You are on page 1of 60

An Integrated Hillslope and Channel Evolution Model as an Investigation and Prediction Tool

Year 2 annual report DACA88-95-R-0020

Prepared by Greg Tucker Nicole Gasparini Stephen Lancaster Rafael Bras, PI

Department of Civil and Environmental Engineering Massachusetts Institute of Technology Cambridge, MA

August, 1997

An Integrated Hillslope and Channel Evolution Model as an Investigation and Prediction Tool
Year 2 Annual Report
Introduction ............................................................................................................................... 1-1 Overview of Model Components ............................................................................................. 2-1 Grid Elements ................................................................................................................. 2-3 Storm Size, Duration, and Frequency ............................................................................. 2-5 Flow Routing and Runoff ............................................................................................... 2-6 Bedrock Weathering (Regolith Production) ................................................................... 2-7 Hillslope Sediment Transport: Continuous Processes .................................................... 2-7 Hillslope Sediment Transport: Mass Movement ............................................................ 2-8 Stream Erosion and Deposition ...................................................................................... 2-8 Grain-Size Sorting ........................................................................................................ 2-10 Lateral Stream Channel Erosion (Meandering) ............................................................ 2-10 Vegetation ..................................................................................................................... 2-10 Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh ................ 3-1 Grid Elements and Data Structures ................................................................................. 3-3 Drainage Networks on a Triangulated Irregular Mesh ................................................... 3-7 Numerical Algorithms .................................................................................................... 3-8 Variable Resolution for Modeling River Meandering .................................................. 3-10 A Stochastic Approach to Modeling Drainage Basin Evolution ........................................... 4-1 Model Description .......................................................................................................... 4-2 Sensitivity of Erosion Rate to Rainfall Variability ......................................................... 4-5 Morphologic Consequences of Rainfall Variability: Numerical Example ..................... 4-9 Dynamics of Vegetation and Runoff Erosion ......................................................................... 5-1 Model of Vegetation and Erosion ................................................................................... 5-2 Characteristic Form Profiles ........................................................................................... 5-6 Numerical Examples ..................................................................................................... 5-11 Understanding the Interactions of Multiple Grain Sizes ...................................................... 6-1 Meandering: A Simple Model ................................................................................................. A-1

CHAPTER 1

Introduction

Understanding the dynamics of landscape evolution is a challenging problem, for two reasons. First, the processes involved are inherently destructive, and therefore the geologic record of landscape development is often fragmentary. Second, the sculpture of terrain involves a fascinating but complex set of interacting nonlinear processes. While challenging, however, the problem is not intractable. Information on landscape history is still preserved in the form of topography itself, and often also in the form of associated sedimentary deposits such as alluvial valley lls. And despite the complexity of geomorphic processes and their interactions, the resultant landforms often exhibit an underlying similarity associated with different geologic and climatic settings. One of the important challenges in modern geomorphic research is to investigate landscape dynamics by developing simulation models that incorporate the process interactions we observe in nature. This report describes progress to date on the development of one such model. The Channel-Hillslope Integrated Landscape Development model (CHILD) is a descendant of earlier modeling efforts at MIT and at the Pennsylvania State University, and it incorporates many of the important processes not

Channel Hillslope Integrated Landscape Development Model

1-1

Introduction

included in previous models. Its primary near-term purpose is to serve as a quantitative tool for investigating the impact of Quaternary-scale climate change on hillslope, channel, and oodplain evolution in non-glaciated drainage basins. This report focuses on the theory and implementation of the model itself, and on its application to several theoretical problems that are closely related to the problem of geomorphic responses to environmental change. The CHILD model is designed to simulate the evolution of uvially-dominated landscapes formed chiey by physical erosion (thus, it does not include glacial erosion or karst development, for example). It simulates the interaction of two general types of process: uvial processes, a category which encompasses erosion or deposition by runoff cascading across the landscape (including slope wash and channel and rill erosion), and hillslope processes, which includes weathering, soil creep, and other slope transport processes. The computer code is written to maximize exibility and modularity, so that the user is free to formulate the problem in a way that is suitable to the problem at hand, rather than being dictated by the software itself. To this end, the model provides a number of different options for activating or deactivating different processes and/or different formulations of the same process. The various components in the model are diagrammed in Figure 1-1. Those shown in gray are planned components that have not yet been fully implemented as of this writing. In some cases, prototypes of process modules have been developed within the framework of the xed-grid GOLEM model but not yet adapted to the variablemesh data structure used in CHILD. Some of the important improvements over previous models include:
Adaptive, irregular mesh: the model uses an irregular nite-difference gridding

method based on the model of Braun and Sambridge (1997) to represent the landscape surface. Use of an irregular simulation mesh allows for the incorporation of lateral stream erosion and makes it possible to represent different parts of the landscape at different spatial resolutions. It also eliminates some of the gridding artifacts associated with xed-grid methods, and opens up the possibility of coupling the model with three-dimensional tectonic deformation models (e.g., strikeslip or thrust faulting). Implementation of the irregular mesh is described in Chapter 3.

1-2

Channel Hillslope Integrated Landscape Development Model

CLIMATE MODULE

RUNOFF AND FLOW ROUTING

STORM GENERATOR

HILLSLOPE TRANSPORT (soil creep, landsliding, etc.)

FLUVIAL EROSION & DEPOSITION

LANDSCAPE STATE VARIABLES: Elevation z(x,y,t) Drainage area A(x,y,t) Surface runoff Q(x,y,t) Regolith thickness C(x,y,t) Percent gravel D%(x,y,t) Vegetation cover V(x,y,t)

WEATHERING

SELECTIVE GRAIN SIZE TRANSPORT

VEGETATION LATERAL EROSION (MEANDERING)

MESH UPDATER (move, add, delete points)

Schematic illustration of the state variables and process modules in the CHILD model. Boxes shown in gray are still under development.
FIGURE 1-1.

Channel Hillslope Integrated Landscape Development Model

1-3

Introduction

Stochastic rainfall forcing: Previous models have always modeled sediment trans-

port by using a single effective rainfall or runoff rate that represents a geomorphic average. CHILD relaxes that assumption by providing the option of stochastic rainfall input. The stochastic rainfall model is described further in Chapter 4. Stream meandering: this is the rst landscape evolution model in which the processes of vertical stream erosion and lateral channel migration (meandering) are coupled. Meandering is simulated using the model developed by Lancaster and others (in prep) (a copy of which appears in Appendix A). The interface between CHILD and the meander model is currently under development; it is discussed briey in Chapter 3. Multiple sediment sizes: A model of mixed-size sediment transport has been developed within the framework of the GOLEM model, and is currently being adapted for incorporation into CHILD. The multi-size model is discussed in Chapter 6. Vegetation: CHILD provides a module to simulate the interaction of vegetation growth and runoff erosion. This module simulates the effect of vegetation on surface resistance to overland ow, vegetation removal due to runoff erosion, and vegetation regrowth during interstorm periods. The vegetation model is discussion further in Chapter 5. The CHILD code is written in C++ in order to exploit the advantages of an objectoriented programming language. This report is divided into two parts. The rst part describes the present version of the model and its implementation. Chapter 2 presents an overview of the basic theory and a summary of the different model components. Chapter 3 gives a description of the algorithms and data structures used to implement the irregular mesh. The second part of the report consists of a three short chapters that address the role of different processes incorporated in the model. Essentially, each of these chapters focuses on the role of one process, and includes both a description of the underlying theory and a sensitivity analysis. These chapters are built around the following set of guiding questions:

1-4

Channel Hillslope Integrated Landscape Development Model

What is the role of rainfall variability in controlling drainage basin morphology

and evolution? (Chapter 4) What is the nature of the interaction between vegetation growth and runoff erosion in catchment evolution? (Chapter 5) How do interactions between hillslope and channel processes and between different branches of a drainage network inuence spatial variations in grain size? (Chapter 6) We do not expect the model to provide denitive answers to these questions. What the model can do, however, is provide insights that can be further tested through eld studies. The sensitivity analyses described in these chapters are also important prerequisites to calibrating and applying the model to a specic eld site, because they provide a basis for understanding the models behavior under idealized cases.

Channel Hillslope Integrated Landscape Development Model

1-5

Introduction

1-6

Channel Hillslope Integrated Landscape Development Model

CHAPTER 2

Overview of Model Components

The CHILD model simulates landscape evolution by tracking the passage of water and sediment across an irregular lattice of points that represents the landscape surface. Figure 2-1 shows a typical simulation, highlighting the drainage networks that form naturally when converging ow excavates valleys and leads to further ow convergence. The model tracks four basic state variables that determine the depth of erosion or deposition at each point during a given iteration: elevation, slope, drainage area, and surface runoff (Figure 1-1). In addition, at the users option the model can track several additional state variables, including regolith thickness (depth to bedrock), fractional vegetation cover, percent gravel in the active surface sediment layer, and the number, thickness, and composition of sediment layers previously deposited at a given point.1 In generic mathematical terms, the model can be expressed by the following equation:

1. At this writing, the module for selective transport of different sediment sizes is currently under development, as noted below.

Channel Hillslope Integrated Landscape Development Model

2-1

Overview of Model Components

(a)

(b)

Drainage basin simulated using the CHILD model. (a) Perspective view of landscape. (For graphical display purposes, the mesh has been interpolated onto a regular grid. Visualization is from the SG3D module of GRASS.) (b) Plan view of drainage network and irregular mesh. The catchment outline is that of the Forsyth Creek watershed, Fort Riley, Kansas, and represents an area of about 11.5 km2.
FIGURE 2-1.

z = U ( [ x, y, t ] ) F ( Q, S, [ Q s, C, V , D ] ) H ( z ) t

(EQ 1)

where z is surface elevation, t is time, and terms in brackets [] denote optional dependencies. The rst term, U(), represents tectonic uplift (or equivalently, baselevel lowering), which may vary in time and space. The second term, F(), represents runoff erosion, and is a function of the runoff rate, Q, surface slope in the direction of ow, S, and optionally also of sediment ux, Qs, regolith thickness, C, vegetation cover, V, and percent gravel in the substrate, D. No distinction is made between erosion by overland ow and by fully channelized ow. The surface runoff rate, Q, is a function

Channel Hillslope Integrated Landscape Development Model

2-2

Grid Elements

of drainage area, A, and optionally of slope, S (which can inuence the likelihood of surface saturation and therefore of runoff production). Surface runoff can be modeled using a series of random storm events, or it can be modeled as a constant runoff rate that represents an average effective geomorphic event. The third term, H(), represents sediment transport by hillslope processes such as soil creep, raindrop impact, and landsliding. In the present version of the model, hillslope transport depends solely on local topography; however, future versions may incorporate landsliding and its dependence on soil pore pressure. To compute the effects of these different processes, the model iterates through a series of discrete storm events and interstorm periods. These storm events can be modeled stochastically, with the intensity and duration of each chosen at random from a distribution, or they may be uniform, in which case the storm properties represent a geomorphically average event (the assumption made in most previous models [e.g., Ahnert, 1976; Kirkby, 1987; Willgoose et al., 1991; Howard, 1994; Tucker and Slingerland, 1994; Moglen and Bras, 1995]). The model provides exibility in the way that each of the process terms are modeled. Implementation of the various processes within the model is diagrammed in the form of a ow chart in Figure 2-2. The component process models are summarized below. Further discussion of many of these components is provided in later chapters.

Grid Elements
The irregular grid used in the model is described in detail in Chapter 3, but a brief overview now will help to clarify some of the concepts that follow. The model grid consists of a set of points N that are connected to form a mesh of triangles (Figure 23). Elevation, drainage area, and other state variables are computed at the points rather than within the triangles (in other words, the model uses a nite-difference rather than a nite-element approach). Points (or nodes) are connected using the Delaunay triangulation, which is the unique set of triangles that connect a given set of points in such a way that a circle passing through the three points in any triangle will contain no other points. Each node Ni is associated with a surrounding Voronoi Region (or Voronoi Cell). The Voronoi Region for a node Ni is the region within
Channel Hillslope Integrated Landscape Development Model

2-3

Overview of Model Components

INITIALIZE UPDATE MESH1 GENERATE NEW STORM2 COMPUTE EDGE SLOPES DRAINAGE DIRECTIONS DRAINAGE AREA3 SURFACE FLOW3 FIND OUTLETS FOR CLOSED DEPRESSIONS4 ADD NODES TO MEANDERING CHANNELS5
no

STREAM EROSION/ DEPOSITION DURING STORM (INCL. EROSION OF VEG. COVER)

CHANNEL MEANDERING6

INTERSTORM HILLSLOPE TRANSPORT

VEGETATION REGROWTH

BASELEVEL CHANGE

OUTPUT7

Notes: 1. Only on rst iteration or if mesh has changed during the previous iteration. 2. If option for stochastic storms is selected; otherwise, mean storm properties are used. 3. If at least one ow direction has changed, or if the mesh has been updated. 4. If the option for lakelling is selected; otherwise, ow entering a closed depression is assumed to evaporate. 5. If the meandering option is selected and higher resolution is needed. 6. If the meandering option is selected. 7. Only at selected output intervals.
FIGURE 2-2. Flow chart showing the sequence of computations in the model.

DONE?
yes

END

Channel Hillslope Integrated Landscape Development Model

2-4

Storm Size, Duration, and Frequency

Points (nodes) Voronoi Cell Edge


FIGURE 2-3.

Schematic illustration of model grid components.

which any arbitrary point Q would be closer to Ni than to any other node on the grid. The boundaries between Voronoi Cells are lines of equal distance between adjacent grid points. Each Voronoi Cell has surface area Av.

Storm Size, Duration, and Frequency


Each model iteration represents a storm event. Each storm event is associated with a rainfall (or runoff) intensity, P, a duration, Td, and an interstorm period before the next event, Ti. These parameters may be chosen at random for each storm using the model of Eagleson (1978), or they may be held constant throughout a simulation. In either case, storms are approximated as having constant rainfall intensity throughout their duration, and the same assumption is applied to the resulting hydrographs (see below). For variable storms, the storm properties are chosen at random from the following distributions:
Rainfall (runoff) intensity 1 P - exp -f ( P ) = - P P T d 1 - exp ----f ( T d ) = ----- Td T d
(EQ 2)

Storm duration

(EQ 3)

Channel Hillslope Integrated Landscape Development Model

2-5

Overview of Model Components

Interstorm period

T i 1 - exp ---f ( T i ) = ---- Ti T i

(EQ 4)

where P, T s, and T i are mean storm intensity, duration, and interstorm period, respectively. If stochastic storm generation is not used, the mean values are applied during each iteration (in which case the duration of each iteration is T s + T i ).

Flow Routing and Runoff


Surface ow collected at a point on the grid is routed downslope toward one of its adjacent neighbor nodes, following the edge that has the steepest downhill slope. If a pit occurs on the grid, with no downhill route away from a given node, water is either assumed to evaporate at that point, or a lake-lling algorithm is invoked to nd an outlet for the closed depression (see Chapter 3). The local contribution from rainfall at a node can be computed in one of two ways. In the rst method, the rainfall associated with a node Ni is equal to the effective runoff rate times the nodes Voronoi area, Av. Alternatively, ow can be collected on each triangular element and routed downslope toward the lowest of the triangles three vertices. The drainage area, A, for a node is the sum of the area of all Voronoi cells (or triangles, if the second method is used) that contribute ow to that node. Surface discharge is computed from drainage area in one of the following ways: (1) Hortonian (inltration-excess) runoff: Runoff production (rainfall minus inltration) is assumed to be uniform across the landscape. Assuming steady-state ow, the surface discharge at any point is equal to
Q = (P Ic) A ,
(EQ 5)

where Ic is inltration capacity (Q = 0 if P < Ic).

Channel Hillslope Integrated Landscape Development Model

2-6

Bedrock Weathering (Regolith Production)

(2) Saturation-excess runoff: An option is currently under development for computing direct runoff from saturated areas, with saturation dependent on position in the landscape, groundwater ux rate, and elapsed time since the last storm.

Bedrock Weathering (Regolith Production)


The model allows for two basic types of material, bedrock and regolith. The term regolith is used to represent any disaggregated sediment material, and includes soil, colluvium, and stream sediment. Conversion of bedrock to regolith by weathering processes is assumed to take place at the bedrock-regolith contact. The rate of conversion, expressed in units of depth per unit time, depends on the thickness of the regolith mantle, C, according to the exponential function
C t C = k w exp ----- , C 0
(EQ 6)

weathering

where t is time, kw is the rate of bedrock-to-regolith conversion when C=0, and C0 is a parameter that controls the rate of decrease in weathering rate with increasing regolith thickness. Note that for purposes of computational efciency, equation (6) is only computed when the model is run in bedrock-alluvial mode (see Stream Erosion and Deposition on page 8).

Hillslope Sediment Transport: Continuous Processes


Sediment transport by continuous hillslope processes such as soil creep and raindrop impact is modeled using the well-known geomorphic diffusion equation (e.g., Culling, 1960),
z t z z = kd 2 + 2 . x y
2 2

(EQ 7)

creep

Numerical solution of equation (7) on an irregular mesh is discussed in Chapter 3.

Channel Hillslope Integrated Landscape Development Model

2-7

Overview of Model Components

Hillslope Sediment Transport: Mass Movement


A module for landsliding presently exists within the GOLEM model, and will be incorporated into CHILD in a future update.

Stream Erosion and Deposition


The model distinguishes between detachment of material from a stream bed and transport of the detached material. The maximum detachment rate depends on local slope and discharge according to
D b = k b ( k t Q b S b cb ) ,
m n pb

(EQ 8)

where Db is the detachment (erosion) rate, cb is a threshold, and kb, kt, mb, nb, and pb are parameters. Note that with suitably chosen parameters, equation (8) can reprep sent excess shear stress, with D b = k b ( cb ) b ( = bed shear stress, cb = critical shear stress for erosion). The transport capacity for detached sediment material (referred to hereafter as regolith, a term which includes both sediment weathered from bedrock and eroded from the channel bed) is
C s = Wk f ( k t Q f S f c ) ,
m n pf

(EQ 9)

where Cs is transport capacity, W is channel width, and kf, kt, mf, nf, and pf are parameters. As with equation (8), equation (9) can be expressed in terms of excess bed shear stress using suitably-chosen values for kf, mf, and nf. In the present version of the model, channel width is computed using the empirical relationship m W = k cw Q cw (the constant kcw is assumed to be absorbed into the transport coefcient kf).

Channel Hillslope Integrated Landscape Development Model

2-8

Stream Erosion and Deposition

Two end-member cases and one intermediate case arise from equations (8) and (9), and special subroutines are provided in the model to handle each of these cases separately: Detachment-limited: If the sediment transport capacity is everywhere greater than the sediment ux, the rate of stream erosion is simply equal to the maximum detachment rate,
z b = Dc , t
(EQ 10)

where zb represents elevation of the channel bed above a datum within the underlying rock column. This formulation has been used in a number of studies to represent bedrock channel erosion (or more generally, detachment-limited erosion) (e.g., Seidl and Dietrich, 1992; Anderson, 1994; Howard et al., 1994; Seidl et al., 1994; Tucker and Slingerland, 1994; Moglen and Bras, 1995; Humphrey and Heller, 1995). It has the practical advantage of being simple and efcient to integrate numerically. Transport-limited: If sufcient sediment is always available for transport and/or the bed material is easily detached, streams can be assumed to be everywhere at their carrying capacity. Under this condition, continuity of mass gives the local rate of erosion or deposition as
z b 1 C s = ---, s x t
(EQ 11)

where s is sediment bulk density and x is a vector oriented in the direction of ow. These two different cases can be invoked in the model by selecting the detachmentlimited or transport-limited stream erosion option, respectively. Alternatively, the model may be run in bedrock-alluvial mode. This represents the most general case. In bedrock-alluvial mode, the rate of erosion is computed as the lesser of (1) the excess sediment carrying capacity or (2) the maximum detachment rate. Regolith is assumed to have an effectively innite detachment capacity; detachment capacity only becomes a limiting factor when bedrock is exposed in a channel.

Channel Hillslope Integrated Landscape Development Model

2-9

Overview of Model Components

Grain-Size Sorting
A model of transport and sorting of multiple sediment sizes has been developed within the framework of the GOLEM xed-grid landscape evolution model by N. Gasparini. The sorting model uses a two-phase sand-gravel transport formula developed by P. Wilcock at Johns Hopkins University. The multi-size transport model includes a set of routines for recording the depth and composition of previously deposited sediment layers. The multi-size model is described further in Chapter 6. Adaptation of the multi-size model to the CHILD framework is currently in progress.

Lateral Stream Channel Erosion (Meandering)


An interface is currently under development that will allow the CHILD model to be coupled with a 2D model of river meandering in order to compute later river erosion and oodplain widening. The meander model, written by S. Lancaster, is based on topographic steering of channel ow. The model parameterizes bank shear stress as a function of the cross-channel transfer of uid momentum due to topographic steering by point bars. The meander model is described in greater detail by Lancaster et al. (Appendix A).

Vegetation
Vegetation is modeled in terms of a percent vegetation cover on the surface, V. Vegetation increases the threshold shear stress that must be exceeded before runoff erosion can occur, according to
c ( V ) = cs + V cv ,
(EQ 12)

where cs is the critical shear stress for an unvegetated surface (primarily a function of grain size) and cv is the added critical shear stress under 100% vegetation cover.

Channel Hillslope Integrated Landscape Development Model

2-10

Vegetation

During storms, vegetation cover is eroded at a rate proportional to excess shear stress,
dV (erosion) = k vd V ( c ) , dt
(EQ 13)

where kvd is a constant that controls the rate of vegetation removal. The rate of vegetation regrowth during interstorm periods is modeled as a linear function of the amount of vegetation cover present,
dV (growth) = k vg ( 1 V ) . dt
(EQ 14)

The vegetation module is discussed further in Chapter 5.

Channel Hillslope Integrated Landscape Development Model

2-11

Overview of Model Components

Channel Hillslope Integrated Landscape Development Model

2-12

CHAPTER 3

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

Continuing advances in computing technology have made three-dimensional modeling an attractive tool for investigating landscape evolution. Most landscape evolution models represent three-dimensional terrain using a regular matrix of points, the same representation that is used in digital elevation models (e.g., Ahnert, 1976; Kirkby, 1987; Beaumont et al., 1992; Willgoose et al., 1991; Chase, 1992; Slingerland et al., 1993; Howard, 1994; Rigon et al., 1994; Tucker and Slingerland, 1994; Sinclair and Ball, 1996). Although signicant insights have been gained from models using this type of regular spatial discretization, the technique suffers from several drawbacks: (1) landform elements must be represented at a uniform spatial resolution, which in practice means the highest resolution required by any feature or process of interest; (2) drainage directions are restricted to 45 degree increments (though for watershed-scale applications, this limitation may be reduced by using multiple-ow algorithms (e.g., Costa-Cabral and Burgess, 1994; Tarboton, 1997)); (3) under certain circumstances, use of a regular grid introduces anisotropy that can lead to bias in drainage network patterns; and (4) use of a xed grid makes it difcult or impossible to model processes with a signicant horizontal component, such as thrust propagation or river meandering.

Channel Hillslope Integrated Landscape Development Model

3-1

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

The last of these constraints is especially signicant. Although we conventially speak of uplift, most crustal deformation processes involve a signicant amount of horizontal translation. Previous coupled models have either incorporated only the vertical component of deformation (e.g., Tucker and Slingerland, 1996; Kooi and Beaumont, 1996) or have represented lateral translation by simply offsetting two xed grids (e.g., Anderson, 1994). Coupled models of deformation, erosion, and sedimentation promise to yield important insights into such issues as the relationship between deformation and the stratigraphic record, but such models ultimately require the ability to model deformation in three dimensions. Similarly, erosional processes often have a signicant horizontal component that is neglected in current models. One of the most important horizontal erosion processes is lateral stream erosion, which by widening a valley can signicantly alter the depositional geometry within a oodplain over geologic time. One alternative to regular grids is the use of triangulated irregular networks (TINs) for representing topographic surfaces. Triangulated irregular networks, which are often based on the so-called Delaunay triangulation, are commonly used for constructing nite element meshes and for representing surfaces within geographic information systems. Techniques for constructing Delaunay triangulations and their corresponding Voronoi (or Thiessen) diagrams are well established in the eld of computational geometry (e.g., Guibas and Stol, 1985; Sloan, 1987; Knuth, 1992; Sugihara and Iri, 1994; Sambridge et al., 1995; Du, 1996). Recently, Braun and Sambridge (1997) adapted an existing large-scale landscape model to operate with a triangular irregular spatial discretization that overcomes many of the limitations associated with regular grids. They demonstrated that use of a triangulated irregular network has the advantages of eliminating anisotropy in drainage directions and allowing for a variable-resolution representation of topography. In this chapter, we describe the irregular mesh technique used in the CHILD model of hillslope and channel evolution. The mechanics of the model are described elsewhere (see Chapters 2 and 4-7); here we focus on implementation of the adaptive mesh. We describe an efcient data structure for implementing the irregular mesh, and briey discuss how the technique makes it possible to model three-dimensional valley formation by stream erosion.

Channel Hillslope Integrated Landscape Development Model

3-2

Grid Elements and Data Structures

Grid Elements and Data Structures


The irregular mesh used by the CHILD model is illustrated in Figure 3-1. Topography is represented in the model by a set of nodes N that are connected to form a mesh of triangles using the Delaunay triangulation of N. The Delaunay triangulation is a unique set of triangles that connect a set of points in such a way that a circle passing through the three points of any triangle will contain no other points. Elevation, drainage area, and other state variables in the model (see Figure 1-1) are computed at the nodes rather than within the triangles; in other words, the model uses a nite-difference rather than a nite-element approach. Each node Ni is associated with a Voronoi Region (or Voronoi cell, also known as a Thiessen polygon). The Voronoi Region for a node Ni is the region within which any arbitrary point Q would be closer to Ni than to any other node on the mesh. The boundaries between Voronoi Cells are lines of equal distance between adjacent grid points. A number of algorithms and data structures for representing topography and hydrology on a xed grid have been developed, and are generally fairly straightforward

400 200 0 0 1000 2000 3000


FIGURE 3-1. Isometric view of a simulated 4000 catchment, showing irregular mesh.

5000
Channel Hillslope Integrated Landscape Development Model

3-3

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

(e.g., Jenson and Domingue, 1988; Tarboton et al., 1991). Designing efcient data structures for an irregular triangulated mesh is more complicated. Unlike regular matrices, where each node is connected to either four or eight adjacent neighbors, the number of neighbors connected to a given node in a triangulated irregular network may be arbitrarily large. Ideally, a data structure should represent this variable connectivity in a way that (1) provides rapid access to adjacent mesh elements without demanding excessive storage space, and (2) is exible enough to handle dynamic changes in the mesh itself. The CHILD model uses a DualEdge data structure that provides an efcient way to satisfy these requirements. The DualEdge structure is adapted from the QuadEdge data structure of Guibas and Stol (1985), and consists of three geometric elements: nodes, triangles, and directed edges (Figure 3-2). The data structure is summarized in Figure 3-3.

(a) C b a D B

(b) C b a D
B.edg = BA BA.nextedg = BD BA.vvertex = b

Voronoi Cell

A.edg = AB AB.nextedg = AC AB.vvertex = a

FIGURE 3-2. Illustration of the dual-edge data structure, showing triangular lattice (black) and corresponding Voronoi diagram (gray). (a) Directed edge AB, counterclockwise nextedg AC, and right-hand Voronoi vertex a. (b) Complementary directed edge BA, counterclockwise nextedg BD, and right-hand Voronoi vertex b.

Channel Hillslope Integrated Landscape Development Model

3-4

Grid Elements and Data Structures

Directed Edges

Each edge of a triangle is associated with two directed edges, which share the same endpoints but point in opposite directions (Figure 3-2). Directed edges are dened by their origin and destination nodes; two directed edges that share the same endpoints are termed complementary edges. In addition to the origin and destination nodes, each directed edge data object includes a reference to the directed edge that lies immediately counter-clockwise relative to its origin node (Figure 3-2). This representation makes it possible to rapidly access all of the edges and neighboring nodes connected to a given node. Each directed edge object also includes the coordinates of the Voronoi vertex immediately clockwise. A Voronoi vertex is dened as the intersection point of three Voronoi cells (Figure 3-2). In general, each triangle is associated with a Voronoi vertex; each directed edge object includes a reference to the Voronoi vertex associated with the triangle on its right-hand side (clockwise). Pseudo-code for the DirectedEdge data structure is shown in Figure 3-3.
Nodes

Each node object includes x, y, z coordinates, the number of adjacent nodes, and a pointer to one of its directed edges (along with other information relevant to hydrologic routing and physical parameters, not shown) (Figure 3-3). The format of the Node and DirectedEdge objects makes it possible to efciently nd the Voronoi polygon associated with a node, using the following algorithm:
FindVoronoiPolygon( Node thenode ) BEGIN XYPoint voronoi_polygon(1..thenode.nnbrs) current_edge = thenode.edg FOR i=1,thenode.nnbrs DO voronoi_polygon(i).x := current_edge.vvertex_x voronoi_polygon(i).y := current_edge.vvertex_y current_edge := current_edge.nextedg END END

Channel Hillslope Integrated Landscape Development Model

3-5

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

Class Node x y z edg nnbrs Class DirectedEdge org dest nextedg vvertex_x vvertex_y Class Triangle p(3) t(3) e(3)

// x-coordinate // y-coordinate // z-coordinate (elevation) // rst connected edge // number of neighboring nodes

// origin node // destination node // next directed edge counterclockwise // x-coordinate of right-hand Voronoi vertex // y-coordinate of right-hand Voronoi vertex

// Vertex nodes // Adjacent triangles (t(1) is opposite p(1), etc.) // Clockwise-oriented directed edges

FIGURE 3-3. Pseudo-code summary of the DualEdge data structure, showing the data members belonging to Node, DirectedEdge, and Triangle objects.

Similarly, a list of neighboring nodes can be obtained by accessing current_edge.dest for each edge originating at a given node.
Triangles

Triangle objects include pointers to the three nodes in the triangle, the three neighboring triangles, and the three directed edges that are oriented clockwise with respect to the triangle. The nodes and neighboring triangles are numbered in such a way that the nth neighboring triangle lies opposite the nth vertex (see Figure 3-4). Each triangle is associated with a Voronoi vertex. This vertex is the intersection of the three Voronoi cells associated with each of the triangles vertices (e.g., Figure 32). Because the three edges are clockwise-oriented, their right-hand Voronoi vertex

Channel Hillslope Integrated Landscape Development Model

3-6

Drainage Networks on a Triangulated Irregular Mesh

P0 T2 P1 e1 e2 T0 e0 T1 P2

FIGURE 3-4. Illustration of numbering of triangle nodes, adjacent triangles, and clockwise edges in Triangle data objects.

is the Voronoi vertex associated with the triangle itself. In general, Voronoi vertices can be found by locating the intersection of the perpendicular bisectors of the triangles edges (note that there exist special cases in which this is not true; these cases must be detected and handled differently).

Drainage Networks on a Triangulated Irregular Mesh


The surface discharge at each node is computed as a function of the upstream contributing area (see Chapters 2 and 4). The model provides two alternative methods of nding the upstream contributing area at a node. In both methods, each node is assigned a drainage direction along the steepest path (edge) toward one of its neighboring nodes. In the rst method, the area of each triangle is assigned to the lowest of the triangles three nodes and to all nodes downstream of that point. In the second method, the contributing area at a node i is equal to the sum of the Voronoi areas of all nodes that ow to i (including i itself). In some cases a node may form a local depression, with no neighbors lower than itself. This case can be handled in one of two ways in the model. The simplest method assumes that all water entering a sink evaporates at that point. Alternatively, an outlet can be found for each sink using the Lake Fill algorithm. The Lake Fill algorithm starts by creating a stack of contiguous ooded nodes, which initially

Channel Hillslope Integrated Landscape Development Model

3-7

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

contains just the sink itself. The perimeter of the ooded region (lake) is then iteratively searched to identify the lowest node along the perimeter. If this node can drain downhill to a location other than the lake itself, it is agged as the outlet point for all nodes on the stack. If not, it is added to the stack. If a node is encountered that is part of a pre-existing lake (one initiated at a different sink), it is also added to the stack. An example of a lake computed using the Lake Fill algorithm is shown in Figure 3-5. The algorithm is robust enough to handle any arbitrary initial condition, and is useful for modeling a rising baselevel or the damming of water and sediment behind an uplifting block. For a mesh with numerous sinks, the lake-lling algorithm is probably slower than the cascade algorithm of Braun and Sambridge (1997). However, in the more typical case of a few isolated sinks, the lake-lling algorithm is likely to be faster than the cascade algorithm. The number of iterations needed by the lake-ll algorithm depends on the number of ooded points, whereas the cascade algorithm requires a number of iterations equal to the maximum number of segments along any continuous stream regardless of the number or depth of sinks.

Numerical Algorithms
Fluvial erosion and deposition at each node in the mesh is computed numerically from
Q s j Q si j = 1 dz i = ----------------------------------dt Av i

(EQ 1)

where zi is the elevation at node i, t is time, n is the number of nodes that ow directly to i, Qs is sediment ux, and Avi is the Voronoi area of node i (the depositional area associated with each node is always its Voronoi area, regardless of the method used to compute contributing drainage area). The sediment outux Qsi depends on discharge, slope, and (possibly) sediment inux Qsj, as discussed in Chapter 2. (Note that this equation only applies to the case in which sediment ux is

Channel Hillslope Integrated Landscape Development Model

3-8

Numerical Algorithms

FIGURE 3-5. Example of lake formation in the model. The green asterisks denote lake nodes. The lake has formed in response to a digital dam that was created by articially raising the elevation of nodes near the catchment outlet. The lake outlet is indicated by the thick blue line at the right-hand edge.

tracked; in the detachment-limited end-member case, the erosion rate dz/dt is assumed to be independent of Qs, and is computed directly as a function of discharge and slope). The system of ordinary differential equations described by equation (1) is solved using a predicted-corrector method (e.g., Acton, 1970). The hillslope diffusion equation (Chapter 2, equation (7)) is solved in the following way. The downslope sediment ux per unit slope width is given by
qs = k d z , x
(EQ 2)

where the vector symbol denotes orientation in the downslope direction. The slope width between two adjacent nodes is approximated by the length of their shared Voronoi cell edge, ij. Erosion or deposition at a node due to diffusion is then approximated numerically by

Channel Hillslope Integrated Landscape Development Model

3-9

Modeling Landscape Evolution Using an Adaptive Irregular Simulation Mesh

ij S ij dz i j=1 = k d ---------------------Av i dt

(EQ 3)

where Sij is the slope from node i to node j, dened as positive downwards. This system of equations is solved using a simple forward-difference method. Note that because diffusive mass exchange takes place along mesh edges, the equation can be solved efciently by rst computing the mass exchange along each physical edge (or equivalently, across each Voronoi polygon face), then updating the node elevations accordingly.

Variable Resolution for Modeling River Meandering


Lancaster et al. (Appendix A) describe a physically-based model of stream meandering that represents a meandering stream as a series of (x,y) points. Use of an irregular, deformable mesh allows the meander model to be incorporated into the threedimensional landscape evolution model. An interface between CHILD and the meander model is currently being tested and debugged. The interface includes routines to (1) identify meandering channels on the basis of a threshold drainage area; (2) collect the nodes within a meandering channel reach and arrange them in data structures that can be passed to the FORTRAN program meander.f; and (3) move, add, and/or delete nodes in response to channel migration. Coupling the two models will make it possible to investigate the dynamics of valley and oodplain widening in an eroding landscape.

Channel Hillslope Integrated Landscape Development Model

3-10

CHAPTER 4

A Stochastic Approach to Modeling Drainage Basin Evolution

Although it is often modeled as a continuum process, landscape evolution is in fact driven by discrete events. The topography of a typical mountain range, for example, is shaped by a quasi-random sequence of oods, earthquakes, and landslides, with each process having its own characteristic frequency distribution, and with each frequency distribution perhaps varying in time and space as well. The importance of the frequency spectra of geomorphic events is a fundamental problem in geomorphic research. Much of the previous research on this problem has focussed on dening the recurrence interval of the most effective geomorphic event, with effectiveness dened either on the basis of denudation rate (e.g., Wolman and Miller, 1960; Andrews, 1980; Webb and Walling, 1982; Ashmore and Day, 1988) or landform genesis (e.g., Baker, 1977; Harvey, 1977; Wolman and Gerson, 1978). Scant attention has been paid, however, to the question of how variability in geomorphic forces impacts the morphology and rate of evolution of landforms. For example, the relative geomorphic signicance of climate variability as opposed to mean climate has been widely debated and is not well understood. In this chapter, we address the problem of rainfall variability and its impact on catchment geomorphology. We do so by developing a stochastic theory for erosion and sedimentation in a drainage basin,

Channel Hillslope Integrated Landscape Development Model

4-1

A Stochastic Approach to Modeling Drainage Basin Evolution

and exploring the consequences of that theory in the framework of the CHILD model. The stochastic model is based on the rainfall model of Eagleson (1978) and describes the probability distribution of storm depth, duration, and frequency. Combining the stochastic rainfall model with the landscape evolution model enables us to simulate the long-term geomorphic impact of natural variability in storm size. We focus in particular on two questions: What is the predicted sensitivity of long-term average sediment transport rates to the degree of variability in rainfall intensity? To what degree does the importance of variability depend on the presence of thresholds in the landscape, such as a threshold for sediment entrainment? 2. What are the morphologic consequences of rainfall variability?
1.

Model Description
Stochastic Rainfall Model

Rainfall is modeled as a series of discrete random storm events, using the model of Eagleson (1978). Each storm event is treated as having a constant rainfall rate R that lasts for a duration Td and is separated from the next event by an interstorm interval Ti. The probability density functions for storm intensity, duration, and interstorm interval are given by
Rainfall (runoff) intensity 1 P - exp -f ( P ) = - P P T d 1 - exp ----f ( T d ) = ----- Td T d Ti 1 - exp ---f ( T i ) = ---- Ti Ti
(EQ 1)

Storm duration

(EQ 2)

Interstorm period

(EQ 3)

Channel Hillslope Integrated Landscape Development Model

4-2

Model Description

In order to derive a distribution of runoff rates f(R), we start by assuming runoff is Hortonian and uniform across the landscape. (However, the formulation could be modied to account for saturation-excess runoff production [Dunne, 1978]). Runoff rate, R, is dened as precipitation minus losses to inltration, evaporation, and/or canopy interception:
R = P I, I = Ic + Il ,
(EQ 4)

where Ic = inltration capacity and Il = evaporation and canopy interception losses. The derived distribution for R is obtained from
f (R) = f (P) dP , dR
(EQ 5)

1 ( R + I ) -- exp ---------------- , R > 0; P f (R) = P 0 otherwise.

(EQ 6)

The mean runoff rate is therefore


I R = P exp -- . P
(EQ 7)

Sediment Transport by Runoff

The instantaneous rate of sediment transport by runoff is modeled as a function of excess shear stress,
Qs = k f W ( c ) ,
p

(EQ 8)

where Qs is the sediment transport rate integrated across the width of a channel or rill, W is channel width, is average bed shear stress, c is critical shear stress for sediment entrainment, kf is a transport coefcient, and p is an exponent typically on the order of 1.5-3 for bedload (e.g., Yang, 1996) and higher for suspended load

Channel Hillslope Integrated Landscape Development Model

4-3

A Stochastic Approach to Modeling Drainage Basin Evolution

(Whipple et al., in review). Assuming steady, uniform ow and adopting an empirical bed friction relationship (such as the Manning-Strickler or Darcy-Weisbach relation), shear stress can be expressed as a power function of discharge, Q, and slope, S,
Q n - S , = k t ---W
(EQ 9)

with kt, , and n as parameters. (Derivations are given by Willgoose et al., 1991; Howard et al., 1994; Tucker and Slingerland, 1997). Substituting the empirical width-discharge relationship W = k w Q ,
Qs = k f k w Q ( k t k w Q S c ) ,
m n p

(EQ 10)

where m = (1-). To write Q in terms of runoff, we use the simple steady-state relationship
Q =

R ( a ) da ,
A

(EQ 11)

where R is runoff per unit area and A is drainage area. If R is uniform across the basin, as we assume in the analysis below, this becomes
Q = RA .
(EQ 12)

This is clearly a simplication, because it neglects hydrograph attenuation by equating hydrograph duration with storm duration. These effects can in principle be accounted for by relating hydrograph duration and peak attenuation to basin geometric parameters such as total stream length, though for the sake of simplicity we do not do so here.

Channel Hillslope Integrated Landscape Development Model

4-4

Sensitivity of Erosion Rate to Rainfall Variability

Sensitivity of Erosion Rate to Rainfall Variability


Fluvial sediment transport can be viewed as a random process in time which in turn is a function of another random process, runoff. We can combine the models derived above for rainfall and sediment transport to analyze the sensitivity of long-term mean sediment transport rate to the degree of temporal variability in rainfall. We start by considering only sediment transported during storm events, which for most rivers constitutes the bulk of sediment carried. Under this condition, the mean annual sediment ux is equal to the mean transport rate produced by a storm times the mean storm duration times the number of storms per year, or
Q s = N T d Q s ,
(EQ 13)

where Q s is the long-term mean transport rate (mean annual, if Td is in years), N is the average number of storms per year, Td is mean storm duration, and Qs is the mean transport rate produced by a storm. Note that Q s can be related to mean annual rainfall. Mean annual rainfall, P , is equal to mean storm rainfall times storm frequency times mean storm duration,
P = PN T d .
(EQ 14)

Combining equations (13) and (14),


P - Qs . Q s = -------P
(EQ 15)

All else being equal, equation (15) implies that the long-term mean sediment transport rate should be linearly related to mean annual rainfall. It remains now to determine the mean storm sediment transport rate, Q s . We make the following assumptions:
1.

Each storm can be approximated as having a constant rainfall rate throughout its duration.

Channel Hillslope Integrated Landscape Development Model

4-5

A Stochastic Approach to Modeling Drainage Basin Evolution

At each point in the landscape, runoff (if nonzero) produces a constant, steady discharge equal to the runoff rate times contributing area. 3. Runoff rate per unit area has the exponential distribution given in equation (6).
2.

The steady discharge assumption allows us to write the average storm sediment transport rate as
Qs =

f ( R ) Q s ( R ) dR .

(EQ 16)

Substituting the sediment transport formula (equations (10) and (12)) and the runoff distribution (equation (6)),
k f kw A Q s = -----------------P

p ( R + I ) m m n exp ---------------- R ( k t k w A R S c ) dR . P

(EQ 17)

We do not know of an analytical solution to this equation, but an analytical solution can be found for the special case c=0, I=0. In that case,
ktk f kw A S Q s = --------------------------------------P
(1 ) np

R exp -- R dR , P

(EQ 18)

where = mp+. The integral term can be solved by the substitution u = R P , which gives
Qs = k t k f k w
(1 )

A S P

np

exp ( u ) u du = k t k f k w

(1 )

A S P ( + 1) ,

np

(EQ 19)

where () is the gamma function. Combining with equation (15) gives the expected mean annual transport rate:
Q s = P P
1

ktk f kw

(1 )

A S ( + 1) .

np

(EQ 20)

Assuming that ranges from ~1-2, as is typical, this equation predicts that long-term average sediment transport rates should generally be more sensitive to mean annual

Channel Hillslope Integrated Landscape Development Model

4-6

Sensitivity of Erosion Rate to Rainfall Variability

rainfall than to mean rainfall intensity, as long as the threshold and inltration terms are negligible.
Nondimensionalization

To facilitate numerical analysis of equation (19), the equation is nondimensionalized as follows. Rainfall and runoff intensity are normalized by mean annual rainfall, and shear stress and critical shear stress are normalized by the shear stress that would be produced by the mean annual rainfall (at a given slope and contributing area).
Nondimensional rainfall intensity Runoff rate Channel width P' = P P R' = R P Q W ' = K w ------ Q p Q W Q Q m - = ------ ------= ------ ------ = R' ' = ---Q p W p Q p Q p p c c' = ---p I I ' = ------- P W ( c ) p m Q s' = ------------------------- = R' ( R' c' ) W p p 1 I' Q s ' = ---- exp ------------- Q' I ' + R' s R'
p

(EQ 21) (EQ 22)

(EQ 23)

Shear stress

(EQ 24)

Critical shear stress

(EQ 25)

Inltration rate

(EQ 26)

Storm transport rate

(EQ 27)

Mean annual transport rate

(EQ 28)

Here, Wp refers to W(<P>), and similarly for Qp and p. Parameter m=(1-). The nondimensionalization reduces the equation to three dimensionless parameters: I, c, and either P or R (in addition to the exponent terms).

Channel Hillslope Integrated Landscape Development Model

4-7

A Stochastic Approach to Modeling Drainage Basin Evolution

Sensitivity to Runoff Variability: Numerical Solutions

In order to analyze the relationship between mean transport rate and the three parameters describing rainfall intensity, inltration capacity, and critical shear stress, the nondimensional form of equation (17) is solved using a Monte Carlo approach. For each set of parameters P', I ', and c' , 10,000 random values of P were chosen and the corresponding Qs was computed for each. The random Qs values were then averaged and multiplied by 1 P to obtain Q s ' (see equation (28)). The results are plotted in Figure 4-1.
Monte carlo solution to Qs=f(P) 3 3.5 Monte carlo solution to Qs=f(P)

2.5

(a)
Normalized <Qs>prime

(b)

Tc = 0 Tc = 1 Tc = 2 Tc = 4

2.5 Normalized <Qs>prime 2

1.5

1.5

1 I=0 I = 10 I = 100 10 20 30 40 50 Pprime 60 70 80 90 100

0.5

0.5

0 0

0 0

10

20

30

40

50 Pprime

60

70

80

90

100

FIGURE 4-1. Plot of normalized mean sediment ux Q s ' versus the rainfall variability parameter P' , showing how the relationship changes as a function of (a) inltration rate I ' and (b) critical shear stress c' . Each plotted point represents an average of 10,000 realizations of Q s' . To facilitate comparison between the curves, each curve is normalized by the mean value of Q s ' . Exponent parameters are = 0.5, m = 1 3, and p = 3 .

Figure 4-1a illustrates how the inltration parameter I inuences the relationship between mean sediment discharge and rainfall variability. With I=0 and c=0, mean sediment discharge increases as the square root of rainfall intensity, as predicted by equation (20) for =+mp=3/2. By comparison, equation (20) predicts a linear relationship between mean sediment discharge and mean rainfall, so that for low I and c, sediment discharge is more sensitive to the total amount of rainfall than to rainfall variability. As I increases, however, sensitivity to rainfall variability rises as increasingly larger rainfall events are required to produce signicant runoff.

Channel Hillslope Integrated Landscape Development Model

4-8

Morphologic Consequences of Rainfall Variability: Numerical Example

Which of these curves is most appropriate to natural catchments? A rough estimate can be obtained from published gures. Dunne (1978) reports measurements of inltration capacity on the order of 0.2 - 6 cm/hr for midwestern agricultural silt-loam soils, and on the order of 8 cm/hr for vegetated forest soils. Given a typical mean annual rainfall of ~1 meter, these values correspond to I ~ 20 - 700, implying that vegetated soils may have the effect of amplifying the importance of rainfall variability. On the other hand, this analysis does not account for direct runoff from saturated areas, which would tend to increase the importance of smaller storms. The analysis also does not account for antecedent soil moisture. In natural catchments, the distribution of interstorm periods inuences the likelihood of precipitation falling on already-saturated soils, and is therefore also a potentially important variable. The relationship between mean sediment discharge and rainfall variability is quite sensitive to the shear stress threshold c, underscoring the point made by Baker (1977) that extreme events become increasingly important in bedrock channels and/ or channels bearing very coarse bedload material. Because c depends on the meanow shear stress of the stream in question, it is difcult to judge what typical values might be. An additional complication is the nding by Parker (1978) that channel geometry in streams with mobile bed and banks tends to adjust such that the ratio / c ~ Kch = constant. Clearly, however, there exists the possibility for high sensitivity to climate variability in the case of boulder-bed or bedrock channels.

Morphologic Consequences of Rainfall Variability: Numerical Example


The consequences of rainfall variability for catchment evolution are next explored through simulations with the CHILD model. Figure 4-2 shows a synthetic catchment formed by a combination of steady tectonic uplift relative to the outlet at a rate of 2 x 10-5 m/yr and erosion by a sequence of random storm events. In this simulation, mean annual rainfall is 1 meter, mean rainfall intensity is 10 m/yr (~27.4 mm/day), and inltration capacity Ic = 10 m/yr. The topography consists of a main valley axis anked by a series of hollows (rst-order valleys).

Channel Hillslope Integrated Landscape Development Model

4-9

A Stochastic Approach to Modeling Drainage Basin Evolution

400 200 0 0 1000 2000 3000


FIGURE 4-2. Simulated drainage 4000of basin formed by a combination steady tectonic uplift and storms of variable intensity and duration.

5000 2000 6000

Figure 4-3 illustrates what happens when the catchment in Figure 4-2 is subjected to a tenfold increase in mean rainfall intensity, with mean annual rainfall held constant, for a total duration of 100,000 model years. The increased efciency of stream erosion (predicted by equation (20)) leads to accelerated erosion of the channel network (Figure 4-3b). Erosion is greatest within the rst-order valleys, which extend headward to produce a signicant increase in drainage density (Figure 4-3a). When the mean rainfall intensity later reverts to its original value (Figure 4-4), the pattern reverses as the valley network begins to ll in with sediment. The morphologic changes illustrated in Figures 4-24-4 are apparent on a plot of local slope versus contributing area (Figure 4-5). Initially, the slope-area relationship is close to the equilibrium value predicted by equation (20) (the small difference is due to the fact that the simulation has not yet reached complete equilibrium between uplift and erosion) (Figure 4-5a). The increase in rainfall intensity (Figure 4-5b)

Channel Hillslope Integrated Landscape Development Model

4-10

Morphologic Consequences of Rainfall Variability: Numerical Example

400 200 0 0 1000


FIGURE 4-3. Effect of an increase in rainfall intensity on the synthetic 2000 catchment shown in Figure 4-2. (a) Topography. (b) Pattern of erosion: red = erosion, green = little or no change, blue 3000 (not shown) = deposition or uplift. (Spike-like features in (b) are distorted Voronoi cells created by boundary 4000 effects in the triangulation).

5000 2000 6000 7000 0 1000

Channel Hillslope Integrated Landscape Development Model

4-11

A Stochastic Approach to Modeling Drainage Basin Evolution

000

000

000

000
FIGURE 4-4. Erosion and deposition pattern following a decrease in mean rainfall intensity (to its original value in the simulation show in Figure 4-2). Initial condition is the topography in Figure 4-3. Catchment is shown 10,000 model years after decrease in mean rainfall intensity. Blue = net deposition, red = net erosion, and green = little or no change. 0

leads to a decrease in gradient along the main valley network as the catchment adjusts toward a new equilibrium (dashed line). At the same time, however, gradients along the rst-order valleys and hillslopes increase in response to headward valley erosion and a consequent steepening of side-slopes. This numerical example illustrates the predicted relationship between rainfall intensity, catchment relief, and drainage density. One implication of the model is that all else being equal, higher rainfall intensity should be correlated with lower overall relief, higher drainage density, and/or higher sediment ux. The model also implies a pattern valley erosion and valley inlling, respectively, in response to variations in rainfall intensity, though the nature and pattern of response may vary depending on the magnitude of erosion thresholds (Tucker and Slingerland, 1997) and vegetation cover (Chapter 5). These predictions thus provide a basis for comparison with morphologic and sediment ux data.

Channel Hillslope Integrated Landscape Development Model

4-12

Morphologic Consequences of Rainfall Variability: Numerical Example

10

(a)

10 Numerical simulation Theory

(b)

Numerical simulation Previous equilibrium New equilibrium

Slope

10

Slope
3 4 5 6 7 8

10

10

10

10

10

10 10 Drainage area (m2)

10

10

10

10

10 10 Drainage area (m2)

10

10

FIGURE 4-5. Plot of local slope versus contributing area for the numerical simulations shown in (a) Figure 4-2 and (b) Figure 4-3. Lines show the equilibrium slope-area relationship calculated by solving equation (20) for slope, under the condition <Qs> = UA (uplift rate times drainage area).

Climatic factors such as mean rainfall and rainfall variability are often closely correlated with vegetation cover, a variable that also can clearly have important geomorphic consequences. In the next chapter, we develop a model for the interaction of vegetation and erosion and use the model to explore some of those consequences.

Channel Hillslope Integrated Landscape Development Model

4-13

A Stochastic Approach to Modeling Drainage Basin Evolution

Channel Hillslope Integrated Landscape Development Model

4-14

CHAPTER 5

Dynamics of Vegetation and Runoff Erosion

Most models of landscape evolution emphasize physical rather than biological processes, yet the biosphere, and vegetation in particular, clearly plays an important role in landscape evolution. Vegetation inuences physical erosion both directly, by increasing surface resistance to wash erosion, and indirectly, by inuencing inltration, runoff, and evapotranspiration. For example, the data of Melton (1958) show an inverse correlation between drainage density and humidity, a nding that has been attributed to the stabilizing effects of vegetation (e.g., Moglen et al., in press). By imposing a signicant threshold for runoff erosion, vegetation cover may effectively impose an upper limit to channel network extent (e.g., Horton, 1945; Montgomery and Dietrich, 1989; Dietrich et al., 1993; Kirkby, 1994). Similarly, vegetation is widely believed to contribute to the observed relationship between climate and sediment yield. A number of data sets show a decrease in sediment yield with increasing mean annual rainfall in semiarid to humid climates, despite the presumed increase in erosive energy associated with more humid climates (Langbein and Schumm, 1958; Douglas, 1967; Wilson, 1973; Moglen et al., in press). Finally, vegetation is potentially important in governing landscape responses to environmental change (e.g., Tucker and Slingerland, 1997; Howard, 1996).

Channel Hillslope Integrated Landscape Development Model

5-1

Dynamics of Vegetation and Runoff Erosion

In fully vegetated landscapes, there is a dynamic competition between vegetation growth and the disruption of vegetation by runoff erosion. Within well-established channels, runoff erosion clearly wins, while near drainage divides it is the vegetation that wins. The interesting part is what happens in between these two extremes. The outcome of the competition between erosion and vegetation growth in the vicinity of rst order channels may have a signicant effect on drainage density (in some cases, it may be the determining factor), and often also on sediment yield. In this chapter, we develop a simple theory to describe the dynamic interaction between vegetation and erosion, and use that theory to explore the nature of that interaction on geomorphic time scales (by which we mean time scales relevant to morphologic development as opposed to, say, the time scale for signicant erosion of an agricultural eld). In the rst part, we present the theory and explore some simple outcomes for slope prole development. In the second part, we present several different numerical examples using the CHILD model.

Model of Vegetation and Erosion


Conceptual model

In the context of drainage network development, one of the most important issues is the nature of the hillslope-channel transition. In humid or semi-humid landscapes, hillslopes are typically vegetation-covered, while low-order ephemeral channels contain sparse, intermittent vegetation (clumps of grass, bushes, etc.), and higherorder channels with year-round ow are free of vegetation except along the banks and bars. In deriving a model to describe vegetation-erosion dynamics, we focus on the physical interaction of vegetation and runoff erosion within rills and channels. The most important aspects of that interaction are (1) the effect of vegetation cover on soil/sediment erodibility, (2) the disruption of vegetation due to erosive overland ow, and (3) the rate at which vegetation regrows after being disrupted. Below we attempt to quantify these three interacting processes.

Channel Hillslope Integrated Landscape Development Model

5-2

Model of Vegetation and Erosion

Modeling approach Vegetation reduces erodibility

There are a number of ways in which the effects of vegetation on surface erodibility might be modeled. A simple but physically plausible approach is to assume that vegetation increases the both the effective shear stress for runoff erosion and the critical shear stress for particle entrainment. Many sediment transport equations have the form
qs = k ( c ) ,
p

(EQ 1)

where qs is sediment transport capacity per unit ow width, is uid shear stress, and c is a threshold for particle entrainment. By binding the soil with roots and leaf cover (grasses, for example), vegetation effectively increases c, an idea consistent with the threshold channel initiation hypothesis (e.g., Horton, 1945; Willgoose et al., 1990; Montgomery and Dietrich, 1989) and also with Fosters (1982) soil erosion model, which forms part of the WEPP agricultural model (Foster et al., 1995). Vegetation can be considered in terms of a fractional ground cover, V, which ranges from zero to one (or more generally from zero to V max 1 , where Vmax is the maximum percent cover that can be supported in a given environment). Few data sets exist with which to constrain the relationship between c and V (Foster 1982 provides some eld-calibrated values; Dietrich et al., 1993 give estimates of c at channel heads, based on DEM analysis). In the absence of better information, we assume a linear relationship, recognizing that this may be an oversimplication:
c ( V ) = cs + V cv ,
(EQ 2)

where cs is the critical shear stress for an unvegetated surface (primarily a function of grain size) and cv is the added critical shear stress under 100% vegetation cover. Note that this approach assumes, via equation (1), that transport capacity depends on vegetation cover. We account for changes in the total shear stress due to increased roughness and for changes in the fraction of shear stress applied to the soil surface using the approach

Channel Hillslope Integrated Landscape Development Model

5-3

Dynamics of Vegetation and Runoff Erosion

discussed by Foster (1982). The total shear stress and effective shear stress f applied to the soil are assumed to be related by
fs - , f = ---ft
(EQ 3)

with fs/ft being the ratio of the friction factor produced by the soil alone to the total friction factor (including vegetation cover). The ratio Rf=fs/ft is parameterized as a function of V, again assuming a linear relationship,
R f = k rv V , k rv 1 .
(EQ 4)

Runoff erosion disrupts vegetation

Surface runoff clearly disrupts vegetation when it becomes strong enough, but only recently have there been attempts to model the process quantitatively (Thornes, 1990; Kirkby, 1995; Kirkby and Cox, 1995). We speculate that the rate of vegetation destruction by rill or channel runoff is proportional to excess shear stress and to the fractional vegetation cover remaining:
dV (erosion) = k vd V ( c ) , dt
(EQ 5)

with kvd being the rate of vegetation loss per unit excess shear stress (or stress to a power) at 100% vegetation cover. This has the potential for a self-enhancing feedback: removal of vegetation decreases c, which in turn increases the rate of vegetation loss (but which is also compensated for by reduction in V). Solutions to (5) are sketched as a function of the dimensionless parameter G = /cv (ratio of applied shear stress to vegetation-added critical shear stress) in Figure 5-1. The humped curves reect the dual inuence of vegetation cover on critical shear stress and on vegetation erodibility; when cover is sparse, there is little vegetation to erode, but when cover is dense, the erosion potential is reduced due to the increase in c.

Channel Hillslope Integrated Landscape Development Model

5-4

Model of Vegetation and Erosion

1 0.9 Nondimensional vegetation destruction rate 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.5 1 2

0.1

0.2

0.3

0.4 0.5 0.6 % vegetation cover (V)

0.7

0.8

0.9

FIGURE 5-1. Vegetation destruction rate as a function of percent vegetation cover, for different values of G (nondimensional shear stress).

Vegetation regrowth

After vegetation is disrupted (for example, during a storm) it will begin to regrow. The rate of regrowth of vegetation biomass Vb is sometimes modeled using a logistic growth equation, in which the rate of growth approaches zero as the biomass becomes large, and also approaches zero if the biomass is very small (in which case there are few or no organisms to reproduce) (e.g., Thornes, 1990). For the present purposes, however, it is reasonable to assume that there will always be a ready supply of seeds and colonizing roots near a devegetated area (recall that V represents vegetation cover within a rill or channel; the surrounding hillsides are assumed to have V=Vmax). In view of this, a simpler linear approach seems warranted, with the rate of regrowth being proportional to the existing cover:
dV (growth) = k vg ( 1 V ) . dt
(EQ 6)

The parameter kvg is the rate of regrowth on a bare surface, and would be a function of solar radiation, soil moisture, nutrients, and so on. Note that this approach is consistent with the crop growth model used in the WEPP agriculture model (Arnold et al., 1995), which assumes a linear rate of biomass growth and an exponential rela-

Channel Hillslope Integrated Landscape Development Model

5-5

Dynamics of Vegetation and Runoff Erosion

tionship between plant cover and biomass. The cover-biomass relationship in WEPP is
V = 1 exp ( V b B )
(EQ 7)

and the biomass growth rate is


dV b d = ( B ln ( 1 V ) ) = k , dt dt
(EQ 8)

which implies that


k dV -(1 V ) . = -B dt
(EQ 9)

Equations (8) and (9) say that although the biomass can continue to grow indenitely, it eventually reaches a point where additional biomass growth does not significantly increase the fractional vegetation cover on the surface. Combining the terms for vegetation growth and erosion,
dV = k vg ( 1 V ) k vd V ( R f c ) . dt
(EQ 10)

In the next section, we explore analytical solutions to this equation. In doing so, we make the further simplications Rf=1 (i.e., vegetation does not affect total shear stress), =1, and cs=0.

Characteristic Form Proles


Nondimensionalization

It is interesting to consider what a vegetation prole across a steady-state (characteristic form) hillslope might look like, and how the presence of vegetation inuences the shape of the hillslope. We can obtain a steady-state vegetation prole by solving

Channel Hillslope Integrated Landscape Development Model

5-6

Characteristic Form Proles

equation (10) for the case dV/dt=0, and solving equation (1) for the case qs = Ex, where E is an erosion rate that is constant along the prole. To facilitate the analysis, we introduce the following nondimensionalization:
Vegetation growth time scale Nondimensional time Nondimensional shear stress Nondimensional distance along slope prole Vegetation number T v = 1 k vg t' = t T v ' = cv x' = x L, k vd cv N v = ------------k vg L = slope length
(EQ 11) (EQ 12) (EQ 13) (EQ 14)

(EQ 15)

The vegetation number represents the efciency of vegetation growth relative to destruction; low values represent fast-growing and/or destruction-resistant vegetation, and vice-versa. Introducing these dimensionless numbers into equation (10), we have
dV 2 = N v V ( 1 + N ' ) V + 1 , d t'
(EQ 16)

which is a quadratic equation that can be solved if dV/dt is constant or zero.


Wash Prole

The equilibrium vegetation prole depends in part on the rate at which it is being disrupted by runoff erosion, which may vary along a slope. First we consider the case of a hillslope prole that is eroded solely by wash (equation (1)), at a uniform rate E such that qs = Ex. Under that condition, the equilibrium shear stress is
' = N E x' + V ,
1 -p 1 -p

LE N E = --------k cv

(EQ 17)

Substituting this into equation (16) yields

Channel Hillslope Integrated Landscape Development Model

5-7

Dynamics of Vegetation and Runoff Erosion

1 V = -----------------------------, q q 1 + N v N E x'

1 q = --. p

(EQ 18)

Similarly, we can solve for equilibrium slope along the prole using the shear stress relationship
= k t L x' S .
m m n

(EQ 19)

Solving for the equilibrium condition,


NE q m 1 - x' S = -----+ -----------------------------------------------q q m Nt N x' ( 1 + N N x' )
t v E q 1n

(EQ 20)

where Nt = ktLm/cv. Figure 5-2 depicts solutions to (18) and (20) for different values of the dimensionless parameters Nv, NE, and Nt, assuming typical values of m=2/3, n=2/3, q=1/p=2/3 (e.g., the Meyer-Peter and Mueller relation) or q=1/p=1/3 (the Einstein-Brown relation). Vegetation cover decreases downslope in response to the increasing shear stress, asymptotically approaching zero as x becomes large. This vegetation gradient has the effect of increasing the concavity of the hillslope prole, particularly when the erosion number NE is small relative to the other parameters (compare Figure 5-2, top and bottom). The vegetation number Nv controls the downslope rate of reduction in vegetation cover. When Nv is large (Figure 5-2, middle), there is an abrupt reduction in vegetation cover in the upper part of the prole that corresponds to a sharp concavity in the topography.
Diffusive Prole

The wash model ignores the effect of creep-related (slope-dependent) processes that would tend to produce convex-upward slopes near a drainage divide (e.g., Gilbert, 1909). The prole shape under the action of both creep and wash processes cannot be solved analytically, but we can obtain some idea of what a vegetation prole might look like on a diffusion-dominated slope by solving for the characteristic-form slope as if diffusion were the sole process. Essentially, we assume that wash controls the amount of vegetation but does not contribute signicantly to shaping the topography.

Channel Hillslope Integrated Landscape Development Model

5-8

Characteristic Form Proles

Characteristic form wash profile with vegetation 1 0.9 Nondimensional elevation / % vegetation 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 Slope profile Vegetation profile Slope profile w/o veg Nondimensional elevation / % vegetation 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.1 0.2 0.3 0.4 0.5 0.6 Nondimensional distance 0.7 0.8 0.9 1 0 0

Characteristic form wash profile with vegetation (p=3) Slope profile Vegetation profile Slope profile w/o veg

0.1

Characteristic form wash profile with vegetation (Nv=10) 1 0.9 Nondimensional elevation / % vegetation 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 Slope profile Vegetation profile Slope profile w/o veg

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Nondimensional distance Characteristic form wash profile with vegetation (Nv=4, p=3) Slope profile Vegetation profile Slope profile w/o veg

1 0.9 Nondimensional elevation / % vegetation 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

0.1

0.2

0.3

0.4 0.5 0.6 Nondimensional distance

0.7

0.8

0.9

0.1

0.2

0.3

0.4 0.5 0.6 Nondimensional distance

0.7

0.8

0.9

Characteristic form wash profile with vegetation (Ne=10) 1 0.9 Nondimensional elevation / % vegetation 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 Slope profile Vegetation profile Slope profile w/o veg Nondimensional elevation / % vegetation 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

Characteristic form wash profile with vegetation (Ne=10, p=3) Slope profile Vegetation profile Slope profile w/o veg

0.1

0.2

0.3

0.4 0.5 0.6 Nondimensional distance

0.7

0.8

0.9

0.1

0.2

0.3

0.4 0.5 0.6 Nondimensional distance

0.7

0.8

0.9

FIGURE 5-2. Solutions to equations (18) and (20) for wash-dominated hillslope proles, showing the effect of vegetation on prole shape.

Channel Hillslope Integrated Landscape Development Model

5-9

Dynamics of Vegetation and Runoff Erosion

Assuming that the transport rate by creep is given by qs = kd S, the characteristic form prole is
UL - x' = N d x' . S = ------kd
(EQ 21)

Substituting into equation (19), the shear stress is given by


kt L N d m + n m+n - x' ' = ----------------= N dt x' . cv
m n

(EQ 22)

Combining with (16) (for dV/dt=0) gives the vegetation prole


N v V ( 1 + N v N dt x'
2 m+n

)V + 1 = 0 .

(EQ 23)

Example solutions to (21) and (23) are shown in Figure 5-3. Note that equation (23) has two roots. In the example shown in the gure, one of the roots implies full vegetation cover along the prole, while the other implies decreasing cover. This example of bi-stability is similar to the bistable landscape states explored by Howard (1996), as well as the multiple phase-states in the soil erosion model of Thornes (1990). Depending on the initial conditions, the system may evolve toward a different equilibrium state. This analytical example of a diffusion-dominated prole is of course rather articial because it does not incorporate the linkage between slope form and

Nondimensional elevation / Vegetation cover %

0.8

0.6

0.4

FIGURE 5-3. Solutions to equation (23) for equilibrium vegetation cover on a diffusion-dominated hillslope on which vegetation growth is limited by wash erosion.

0.2 Hillslope profile Vegetation (solution 1) Vegetation (solution 2) 0 0 0.1 0.2 0.3 0.4 0.5 0.6 Nondimensional distance 0.7 0.8 0.9 1

Channel Hillslope Integrated Landscape Development Model

5-10

Numerical Examples

wash erosion. To explore the complete interaction of wash, diffusive transport, and vegetation growth, numerical simulations are needed.

Numerical Examples
Figure 5-4 shows the topography and percent vegetation cover in a synthetic catchment simulated using the CHILD model. The basin is close to a state of dynamic equilibrium between uplift and erosion. Vegetation cover ranges from 100% on the hillslopes to very low values in the main channels. (The presence of a small vegetation cover within the channels is due to the fact that vegetation always regrows during interstorm periods, regardless of position within the landscape). In general, the downstream transition from full to sparse vegetation cover is a function of two factors: (1) the threshold imposed by the vegetation itself, which retards runoff erosion on upper slopes (as seen in Figure 5-2), and (2) the rounding of hillslope proles due to diffusion, which reduces slopes and thus shear stresses (as in the extreme case of Figure 5-3). In this particular example, the hillslope diffusivity constant kd is sufciently high that the latter effect is more important. (a)
400 200 0 0 1000 2000 3000 4000 5000 2000 6000 7000 0 1000

(b)

FIGURE 5-4. Simulated catchment with variable vegetation cover. (a) Topography. (b) Percent vegetation cover. White = 100% cover, black = 0% cover.

Channel Hillslope Integrated Landscape Development Model

5-11

Dynamics of Vegetation and Runoff Erosion

4000

3000

2000

FIGURE 5-5. Short term-effect of a decrease in cv on percent vegetation cover in the synthetic catchment shown in Figure 1000 5-4.

Figure 5-5 illustrates the effect of a hypothetical decrease in the maximum sustainable vegetation cover. The decrease in sustainable vegetation maximum is modeled by a tenfold decrease in the threshold parameter cv. The decrease in cv is accompanied by an increase in the effectiveness of runoff erosion, which leads to an upslope extension of the sparsely-vegetated tributaries. Vegetation cover typically inuences not only surface resistance to erosion, but also the soil inltration capacity and hence runoff production. Figure 5-6 shows a hypothetical example in which the synthetic catchment shown in Figure 5-4 is subjected to a complete loss in vegetation accompanied by a reduction in inltration capacity (Ic = 0). The gure depicts areas of erosion (red), deposition (not shown), and little or no change (green). The net effect of vegetation loss in this example is an increase in the rate of scour along the main channels. This result is somewhat contrary to intuition, which would suggest that a loss of vegetation should lead to accelerated erosion on hillslopes, where the initial vegetation cover (and thus the threshold) was greatest (this type of behavior was observed by Tucker and Slingerland (1997) in experiments with a constant-threshold model). The explanation appears to lie in the topography of the catchment and the nature of the hillslope-valley transition. With a relatively high diffusivity constant kd, the valley network extent in the simulation is limited by hillslope diffusion rather than by the threshold imposed by vegetation cover. Thus, reducing the threshold does not lead to a rapid expansion of the valley network, as it did in the constant-threshold experiments of Tucker and Slingerland

Channel Hillslope Integrated Landscape Development Model

5-12

Numerical Examples

4000

3000

2000 FIGURE 5-6. Patterns of erosion (red=max. erosion, green=no change) following a complete loss of vegetation and a corresponding 1000 reduction in inltration capacity in the synthetic catchment shown in Figure 5-4. 0

(1997). An implication of this result is that the nature of catchment responses may depend to a large extent on the nature of hillslope-valley transitions, and in particular on whether that transition is governed by a process transition or by a vegetationimposed erosion threshold. There is still much to be learned about what controls the type of response in the model and, by extension, in natural catchments. We anticipate that this question can be fruitfully addressed through a program of systematic model sensitivity experiments and comparison of the model predictions with the Holocene stratigraphic record observed at Fort Riley, Kansas.

Channel Hillslope Integrated Landscape Development Model

5-13

Dynamics of Vegetation and Runoff Erosion

Channel Hillslope Integrated Landscape Development Model

5-14

Chapter 6 Understanding the Interactions of Multiple Grain Sizes

6-1

Appendix A Meandering: A Simple Model

A-1

You might also like