore and D. G. Colliver. 1987. Profile meters for detailed measure­ment of soil surface heights. Applied Engineering in Agriculture, 3(1):47-51.

Hirschi, M. C., B. J. Barfield and I. D. Moore. 1983. Modeling erosion on long steep slopes with emphasis on the rilling process. Research Report No. 148, Water Resources Research Instiblte, University of Kentucky.

Holbrook, K. F., J. T. Ligon and J. C. Hayes. 1986. Comparison of methods for determination of eroded size distribution in sedimentology modeling. Southeast Region Meeting, American Society of Agri­cultural Engineers, St. Joseph, MI.

Page 126: Hydrology and Sedimentology of Dynamic Rill Networks ...

117

Horton, R. E. 1945. Erosional development of streams and their drainage basins, hydrological approach to quantitative morphology. Geol. Soc. Am. Bull. 56:275-370.

Rotes, F. L., K. H. Ateshian and B. Sheikh. 1973. Comparative costs of erosion and sediment conuol con­struction activities. Report No. EPA-430/9-73-016. U.S. Environmental Prolection Agency, Wash­ington, D. C.

Howard, A. D. 1971. Simulation of stream networks by headwall growth and branching. Geogr. Anal., 3:29-50.

Huang, C., J.M. Bradford and J. H. Cushman. 1982. A numerical study of raindroP ilnpact phenomena: the rigid case. Soil Sci. Soc. of Amer. Jour. 46(1):14-19.

Huang, C., J. M. Bradford and J. H. Cushman. 1983. A numerical study of raindroP ilnpact phenomena: the elastic deformation case. Soil Sci. Soc. Amer. Jour. 47:855-861.

Jenson, S. K. and J. 0. Domingue. 1988. Extracting topegraphic structure from digital elevation data for geographic information system analysis. Photogrammetric Engineering and Remote Sensing 54(11):1593-1600.

Julian, P. Y. 1987. Comments on "Sediment transpon capacity of sheet and rill flow: Application of unit stream power theory." By I. D. Moore and G. J. Burch. Water Resources Research, 23(8):1700.

Kalinske, A. A. 1947. Movement of sediment as bed-load in rivers. Trans. Am. Geophy. Union, 28(4):123-141.

.·Karlingcr, M. R. and B. M. Trouunan. 1989. A random spatial network model based on elementary poslll­lates. Water Resources Research, 25(5):793-798.

· Kelly, W. E. and R. C. Gularte. 1981. Erosion resistance of cohesive soils. Journal of the Hydraulics Divi­sion, Proceedings of the ASCE 107(HYI0):1211-1224.

,Kilinc, M. and E. V. Richardson. 1973. Mechanics of soil erosion from overland flow generated by silnu- · lated rainfall. Hydrology Paper No. 63, Colorado State University, FL Collins, CO.

Kinnell, P. I. A. 1974. Splash erosion: Some observations on the splash-cup technique. Soil Sci. Soc. Am. Proc. 38:657-660.

Kinnell, P. I. A. 1976. Splash erosion of primary particles and aggregates. Soil Sci. Soc. Am. J. 40:966-968.

Laflen, J. M., G. R. Foster and C. Onstad. 1985. Simulation of individual stonn losses for modeling the impact of soil-erosion on cropland productivity. In Soil Erosion and Conservation, eds. El-Swaify, Moldenhauer, and Lo. 285-295. Ankeny, IA:SCSA.

Lane, E. W. 1953. Progress report on studies on the design of stable channels of the Bureau of Reclama­tion. ASCE Proc., Irrigation and Drainage Division, Separate No. 280. 31 pp.

Lane, E. W. 1955. Design of stable channels. ASCE Transactions, 120: 1234-1279.

Lane, L. J. and G. R. Foster. 1980. Concentrated flow relationships. In CREAMS A field scale model for Chemicals, Runoff, and Erosion from Agricultural Management Systems. Volume ID. Supporting Documentation, Chapter 11, USDA-SEA, Conservation Report No. 26, pp. 474-485.

Laws, J. 0. and D. A. Parsons. 1943. The relationship of raindrop-size to intensity. Trans. Amer. Geo. Union, 24:452-460.

Laursen, E. M. 1958. Total total sediment load of streams. Jour. of Hyd Div~ ASCE, 84(HY1):1530-l, 1530-36.

Page 127: Hydrology and Sedimentology of Dynamic Rill Networks ...

118

Leighly, J. B. 1932. Toward a theory of morphologic significance of turbulence in the flow of waler in streams. University of California Publications in Geography, 6(1):1-22.

Leopold, L. B and W. B. Langbein. 1962. The concept of entropy in landscape evolution. U.S. Geol Surv. Prof. Pap. 500-A, 1962.

Leopold, L.B., M. G. Wolman and J.P. Miller. 1964. Fluvial Process in Geomorphology. W. H. Freeman and Co. 522 pp.

Li, R., V. M. Ponce and D. B. Simons. 1980. Modeling rill density. lour. Irr. and Drain. Div., ASCE, 106(1R 1):63-67.

Lu, J. Y., G. R. Foster and R. E. Smith. 1987. Numerical simulation of dynamic erosion in ridge- funow systems. Transactions of the ASAE, 30(4):969-976.

Lundgren, H. and I. G. Jonsson. 1964. Shear and velocity distribution in shallow channels. Journal of the Hydraulics Div., Proceedings of the ASCE 90(HYl):l-21.

Lyle, W. M. and E. T. Smerdon. 1965. Relation of compaction and other soil properties to erosion resis­tance of soils. TRANSACI10NS of the ASAE, 8(3):419-422.

McCool, D. K., G. R. Foster, C. K. Mutchler and L. D. Meyer. 1989. Revised slope length factor for the Universal Soil Loss Equation. Trans. of the ASAE. 32(5):1571-1570.

Mannering, J. V. and L. D. Meyer. 1963. The effects of various rares of surface mulch on infiltration and erosion. Soil Sci. Soc. Am. Proc. 27:84-86.

Matheron, G. 1973. The intrinsic random functions and their applications. Adv. Appl. Prob. 5:439-468.

Mantoglou, A. 1987. Digital simulation of multivariate two- and three-<limensional stochastic processes with a spectral turning bands methods. Math. Geol., 19(2):129-149.

Mantoglou, A. and J. L. Wilson. 1981. Simulation of random fields with the turning bands method. Tech. Rep. 264, Ralph M. Parsons Lab., Dept of Civ. Eng., Mass. Inst of Tech., Cambridge, Mass.

Mantoglou, A. and J. L. Wilson. 1982. The turning bands method for simulation of random fields using line generation by a spectral method. Water Resources Research, 18(5):1379-1394.

Mantz, P.A. 1977. Incipient transport in fine grains and flakes by fluids-extended Shields diagram. lour. of Hydr. Division, ASCE, 103 (HY6):601-615.

Meyer, L. D. 1981. How rain intensity affects inrerrill erosion. Transactions of the ASAE. 24(6):1472-1475.

Meyer, L. D. and E. J. Monke. 1965. Mechanics of soil erosion by rainfall and overland flow. Trans. of the ASAE. 8(4):572-577, 580.

Meyer, L. D. and J. M. Romkens. 1976. Erosion and sediment control on reshaped land Proceedings of the Third Federal Inter-Agency Sedimentation Conference.

Meyer, L. D. and W. H. Wischmeier. 1969. Mathematical simulation of the process of soil erosion by water. Transactions of the ASAE. 12(6):754-758, 762.

Meyer-Peter, E. and R. Muller. 1948. Formulas for bed load transport pp. 39-64. IN: Proceedings of the 2nd Congress of the !AHR, Stockholm.

Moore, I. D. 1981. Effect of surface sealing on infiltration. Transactions of the ASAE 24(6):1546- 1552, 1561.

Page 128: Hydrology and Sedimentology of Dynamic Rill Networks ...

119

Moore, I. D. and G. J. Burch. 1986. Sediment transport capacity of sheet and rill flow: application of unit stream power theory. Water Resources Research 22(8):1350-1360.

Moore, I. D. and G. J. Burch. 1987. Reply. Water Resources Research. 23(8):1701-1702.

Moore, I. D., M. C. Hircshi, and B. J. Barfield. 1983. Kentucky Rainfall Simula«I". Transactions of the ASAE, 26(4):1085-1089.

Morisawa, M. 1985. Rivers: Form and Processes. Longman, New York.

Mosley, M. P. 1974. Experimental study or rill erosion. Transactions of the ASAE 17(5):909-913, 916.

Mossaad, M. E. and T. H. Wu. 1984. A stochastic model of soil erosion. International Journal for Numeri· cal and Analytical Methods in Geomechanics, 8:201-224.

Mutchler, C. K. 1967. Parameters for describing raindrop splash. J. Soil and Water Coos. 22(3):91-94.

Mutchler, C. K. and C. L. Larson. 1971. Splash amount from waterdrop impact on a smooth smface. Warec Resources Research. 7(1):195-20.

Nearing, M. A., L. D. Ascough and H. M. L. Chaves. 1989. WEPP model sensitivity analysis. In USDA· Water Erosion Proteetion Project: Hillslope Profile Model Documentation, Chapter 14. NSERL Report No. 2, National Soil Erosion Research Laboratory, USDA-ARS, W. Lafayette, IN.

Nearing, M. A., G. R. Foster, L. J. Lane and S. C. Finkner. 1989. A process-based soil erosion model for USDA-water erosion prediction project technology. Transactions of the ASAE. 32(5):1587-1593.

Ogunlela, A., B. N. Wilson, C. T. Rice and G. Couger. 1989. Rill network development and analysis under simulated rainfall. ASAE Paper No. 89-2112, ASAE, SL Joseph, MI.

Ormsbee, L.E., G.E. Blandford, J.S. Montgomery, L.B. Terrell, BJ. Barfield, and D.E. Storm. 1990. Hydrology and sedimentology or dynamic rill networks. Volume TI: Hydrologic model for dynamic rill networks. Departments of Agricultural and Civil Engineering, University of KenlUCky, Lexing­ton, Kentucky. Final report prepared for the U.S. Geological Survey Matching Grant No. 14-08-001-Gl 147.

Palmer, R. S. 1965. Waterdrop impact forces. Transactions of the ASAE. 8(1):69-70. Palk, S. W., J. K. Mitchell and G. D. Bubcnzer. 1983. Rainfall characteristics and their relation to splash erosion. Transactions of the ASAE. 26(4):795-804.

Piest, R. F., J. M. Bradford and G. M. Wyatt 1975. Soil erosion and sediment transport from gullies. J. Hydr. Div. Proc., ASCE. !Ol(HYl):65-80.

Quansah, C. 1981. The effect of soil type, slope, rain intensity, and their interactions on splash detachment and transport. Journal of Soil Science. 32:215-224.

Raudkivi, A. J. 1976. Loose boundary hydraulics, 2nd ed. Peragamon Press. 397 p.

Rawls, W. J., J. J. Stone and D. L. Brakensiek. 1989. Infiltration. In USDA-Water Erosion Protection Project: Hillslope Profile Model Documentation, Chapter 4. NSERL Report No. 2, National Soil Ero­sion Research Laboratory, USDA-ARS. W. Lafayette, IN.

Renard, K.G., G.R. Foster, and G.A. Weesics. 1990. Predicting soil erosion by water· A guide to conserva­tion planning with the revised universal soil loss equation. U.S. Department of Agriculture, ARS Spe­cial Publication (In preparation).

Replogle, J. A. and V. T. Chow. 1966. Tractive-force distribution in open channels. Jour. Hydr. Div., ASCE, 92(HY2):169-191.

Page 129: Hydrology and Sedimentology of Dynamic Rill Networks ...

120

Rohlf, R. A. 1981. Development of a detenninistic mathematical model for interrill and rill nmoff and ero­sion. M.S. Thesis, University of Kentucky, Lexingron, KY. 84 p.

Romkens, M. J., C. B. Roth and 0. W. Nelson. 1975. Soil erosion on selected high clay subsoils. Journal of Soil and Water Conservation. 30:173-176.

Romkens, M. J., C. B. Roth and D. W. Nelson. 1977. Erodibility of selected clay subsoils in relation 10 physical and chemical properties. Soil Science of America Journal 41 :954-960.

Rose, C. W. 1960. Soil detachment caused by rainfall. Soil Science. 89(1):28-35.

Schumm, S. A., M. D. Harvey and C. C. Watson. 1984. Incised Channels: Morphology, Dynamics, and Con1rol. Water Resources Publication, Littleton, CO.

Schwab, G. 0., R. K. Frenuert, T. W. Edminster and K. K. Barnes. 1981. Soil and Waler Conservalion Engineering. 3rd Edition. New York: John Wiiey and Sons, 683 pp.

Shields, A. 1936. Application of the theory of similarity and turbulence research 10 the bed load movement, Mitt. Preuss. Vers. Wasser Schiff, 26:5-24.

Shultz, J.P., A. R. Jarrett and J. R. Hoover. 1985. Detachment and splash of a cohesive soil by rainfall. Transactions of the ASAE. 28(6):1878-1884.

Simons, D. B. and F. Senturk. 1977. Sediment Transport Technology. FL Collins, CO: Water Resources Publication.

Smerdon, E.T. and R. P. Beasley. 1961. Critical tractive forces in cohesive soils. Agricultural Engineering 42(1):26-29.

Smith, D. 0. and W. H. Wischmeier. 1957. Factors affecting sheet and rill erosion. Transactions American Geophysical Union. 38:889-896. --

Strahler, A. N. 1952. Hyposometric (area-altitude) analysis of erosional topography. 1952. Geol. Soc. Am. Bull. 63:1117-1142.

Stenn, D. E. 1986. Modeling phosphorous transport in surface runoff from agricultural watersheds for non­point source pollution assessment. M.S. thesis, Virginia Polytechnic Institute and State University, Blacksburg, Virginia, 202 p.

Stenn, D.E., BJ. Barfield. and L.E. Onnsbee. 1990. Hydrology and sedimentology of dynamic rill net­works. Volume IV: dynamic erosion model validation. Final Report prepared for the U.S. Geological Survey Matching Grant No. 14-08-001-GI 147. Department of Agricultural Engineering, University of Kentucky, Lexington, KY. ·

Task Committee for Preparation of Sedimentation Manual. 1971. Sediment transport mechanics: H. sedi­ment discharge fonnulas. Journal of the Hydraulics Division, Proceedings of the ASCE 97(HY4):523-567.

Tennekee, H. and J. L. Lumley. 1987. A first course in Turbulence. The MIT Press, Cambridge, MA.

Thomas, W. A. 1976. Scour and deposition in rivers and reservoirs. HEC-6. Hydrologic Engineering Cen­ter, U.S. Army Corps of Engineers.

Thompson, A. F. B., R. Ababou and L. W. Gelhar. 1989. Implementation of the three-dimensional turning bands random field generator. Water Resources Research, 25(10):2227-2243.

Walker, J. S. 1988. Fourier Analysis. Oxford University Press, New York.

Page 130: Hydrology and Sedimentology of Dynamic Rill Networks ...

121

Wang, R. C. T. and H. G. Wenzel. 1970. The mechanics of a drop after striking a stagnallt water layer. Uni­versity of !Uinois, Water Resources Center, Repon No. 30, Urbana, 130 pp.

Ward, A D .• B. N. Wilson. T. C. Bridges and B. J. Barfield. 1980. An evaluation of hydrologic modeling techniques for determining a design storm hydrograph. Proc. 7th lnt'I. Symp. of Urban Stonn Runoff. University of Kentucky, Lexington, Ky.

Weaver, H. J. 1989. Theory of discrete and continuous fourier analysis. John Wiley & Sons, New York.

Williams, J. R. 1976. Sediment yield prediction with Universal Equation using runoff enezgy factor. In: Present and Prospective Technology for Predicting Sediment Yields and Sources. ARS-S-40. USDA-Agricultural Research Service.

Wilson, B.N., BJ. Barfield, and I.D. Moore. 1983. A hydrology and sedimentolgy watershed model. Part I: Modeling techniques. Department of Agricultural Engineering, University of Kentucky, l.eJ<ington, KY.

Wilson, B. N., H. W. Downs, C. T. Haan and G. A. Kranzler. 1988. Transport of eroded soil particles. Uni­versity Center for Water Research. Oklahoma State University, Stillwater, OK. 38 pp.

Wischmeier, W. H. 1959. A rainfall erosion index for a Universal Soil Loss Equation. Soil Sci. Soc. Am. Proc. 23:246-249.

Wischmeier, W. H .• C. B. Johnson, and B. V. Cross. 1971. A soil erodibility nomograph f<r farmland and construction sites. Jour. of Soil and Water Cons., 26(5):189-193.

Wischmeier, W. H. and D. D. Smith. 1958. Rainfall energy and its relationship to soil loss. Trails. Amer. Geo. Union, 39(2):285-291.

Wischmeier. W. H. and D. D. Smith. 1%5. Rainfall erosion losses from cropland east of the Rocky Moun­tains. Agriculmre Handbook No. 282, U.S. Department of Agriculture, Washington, D.C.

Wischmeier, W. H. and D. D. Smith. 1978. Predicting rainfall erosion losses - a guide to conservation planning. USDA, AH-537.

Wu, T. H., K. W. Bedford, M. E. Mossaad, K. C. McLurdy and E. M. Ali. 1982. A stoehastic model of soil erosion. Water Resources Center, Ohio State University. Completion Report No. 711523, NTIS No. PB82-224734. 145 p. .

Wu, C. C. and L. D. Meyer. 1988. Numerical model for nonuniform sediment transport along flatland fur­rows. Am. Soc. of Agr. Eng., No. 88-2531. SL Joseph, MI.

Yalin, M. S. 1963. An expression for bed-load transportation. Jour. of the Hydr. Div., Proc. of the ASCE, 89(HY3):221-250.

Yalin, M. S. 1977. Mechanics of Sediment Transport, 2nd ed. Peragamon Press. 298 pp.

Yang, C. T. 1972. Unit stream power and sediment transport Jour. of Hydr. Div., Proc. of the ASCE, 98(HY10):1805-1826.

Yang, C. T. 1973. Incipient motion and sediment transport Jour. of Hydr. Div., Proc. of the ASCE, 99(HY10): 1679-1704.

Young, R. A. and J. L. Wiersma. 1973. The role of rainfall impact in soil detachment and transport Water Resources Research. 9(6):1629-1636.

Zimmerman, D. A. and J. L. Wilson. 1989. Description of and user's manual for TUBA: A computer code for generating two-dimensional random fields via the turning bands method. Draft #3, Gram, Inc., Albuquerque, NM.

Page 131: Hydrology and Sedimentology of Dynamic Rill Networks ...