Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and...

18
Numerische Methoden 1 – B.J.P. Kaus Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes how erosional processes control the shape of mountains. An interesting question is how the physics of erosion works, and whether erosion controls tectonic processes or whether tectonics controls erosion. There is currently much research done to better understand the physics of erosion. There is general agreement that there are several processes that contribute to erosion. One is ofcourse fluvial erosion, which basically depends on the amount and the speed of water that flows over a landscape (the larger the river, the more its erosional capacity). Another process is hillslope diffusion, which is basically a diffusional processes. Arguably, one of the simplest ways to model erosion is to use a 2-D diffusion-like equation: ∂h ∂t = ∂x k ∂h ∂x + ∂y k ∂h ∂y (1) where h is the elevation of the surface and k the diffusivity, which is now spatially variable. We can include both hillslope and fluvial erosion in k by defining it as: k = k 0 + c * q n (2) where k 0 is the hillslope diffusivity [m 3 /s], n a powerlaw exponent (typically 2), c a constant and q the fluvial discharge rate [m 2 /s], which basically contains all the collected rainwater that is flowing down the landscape. Once we have q at every point in our model, solving equation (1) is the same as solving a 2D diffusion equation with variable conductivity (which we discussed earlier). The tricky part is how to compute q, which we will do here with a simple algorithm. Imagine that a certain amount of rain falls on your landscape (say 1 cm/yr). Water will always flow to the lowest point around it. So the water-stream algorithm to compute q consists of finding the lowest neighbour of each point in your 2D grid. There will be some cases in which there is no lower neighbour, which will be locations where a lake fill form (so the neighbour is the same point). Once you have a list with the lowest neighbour, you give a certain amount rain to every point in your domain. Start with the highest point, and move the rain to it’s lowest neighbour and continue until you have distributed the water over all points. This algorithm can be done in a few lines of matlab, and will give q. 1

Transcript of Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and...

Page 1: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Numerische Methoden 1 – B.J.P. Kaus

Project: Modelling of erosion and landscape formation

Anyone who has visited a mountain-belt such as the European alps realizes how erosional processescontrol the shape of mountains. An interesting question is how the physics of erosion works, and whethererosion controls tectonic processes or whether tectonics controls erosion. There is currently much researchdone to better understand the physics of erosion. There is general agreement that there are severalprocesses that contribute to erosion. One is ofcourse fluvial erosion, which basically depends on theamount and the speed of water that flows over a landscape (the larger the river, the more its erosionalcapacity). Another process is hillslope diffusion, which is basically a diffusional processes. Arguably, oneof the simplest ways to model erosion is to use a 2-D diffusion-like equation:

∂h

∂t=

∂x

(k

∂h

∂x

)+

∂y

(k

∂h

∂y

)(1)

where h is the elevation of the surface and k the diffusivity, which is now spatially variable. We caninclude both hillslope and fluvial erosion in k by defining it as:

k = k0 + c ∗ qn (2)

where k0 is the hillslope diffusivity [m3/s], n a powerlaw exponent (typically 2), c a constant and q thefluvial discharge rate [m2/s], which basically contains all the collected rainwater that is flowing down thelandscape. Once we have q at every point in our model, solving equation (1) is the same as solving a 2Ddiffusion equation with variable conductivity (which we discussed earlier).

The tricky part is how to compute q, which we will do here with a simple algorithm. Imagine thata certain amount of rain falls on your landscape (say 1 cm/yr). Water will always flow to the lowestpoint around it. So the water-stream algorithm to compute q consists of finding the lowest neighbour ofeach point in your 2D grid. There will be some cases in which there is no lower neighbour, which willbe locations where a lake fill form (so the neighbour is the same point). Once you have a list with thelowest neighbour, you give a certain amount rain to every point in your domain. Start with the highestpoint, and move the rain to it’s lowest neighbour and continue until you have distributed the water overall points. This algorithm can be done in a few lines of matlab, and will give q.

1

Page 2: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Numerische Methoden 1 – B.J.P. Kaus

Exercise

• Define a landscape with a given slope (use an area of 20x20 km’s, with a maximum height of 1000m) and some random noise on top of it (use the matlab command ’rand’).

• Develop a stream-flow algorithm that distributed the water to the next lower point (see descriptionabove). Make a 2D plot of q. Can you see river-like patterns already

• Modify your 2D diffusion code to make it work with variable conductivity, and combine it with yourstream-flow algorithm. Make a forward run in time. Do you see an evolving eroding landscape?Use c = 100, n = 2, k0 = 3.2× 10−8

• Experiment with different values of k0, different amplitudes of random noise, and different slopes.

2

Page 3: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Topographic evolution and morphology of surfaces evolving in

response to coupled fluvial and hillslope sediment transport

Guy Simpson and Fritz Schlunegger1

Geology Institute, ETH Zurich, Zurich, Switzerland

Received 21 August 2002; revised 21 January 2003; accepted 11 March 2003; published 14 June 2003.

[1] This paper quantifies how the ratio of sediment transport on hillslopes to sedimenttransport in channels influences surface and channel network morphologies and thedynamics of topographic evolution. This problem is investigated by development andinvestigation of a simple deterministic model incorporating mass balance of sediment andrunoff coupled with a law combining dispersive and concentrative sediment transportprocesses. Our analysis includes the identification of a new nondimensional parameter De

that is a function of rainfall, system size, rock type, and hydraulic regime and that is ameasure of the relative importance of fluvial and hillslope sediment transport. We showthat De has an important influence on the surface morphology (e.g., total exposed surfacearea and interface width which reflects surface roughness and relief), channel networkform (e.g., channel sinuosity), channel spacing, and timescale of surface evolution.Surface and channel network morphologies are also strongly influenced by the overallsurface slope relative to the magnitude of initial topographic roughness. Topographicevolution occurs in distinct phases of relief growth and decay, the transition betweenwhich is controlled by a saturation phenomenon related to the growth of spatialcorrelations. The scaling behavior of simulated topography with respect to both time andspace is obtained and is shown to be independent of De. Roughness exponents are found tobe independent of De but dependent on the magnitude of initial roughness. Interface widthis shown to grow and decay as a logarithm of time. INDEX TERMS: 1625 Global Change:

Geomorphology and weathering (1824, 1886); 1815 Hydrology: Erosion and sedimentation; 1860 Hydrology:

Runoff and streamflow; 3210 Mathematical Geophysics: Modeling; KEYWORDS: erosion, landscape,

topography, drainage network, diffusion, FEM modeling

Citation: Simpson, G., and F. Schlunegger, Topographic evolution and morphology of surfaces evolving in response to coupled

fluvial and hillslope sediment transport, J. Geophys. Res., 108(B6), 2300, doi:10.1029/2002JB002162, 2003.

1. Introduction

[2] In recent years there has been widespread apprecia-tion that some important aspects of landscape morphologyand dynamics result from interaction between hillslopes andchannel networks. Recognizing this, a number of numericalmodels that incorporate coupling between hillslope andchannel processes have been developed [e.g., Kirkby,1986; Ahnert, 1987; Willgoose et al., 1991a; Chase, 1992;Howard, 1994]. These models have been used to explorevarious aspects of landscape development including generaltopographic evolution [e.g., Willgoose et al., 1991b, 1991c;Kooi and Beaumont, 1994; Tucker and Slingerland, 1994],response to climate change [Rinaldo et al., 1995; Tuckerand Slingerland, 1997] and tectonics [Beaumont et al.,1992; Willet et al., 2001], and landscape self-organization[Rigon et al., 1994]. To date, however, relatively little

detailed research based on numerical modeling has beencarried out on the important aspects concerning the couplingitself that exists between hillslope and channel processes(see, however, Tucker and Bras [1998]). Thus a number ofinteresting and important general questions remain unre-solved. For example, how does the relative importance ofchannel and hillslope sediment transport influence land-scape and network morphologies and the timescales oflandscape evolution? What parameter(s) controls channelspacing? Why are some channel networks regular with awell defined channel spacing whereas others are highlybranching? This paper uses numerical modeling as a tool toexplore several problems such as these.[3] In general terms, the morphology of natural land-

scapes (that are not subject to lithospheric deformation)reflects the strength and nature of climatic forcing relative tothe ability of topography to relax by diffusion. In atransport-limited system where erosion rates are proportion-al to the divergence of sediment flux, runoff from rainfallhas the potential to roughen topography by channel incisiondue to the strong dependency of sediment flux on fluiddischarge [Graf, 1971]. This generally concentrative pro-cess is counteracted by sediment transport proportional to

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B6, 2300, doi:10.1029/2002JB002162, 2003

1Now at the Institute of Geological Sciences, University of Bern, Bern,Switzerland.

Copyright 2003 by the American Geophysical Union.0148-0227/03/2002JB002162$09.00

ETG 7 - 1

Page 4: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

the local surface gradient [Culling, 1960] that tends to filldepressions and smooth topography. The presence of twocompeting classes of processes, one concentrative (anddominant in channels) associated with runoff, which causespositive feedback and the other dispersive (and dominant onhillslopes) causing negative feedback, are recognized tocomprise the basic minimum requirements to explain theformation of landscapes containing convex hillslopes andincised channel networks.[4] Importance of the balance between dispersive and

concentrative transport processes in influencing landscapedevelopment is widely appreciated on the basis of field-based and experimental studies [e.g., Schumm, 1956; Dunneand Aubry, 1986; Montgomery and Dietrich, 1992; Has-bargen and Paola, 2000] and has been the focus oftheoretical and numerical investigations [e.g., Smith andBretherton, 1972; Willgoose et al., 1991a, 1991b; Howard,1994; Loewenherz-Lawrence, 1994; Smith et al., 1997a,1997b; Tucker and Bras, 1998]. For example, it is nowgenerally accepted that instability leading to incision of aninitially unchannelled surface is associated with a domi-nance of concentrative over dispersive processes in thetransport of surface sediments. Although it was initiallyshown on the basis of a linear stability analysis that theshortest transverse topographic wavelengths are the fastestgrowing [Smith and Bretherton, 1972], implying that chan-nels should be spaced an infinitesimal distance apart, it hassince been demonstrated that nonlinear effects governsubsequent channelization leading to stable topographicsurfaces with finite channel spacings [e.g., Smith et al.,1997a]. An alternative interpretation is that the failure oflinear models to select a finite channel spacing is an artifactdue to oversimplify the equations governing surface fluidflow. For example, Izumi and Parker [1995, 2000] showthat a linear analysis of the full equations of shallowoverland flow coupled with erosion leads to the selectionof finite channel wavelengths in reasonable agreement withobservations.[5] It is also generally accepted that the balance between

dispersive and concentrative sediment transport processes isreflected in the general morphology of landscapes [e.g.,Montgomery and Dietrich, 1992; Kirkby, 1994]. For exam-ple, highly incised, rough landscapes such as those found insemiarid badlands are thought to be formed by dominantlyconcentrative processes whereas weakly incised, smoothlandscapes characteristic of more temperate humid settingsare believed to be formed dominantly by dispersive sedi-ment transport processes. However, not all hillslope pro-cesses are necessarily dispersive in nature (e.g., some formsof mass movement lead to concentrated deposits). A largeamount of theoretical work has been focused on under-standing how interaction between hillslope and channelprocesses determines landscape properties such as thecontributing area-slope relationship and the drainage density[e.g., Tarboton et al., 1992; Willgoose et al., 1991d; Will-goose, 1994; Tucker and Bras, 1998]. Results of thesestudies are generally consistent with the concept of a ‘‘beltof no fluvial erosion’’ [Horton, 1945] extending somedistance xc down from drainage divides which is controlledby competition between hillslope and water-assisted sedi-ment transport processes. Scaling at distances smaller thanxc is controlled by diffusive processes whereas scaling at

distances larger than xc is controlled by fluvial processes.The distance xc is also related to the drainage density[Tarboton et al., 1992]. Thus the ratio between the magni-tudes of hillslope to fluvial sediment transport is anticipatedto strongly influence the drainage density. This inference isconsistent with field observations showing that whereasbadland drainage densities are typically on the order of1000 km/km2, those of temperate landscapes are typically�1 km/km2 [Kirkby, 1986].[6] The main objective of this paper is to quantify how the

ratio of sediment transport on hillslopes to sediment trans-port in channels influences the overall landscape morphol-ogy and dynamics. As indicated above, some aspects of thistopic have already received investigation in the literature.This paper extends and unifies these studies within aconsistent framework by clearly identifying any importantcontrolling parameters and by systematically studying howthe landscape morphology and dynamics vary as a functionof these parameters. Our analysis includes the identificationof a new nondimensional parameter De that is a function ofrainfall, system size, rock type and hydraulic regime and thatis a measure of the relative importance of fluvial andhillslope sediment transport. We show that De has animportant influence on the surface morphology, channelnetwork form, channel spacing (or drainage density) andthe time scale of surface evolution. We quantify results byanalyzing exposed surface area (reflecting surface rough-ness), channel sinuosity (reflecting how the channel networkis organized on the surface), interface width or root-mean-square height fluctuation (reflecting roughness and relief),and crossover length scales (reflecting the average channelspacing). The scaling behavior of simulated topography withrespect to both time and space is obtained. In addition,surface and channel network morphologies are shown to bestrongly influenced by the overall surface slope relative tothe magnitude of initial topographic roughness.[7] The approach taken in this study is to numerically

solve and investigate solutions of equations that are basedon the conservation of surface water and sediment com-bined with a simple transport law connecting the sedimentflux to the local slope and runoff discharge. This formula-tion is due largely to Smith and Bretherton [1972] and issimilar to the equations studied by Cordova et al. [1982],Loewenherz-Lawrence [1994], and Smith et al. [1997a,1997b]. The governing equations are solved here by thefinite element method.

2. Governing Equations

[8] The formulation considered consists of two equationsfor the two unknown functions h(x, y, t) and q(x, y, t), thesurface elevation and the magnitude of the surface fluiddischarge per unit width, respectively [see Smith andBretherton, 1972]. These equations are derived by assumingthat the masses of moving sediment and surface water areconserved:

@h

@t¼ �r � ðn qsÞ ð1Þ

r � ðn qÞ ¼ a; ð2Þ

ETG 7 - 2 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 5: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

where n (=�r h/S) is a unit vector directed down the surfacegradient, S is the local slope of the surface (= jr hj), a (>0and assumed to be spatially constant) is the effective rainfall(the average rainfall in excess of infiltration) and qs is thesediment discharge per unit width (see Table 1 for notation).Note that qs and q are the vector fields nqs and nq,respectively. The symbolsr � andr represent the divergenceand the gradient operators, respectively. Equation (1)indicates that the time rate of change in surface elevation isbalanced by the divergence of the sediment discharge, whileequation (2) indicates that fluid from steady rainfall flows offthe surface as a gravity-driven flow directed down the localgradient. Neglection of the time derivative in equation (2) ismade on the basis of the observation that the timescale ofoverland flow of runoff is generally far faster than thetimescale associated with transport of topography (sediment)[see also Smith et al., 1997b]. Relative base level changes(not included here) can be incorporated in the formulationeither by including a source term in equation (1) or by directlyconstraining h on the desired boundary.[9] The constitutive law for the sediment discharge can

have a variety of forms depending on the geomorphologic/hydrologic setting and numerous different variations havebeen proposed [e.g., Culling, 1960; Andrews and Bucknam,1987; Graf, 1971; Smith and Bretherton, 1972; Kirkby,1986]. The approach taken here is to use one of the simplestphysically reasonable laws that enable simultaneous devel-opment of both diffusive hillslopes and an incisive drainagenetwork and to apply this law to the entire surface withoutdistinction or prespecification of channel and hillsloperegions. This law can be written as

qs ¼ kS þ cqnS ð3Þ

where k is the hillslope diffusivity, c is the fluvial transportcoefficient, and n is a power law exponent quantifying thedependency of sediment transport on fluid discharge. Thefirst term represents the widely accepted dependency ofsediment flux on local slope [Culling, 1960] whereas thesecond term is included to account for fluvial transport [Graf,1971]. Note that by proposing this transport law, it is assumedthat the system is transport-limited, meaning that there is nolimit imposed by the supply of sediment (see discussions byCulling [1965], Carson and Kirkby [1972], and Howard[1994]). Note also that no distinction is made here betweensuspended and bed load transport. In this sense, the fluvialterm in equation (3) is intended to approximate the totalfluvial load [see also Smith and Bretherton, 1972]. Manytotal transport laws assume a nonlinear dependency ofsediment flux on slope. Inclusion of such a dependency inequation (3) is likely to modify surface and channelproperties (e.g., channel concavity) but will lead to the sametype of general behavior (i.e., nonlinear diffusion). Anobvious and easily justifiable addition to equation (3) wouldbe to include a threshold below which discharge-assistedsediment transport is negligible, for which considerableexperimental and theoretical support exists [Graf, 1971]. Asecond possible addition would be to include a term wherebythe hillslope sediment flux increases rapidly as a critical slopeis approached, which has been suggested to capture somebasic aspects of landsliding [see Andrews and Bucknam,1987; Roering et al., 2001].

[10] It is noteworthy that the adopted transport law iswritten in terms of sediment discharge per unit width qs (m

2

s�1) rather than the total sediment discharge Qs (m3 s�1),

the two being related by Qs =R0WqsdW (where W is the

channel width). This formulation is useful because it avoidsthe explicit introduction of a channel width and is naturalgiven the form of the continuity equation for fluid (alsowritten in terms of discharge per unit width). The reader isreferred to Tucker and Slingerland [1997] for an example ofa more typical formulation involving explicit relationshipsbetween water flux, channel width and water depth.[11] Substituting equation (3) into equation (1), the sys-

tem of governing equations can be rewritten as

@h

@t¼ r � ððkþ c qnÞrhÞ ð4Þ

r � rh

jrhj q� �

¼ �a: ð5Þ

Equation (4) is recognized as a transient nonlinear diffusionequation while equation (5) is a steady nonlinear advectionequation. Note that the grouping k + c qn in equation (4) hasthe role and dimensions of a diffusion coefficient (that canvary in time and space). The entirely diffusive nature of (4)is reflected by the terminology adopted within this studywhereby transport processes have been separated fordiscussion purposes into ‘‘dispersive’’ and ‘‘concentrative’’components [see also Howard, 1994], as opposed to theterms ‘‘diffusive’’ and ‘‘advective’’ more commonly used inthe literature. The dispersive and concentrative processesare sometimes also simply referred to as ‘‘hillslope’’ and‘‘fluvial’’, respectively, due to their respective dominance inthese landscape settings.[12] This study treats the governing equations in their

nondimensional form. This has the advantage that the

Table 1. Notation

Symbol Unit Definition

h m surface elevationq m2 s�1 surface fluid dischargeqs m2 s�1 sediment dischargex,y m horizontal coordinatest s timeS – slope of local topography (= |rh|

=ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið@h=@xÞ2 þ ð@h=@yÞ2

a m s�1 effective rainfallk m2 s�1 substrate diffusion coefficientc (m2 s�1)/(m2 s�1)n coefficient for discharge-dependent

sediment transportn – exponent for dependency of fluid

discharge on sediment transportL m characteristic horizontal dimension

(slope width perpendicular to ridge)~t – dimensionless time (= t k/L2)~x;~y – dimensionless distances (= x/L, y/L)~h – dimensionless elevation (= h/L)~q – dimensionless fluid discharge (= q/(a L))De dimensionless diffusion parameter

(= c anLn/k)hr – maximum initial value of random noiseH – initial topographic height at ridge axisl – dimensionless roughness (= hr/H)

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 3

Page 6: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

number of independent parameters appearing in the govern-ing equations can be reduced from four (i.e., k, c, a, and n) totwo, which greatly facilitates sensitivity analysis. Naturalscales for nondimensionalization of equations (4) and (5) arethe linear diffusion timescale t*, the characteristic lengthscale L (e.g., horizontal slope length), and the fluid dischargeat the characteristic length scale Q*. One may introduce

t ¼ ~t t* ¼ ~t L2=k;

h ¼ ~h L; x ¼ ~x L; y ¼ ~y L;

q ¼ ~qQ* ¼ ~qaL;

where dimensionless quantities are indicated by tildesymbols. Substituting these expressions into equations (4)and (5) yields the dimensionless equations

@~h

@~t¼ ~r � ðð1þ De ~q

nÞ ~r ~hÞ ð6Þ

~r �~r~h

j ~r~hj~q

!¼ �1 ; ð7Þ

and the dimensionless parameter

De ¼can Ln

k: ð8Þ

The parameter De is a measure of the relative importance ofdispersive processes related to hillslope creep versusconcentrative processes related to water-assisted transportacting on the landscape. If De 1, nonlinear diffusion dueto the dependency of the diffusion coefficient on runoffdominates over linear diffusion associated with hillslopecreep. If De � 1 the dependency of the diffusion coefficienton runoff is negligible and topographic evolution reduces toa linear equation controlled by hillslope creep. An alternativeinterpretation of De can be made by defining the character-istic timescales of hillslope and fluvial processes as th = L2/kand tf = L2/(c anLn), respectively, where De can be writtenas the ratio of the two timescales th/tf. Note that De is notan erosion Peclet number since there is strictly noadvective sediment transport. Similar nondimensionalformulations for erosion have been presented by Smithand Bretherton [1972], Willgoose et al. [1991b], and Smithet al. [1997a]. A parameter very similar (though not exactlyequivalent) to De was derived by Willgoose et al. [1991b,equation 27]. Differences between the two nondimensionalparameters are due to differences in the governingequations and in the characteristic scales adopted.

3. Numerical Method and Approach

[13] Equations (6) and (7) form a closed system ofcoupled nonlinear equations that are solved in a classicalmanner by the finite element method on an irregulartriangular mesh [Zienkiewicz and Taylor, 2000a]. Equation(6) is discretized using the Galerkin method with cubic

shape and weighting functions using 7-node triangles.Equation (7) is discretized on 3-node triangles using linearshape functions and streamline (upwind) Petrov-Galerkinweighting [Zienkiewicz and Taylor, 2000b] in order tosuppress numerical oscillations associated with strong ad-vection. This procedure involves the addition of anisotropicdiffusion to regions where the Peclet number locallyexceeds unity, which has the effect of preventing infiniteconcentration of flow in channels, a problem known to existwith equation (7) [see Izumi and Parker, 1995]. The addedartificial diffusion mimics real physical processes (e.g.,eddy viscosity) that have similar stabilizing effects in theNavier-Stokes equations but which have been eliminated inthe simplified formulation adopted. Locations with zerolocal slopes are singular (see equation (7)) and are treatedby setting the fluid flux to zero. Discretization in time wasperformed using linear explicit finite differences whilenonlinearities in both equations are treated by laggingsolution-dependent coefficients by one time step.[14] All results presented were performed on a two-

dimensional spatial domain discretized by �20,000 ele-ments. This number proved to be sufficient in order toresolve the major channel structures but is unlikely tocapture structure beyond channel orders of 4–5. Higher-resolution simulations capable in future studies are antici-pated to reveal more fine-scale channel structure andsharper channel-hillslope boundaries but are not anticipatedto influence landscape dynamics. Dynamic behavior (suchas anastomosing channel networks and stream captureevents) observed in simulations (see section 4.1) persist asthe time step increment is decreased, indicating that thesefeatures are unlikely to be numerical artifacts.[15] Note that a far more commonly applied approach to

solving the general coupled erosion problem is to use adiscrete routing algorithm [e.g., Chase, 1992]. While manyvariants of this method have been developed and applied thegeneral strategy involves progressively routing material(both sediment and fluid) down the computed drainagenetwork between nearest discrete neighbors from the high-est to the lowest elevation. This strategy differs from thecontinuum approach adopted here [see also Smith et al.,1997b] in which sediment and fluid implicitly move downthe local surface gradient. The discrete routing approach hasan advantage over the continuum approach in that it iscomputationally efficient but it suffers from a grid depen-dency introduced by routing between a single nearestneighbor. In principal, this numerical artifact could beeliminated by redistributing fluid among neighboring cellsaccording to the local gradient. An alternative approach is tomask the artifact using high resolution and/or an irregularspatial grid [Braun and Sambridge, 1997]. The continuumapproach has the advantage that grid dependency is reducedand that the routing of sediment and fluid is achievedautomatically as a result of the discretization and solutionprocess rather than manually on the basis of some user-defined redistribution algorithm. The disadvantage of theadopted approach is greater computational expense.

4. Results

[16] Numerical solutions of equations (6) and (7) werecomputed to investigate erosion of, and runoff over a

ETG 7 - 4 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 7: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

simple, initially unchannelled, tilted, planar topographicsurface perturbed by gaussian random noise (Figure 1).Boundary conditions consist of (1) a symmetrical ridgecrest with zero water and sediment flux, (2) zero sedimentand water flux normal to lateral boundaries, and (3) fixedtopographic elevation at the slope base. Note that theseboundary conditions imply that the system is open withrespect to transport of sediment and fluid.[17] Attention is focused on landscape development and

morphology as a function of both time and De with constantvalues for the power exponent n (= 2). This value, which isconsistent with experimental data [see Graf, 1971], impliesthat sediment discharge is strongly dependent on the runoffdischarge. Note that when n � 1, channelization does notoccur in the continuum formulation because the relativesediment carrying capacity does not increase when initialperturbations induce flow convergence [see Smith andBretherton, 1972]. A number of studies have shown thatvarying n (or similarly the area-slope exponent) has a largeinfluence on valley profile concavity and on the generallandscape form [e.g., see Tucker and Whipple, 2002].[18] The morphology of simulated surfaces is quantified

by determining the root-mean-square height fluctuation orinterface width ~w (reflecting the surface roughness, reliefand interface scaling properties), the exposed surface area ~A(reflecting the surface roughness) and the channel sinuosity(reflecting how the channel network is organized on thetopography). One-dimensional profiles are determined byinterpolating from the irregular triangular mesh to a regularcellular mesh with an approximately equivalent resolution(i.e., �150 150). The nondimensional interface width isdefined as [e.g., Barabasi and Stanley, 1995]

~wð~l;~tÞ ¼ w=L ¼ 1

L

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffih½hð~x;~tÞ � hlð~x;~tÞ�2ix

q: ð9Þ

The subscript l means that a single window of data withlength l is selected in the direction parallel to the ridge axis

and used to measure the interface width and mean heighthl(~x;~t) in this window. The angle brackets h ix denote thataverages are taken over x. Many different windows alongthe same ridge-parallel profile are taken and the result isaveraged to obtain ~wð~l;~tÞ. The mean interface width ~w isobtaining by taking the mean of ~w measured in manydifferent ridge-parallel profiles. Note that although theinterface width (as formulated here) is a 1-D measure, itcontains information from the entire 2-D surface. Theinterface width, in addition to giving information concerningthe surface roughness and relief, is also used to measure theroughness exponent (describing how the interface widthscales with the window size), and to quantify the dominantchannel spacing, since the interface width saturates (does notincrease further) once the window size exceeds the channelspacing. The nondimensional exposed surface area isdefined as

~A ¼ A=Ao ¼XNj

j¼1

ffiffiffiffiffiffiffiffiffiffiffiffiffi1þ S2j

q; ð10Þ

where Sj is the average slope of an element j, Nj is the totalnumber of elements, A is the total surface area and Ao is theinitial total surface area. The average slope of an element iscalculated as the average of slopes determined at nodesusing the piecewise element shape function derivatives. Thenondimensional channel sinuosity (or average channellength) is defined here as

Se ¼1

Ni

XNi

i¼1

lsi

loi; ð11Þ

where lsi is the length of stream i, loi is the length of stream ifrom source to outlet in a straight line, and Ni is the totalnumber of streams sampled. The channel sinuosity wascalculated by specifying arbitrary channel sources in areasonably regular fashion over the entire grid and bytracing each ‘‘stream’’ to its respective outlet. Note that thismethod does not require actual definition of channels (andtheir sources) and can easily be carried out automatically bystream tracing down local gradients. In this respect, themeasure may be best thought of as a flow sinuosity ratherthan a channel sinuosity.

4.1. General Surface Evolution

[19] The first results to be presented demonstrate generalfeatures of how topography and runoff evolve through timefor relatively high De in response to constant climaticforcing in the form of uniform steady rainfall (Figure 2).Note that because base level is maintained at a constantlevel throughout all simulations (i.e., the surface is notundergoing uplift) the steady state solution is a flat surfacethat has been completely eroded.[20] The initial stage of topographic development is

governed by the presence and magnitude of initial topo-graphic perturbations (roughness) that have the ability toinhibit accumulation of runoff. Thus, even though De has alarge value indicating an overall dominance of runoff-assisted sediment transport (at the slope length scale L),the pinning of runoff at a scale much shorter than the slopelength scale effectively means that initial topographic evo-

Figure 1. Initial topographic surface and boundaryconditions used in computations. The surface has an initialslope of 1/10 unless stated otherwise. A random perturba-tion (i.e., roughness) is added to this surface to localizerunoff. The influence of the magnitude of initial roughnesson subsequent landscape evolution is investigated in section4.4 and Figure 7.

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 5

Page 8: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Figure 2a. Contour plots showing how topography evolves through time in response to constant baselevel (calculated for De = 103, n = 2 and l = 1; see equation (12)). The values of erosion shown aboveeach box represent the total erosion which has taken place since the experiment was initiated. Thus anerosion of 0.0 represents the initial conditions whereas a value of 1.0 indicates the time when all of theinitial topography has been removed. An erosion of 1.0 also coincides with the steady state solution.

ETG 7 - 6 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 9: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

lution is governed by linear diffusion (since ~q is small). Thisis particularly true when the initial surface is relatively flatand/or the magnitude of initial roughness is large. Theinitially formed channel network has an anastomosing form(e.g., before 0.02 erosion in Figure 2b) that migrateslaterally in an unstable fashion. However, with increasingtime, diffusion (both dispersive and concentrative) efficient-ly decays initial roughness due to its short-wavelengthnature allowing runoff to coalesce into larger streams andfor concentrative sediment transport to increase in impor-tance relative to dispersive transport. This leads to theformation of more stable channels (e.g., after 0.2 erosionin Figure 2b) which further concentrate runoff giving rise topositive feedback and enhanced channel incision.

[21] Several major changes in surface morphology takeplace in response to erosion and the development of a fullydeveloped channel network in the simulations investigated.First, the average channel spacing (quantified as the cross-over length scale in Figure 3a) first decreases and thenslowly increases in a quasi-linear fashion. Second, incisioncauses the overall surface roughness (or total exposedsurface area) to increase sharply before attaining a maxi-mum and decreasing in an exponential fashion (Figure 3b).Together, these observations imply a two-phase stage oftopographic evolution. Surface morphology during theinitial (growth) phase reflects the fact that new channelsare rapidly formed and propagate headward toward theupper boundary. In the later (decay) phase, topographic

Figure 2b. Same as Figure 2a except runoff discharge.

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 7

Page 10: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

evolution is dominated by lateral and vertical incisionwithin the already existing channel network. Interestingly,deposition does not play a major role in the topographicevolution. This is due largely to imposition of a fixed lowertopographic boundary condition.[22] The transition from a growth to a decay phase of

landscape evolution can be interpreted as a saturationphenomenon related to the progressive development ofspatial correlations. During the initial stage of erosion,different parts of the topographic surface act largely inde-pendently of one another and depend mainly on localproperties (local slope and hillslope diffusivity). This stage,which is dominated by linear diffusion, is characterized by ashort correlation length, the magnitude of which is con-trolled mainly by initial topographic perturbations. Aserosion proceeds, runoff of fluid over the surface andassociated fluvial sediment transport transmits informationabout the surface properties over a progressively larger area.Thus, even though erosion occurs locally, the increase incorrelation length enables lateral spread of informationthat is manifest in the surface as headward incision. Whenthe correlation length reaches the size of the system,corresponding to the arrival of headward incision at theupper boundary, the surface properties (e.g., the surfacearea) saturates. Subsequent evolution evolves by the pro-gressive decay of topography developed prior to saturationrather than by the formation of new features. Note thatalthough the initial stage of topographic development, asrepresented by the depinning of runoff trapped on thesurface by topographic perturbations and the transitionthrough anastomosing to a fully developed channel net-

work, is short-lived and involves a small amount of totalerosion, it represents an important period in the overalllandscape development because it involves the formation ofchannels, the location of which tend to remain relativelystable throughout subsequent evolution.

4.2. Influence of De on Landscape Morphology

[23] Simulations indicate that De, a measure of therelative importance of dispersive hillslope and concentrativefluvial transport processes, exerts an important influence onlandscape morphology. Large values of De (i.e., sedimenttransport is dominated by concentrative fluvial processes)favor development of rough, highly incised topography anda dense, branched channel network, whereas lowering thevalue of De, corresponding to increasing the contribution tosediment transport from dispersive hillslope processes,results in smoother topography and straighter, less branchedchannels that have a greater average spacing (Figure 4). Forthe maximum value of De investigated (i.e., 10,000), themaximum exposed surface area ~Amax (i.e., the surface area atsaturation) and the channel sinuosity (or equivalently theaverage channel length) are both enhanced by �30%relative to low values of De (Figures 5a and 5b). Theaverage channel spacing (as reflected by the crossoverlength scale ~lc) varies from a minimum value of �~lc =0.05 corresponding to �20 channels per unit length at De =10,000 to infinity at De � 1 (i.e., no channelizationobserved). Together, these results are consistent with thewidely accepted notion that increasing the importance ofdispersive sediment transport tends to suppress channelincision [e.g., Dunne and Aubry, 1986; Loewenherz-Law-

Figure 3. (a) Evolution of the exposed surface area ~A (surface roughness) and the crossover length scale~lc (average channel spacing) for the simulation shown in Figure 2 (De = 103, n = 2 and l = 1). Theexposed surface area is determined from equation (10). (b) Calculation of the crossover length scale. Forevery time of interest, corresponding to a fixed proportion of erosion, the scaling between the interfacewidth ~w and the spatial window size ~l is determined via equation (9) (three examples of such scalingrelations are shown in Figure 3b). The crossover length scale ~lc corresponds to the length scale at whichthe interface width saturates (or ceases to further increase). This value is obtained as the intersectionbetween two straight lines (shown in Figure 3b for erosion of 0.02). Also shown in Figure 3b is the locusof how the intersections (i.e., the value ~lc) vary through time. This locus of intersections is plotted inFigure 3a as the crossover length scale versus time. Note that the crossover length scale shown in Figure3b also corresponds to the average channel spacing, since the interface width determined in the directionparallel to the ridge axis saturates once the window size reaches the channel spacing.

ETG 7 - 8 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 11: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

rence, 1994] and to decrease the channel density [e.g.,Kirkby, 1986; Tarboton et al., 1992; Willgoose et al., 1992].[24] The relationship between the interface width ~w and

the window length ~l in the region below crossover is of the

approximate form w / lg where the roughness exponent g is�0.8 (Figure 5d). This roughness exponent is within therange of values measured for natural topography [see Doddsand Rothman, 2000]. Note that the magnitude of De exertsno influence on the spatial scaling of the interface width(i.e., on g) while it has an important control on the ultimatemagnitude of the width.

4.3. Influence of De on Landscape Dynamics

[25] The magnitude of De exerts an important influenceon landscape dynamics. For example, increasing De resultsin a decrease in the timescale within which a landscaperesponds to erosion, due to an increase in the bulk diffu-sivity of the system. The response to increasing De isrecognized in landscapes by an increase in the rate ofsurface roughening during the growth phase of topographicevolution associated with more rapid vertical and headwardincision and an increase in the rate of surface smoothingduring the decay phase (Figure 6a). The relationship be-tween the timescale required for saturation ~ts and De isobserved to have the approximate form ~ts / De

�0.78 (Figure6b). Thus increasing De by a factor of 10 decreases theresponse time of the system by a factor of �6. This resultdemonstrates the ability of fluvial processes to greatlyincrease rates of topographic evolution.[26] It is interesting to note that roughening of the surface

does not lead to power law scaling with time, as is observedfor most surface evolution processes [e.g., see Barabasi andStanley, 1995]. The time evolution of the surface morphol-ogy properties such as interface width ~w follows a logarith-mic-type law during the growth phase (i.e., the positive partof the curve in Figure 6c) of the approximate form ~w / mlog ~t (where m is the slope on a linear log plot). The reasonfor this difference may lie in the unstable nature of initialchannel incision that generates an exponential increase inroughness at small times. This result implies that the classicgrowth exponent is not well defined.

4.4. Effects Due to Variations in Initial SurfaceConditions

[27] Given that initial stages of surface evolution arestrongly influenced by the presence and magnitude of initialroughness (see section 4.1), it is anticipated that the initialsurface geometry may have an important influence on finitelandscape morphologies. Indeed, a considerable amount ofwork has already demonstrated that variability in the ran-domness applied as initial conditions can result in signifi-cant variations in the form and statistics of channelnetworks [e.g., Willgoose et al., 1991b; Howard, 1994].Attention here is focused on a different effect, namely thatdue to variations in the relative magnitude of initial noise.This effect has been tested in simulations carried out withdifferent values of the dimensionless parameter

l ¼ hr=H ; ð12Þ

where hr is the maximum value of initial random noise andH is the surface height at the ridge axis. Note that l! 1 asthe regional slope approaches zero. Results show that l hasa strong influence on the form of the channel network, withsmall values of l (corresponding to a steep overall slopeand/or a small magnitude of initial roughness) favoring

Figure 4. Influence of De on surface morphology (erosionof 0.2, n = 2, l = 1). Large values of De indicate sedimenttransport is dominated by fluvial processes, whereas for smallDe, hillslope sediment transport becomes more important.

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 9

Page 12: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

parallel drainage networks and an almost periodic channelspacing (e.g., Figure 7a), whereas large values of l favordevelopment of highly branched channel networks (e.g.,Figure 7d). These results are consistent with field observa-tions showing that dendritic drainage networks tend todevelop in regions with gentle slopes whereas paralleldrainage networks are characteristic of moderate or steepregional slopes [e.g., Howard, 1967]. The influence of lincreases sharply for values greater than �10, as indicatedby a dramatic (�50%) increase in channel sinuosity oraverage channel length as the relative magnitude ofroughness is increased further (Figure 8a).[28] It is interesting to note the apparent increase in the

scale of observation as l is increased (i.e., Figure 7aresembles small-scale rill structures on a hillslope, whereasFigure 7d resembles a large-scale drainage pattern). Thiseffect, although not due to a change in scale, may actuallyreflect one, since at the scale of a single hillslope (in naturalsystems) the ridge height tends to be large relative to themagnitude of roughness whereas at larger scales the region-al slope becomes progressively smaller relative to rough-

ness. As discussed more in section 5, this interpretation hasinteresting implications concerning the dependency of chan-nel morphology on scale.[29] Varying the relative magnitude of initial noise also

influences the interface width, with large magnitudes of lcausing large interface widths or relatively rough surfaces(Figure 8b). A more surprising result is that l significantlyinfluences the interface scaling. The roughness exponentsincrease from �0.7 to �0.8 as l is increased from 0.1 to100 (Figure 8b). Note that l does not influence thecrossover length scale (Figure 8b) or equivalently thecharacteristic drainage spacing (Figure 7).[30] The reason for the observed dependency of the

channel network form on l is likely to be due to theinfluence of surface roughness on local flow directions.When a surface is relatively rough with stochastic noise(as is a surface which is flat on average), runoff flowingdown the surface is strongly perturbed resulting in a highlybranching channel network, whereas when the magnitude ofroughness is relatively small (as is a steeply tilted surface),flow of runoff is relatively unhindered and the resulting

Figure 5. Influence of De on (a) channel sinuosity, (b) maximum exposed surface area,) (c) crossoverlength scale (average channel spacing, and (d) interface width scaling. Channel sinuosity, as defined here,is determined from equation (11). This figure indicates that the ratio of fluvial to hillslope sedimenttransport (i.e., De) has an important influence on surface and channel network properties.

ETG 7 - 10 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 13: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Figure 6. Influence of De on (a) the time evolution ofsurface area, (b) saturation time ~ts, and (c) interface width.The saturation time is defined as the time when the surfacearea saturates (as shown in Figure 6a for De = 100). Note thelinear relations between interface width and the logarithm oftime during growth and decay phases (Figure 6c).

Figure 7. Influence of the relative magnitude of initialroughness (l, see equation (12)) on surface morphologies(De = 1000, n = 2, initial surface slope of 1/100, erosion of0.2). Small values of l correspond to steep initial surfacesand/or small magnitude of initial roughness, whereas largevalues of l correspond to flat initial surfaces and/or largemagnitudes of initial roughness.

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 11

Page 14: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

channel network tends to be more regular and parallel innature. Note that in the extreme case where the magnitude ofinitial roughness is zero (and the initial surface is exactlyplanar), no channelization occurs and the solutions becometruly one-dimensional (i.e., neither the discharge or topog-raphy exhibit variation parallel to the ridge axis).

5. Discussion

[31] One of the major objectives of this study has been tosystematically investigate how the ratio of sediment trans-port on hillslopes to sediment transport in channels influ-ences the overall landscape morphology and dynamics. Tothis end, we have derived a critical nondimensional numberDe which can be interpreted as the ratio of characteristictimescales associated with dispersive and concentrativesediment transport processes, respectively. This number isa function of other parameters such as the characteristicsystem size (or slope length), hillslope diffusivity, fluvialdetachment coefficient, steady state rainfall and hydraulicregime. De is the only independent parameter (apart fromthose related to initial conditions) which can influence

landscape evolution for the set of equations investigated.Simulations demonstrate that De has an important influenceon both surface and channel network morphology and theresponse time of landscapes, with large values of De

(i.e., fluvial-dominated sediment transport) favoring roughhighly incised surfaces containing a dense (closely spaced)channel network that evolves rapidly in time. For lowervalues of De, time evolution is slower, the resulting land-scapes are smoother and the channel network is lessbranching and less dense. Note that high steady state rainfall(large De) corresponds to fluvial-dominated model land-scapes (Figure 4), whereas in nature it is typical thatrelatively dry climates with highly sporadic rainfall arefluvial-dominated. Thus equating ‘‘strongly fluvial’’ and‘‘wet’’ is not possible. This point emphasizes the impor-tance of considering transient effects of climate whenestimating values of De from natural landscapes. Theimportance of a transient climate in influencing erosionrates has been argued by Molnar [2001] on the basis ofgeneral transport laws and the frequency-magnitude distri-bution of floods and has been investigated numerically byTucker and Bras [2000].[32] Our results imply that some important properties of

landscapes such as local relief, drainage density and exposedsurface area are strongly influenced by the dominant processof sediment transport. In particular, fluvial-dominated sys-tems favor generation of large relief (i.e., large interfacewidths) and large total exposed surface area. We emphasizethat even though it is river channels which are capable ofcreating large relief, hillslope sediment transport imposes animportant limit to the total relief through modifying thehillslope length and channel density [see also Montgomeryand Dietrich, 1992]. Thus the finite landscape morphologyreflects the relative importance of sediment transport pro-cesses: competition between the roughening influence ofconcentrative fluvial transport and the smoothing influenceof hillslope transport. Interestingly, the increase in surfacearea due to fluvial incision may have a self-regulatinginfluence on landscape morphology by increasing the pro-portion of substrate available for (weathering and) hillslopetransport. Thus the effective magnitude of De may bebuffered by the landscape. This effect may explain why thechannel spacing (and to a lesser extent the exposed surfacearea) saturates with increasing De beyond �103 (see Figures5b and 5c).[33] The predictable way in which landscape morphology

responds to De is broadly consistent with other studies. Forexample, field-based studies indicate that hillslopes tend todominate landscapes at small spatial scales (small De),whereas fluvial features tend to dominate at large scales(large De) (e.g., Montgomery and Dietrich, 1992). Further-more, whereas high drainage densities (and small channelspacings) are often found in association with semiaridbadland-type (fluvial dominated) landscapes, low drainagedensities (large channel spacings) are typical of more humid(hillslope dominated) landscapes [e.g., Kirkby, 1992]. Sim-ilarly, many studies have shown that catchment propertiessuch as the areas-slope relationship are strongly influencedby the ratio between the importance of hillslope and channelprocesses, the transition (in process dominance) beingcontrolled by change of scale [e.g., Willgoose, 1994], rocktype [e.g., Howard, 1994], climate [e.g., Tucker and Sling-

Figure 8. Influence of l (the relative magnitude of initialroughness) on (a) channel sinuosity Se (see equation (11))and (b) interface width scaling for two of the simulationsshown in Figure 7.

ETG 7 - 12 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 15: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

erland, 1997], or the presence of different dominant hill-slope processes [e.g., Tucker and Bras, 1998]. A majorcontribution of this study has been to put results such asthese into a consistent framework by identifying a singleindependent controlling parameter (i.e., De, which dependson all other model parameters) and by quantifying itsinfluence.[34] The strong dependency of average channel spacing

on De demonstrated here is inconsistent with field datapresented by Hovius [1996], at least at first glance. Hovius[1996] showed that major rivers draining relatively linearmountain belts from around the world display a remarkablyregular spacing in a wide range of climatic settings andsubstrate lithologies. The same data also show a consistentlinear relationship between drainage spacing and the halfwidth of the mountain belt, implying a strong kinematic orgeometrical control on the spacing of major rivers (in well-developed drainage systems). In contrast, simulations indi-cate that the average channel spacing tends to decrease withincreasing De, and therefore with increasing system size.However, the connection between this study and that ofHovius [1996] is not straightforward since attention here isfocused on the average channel spacing, not the spacing ofmajor drainage outlets as investigated by Hovius [1996].The two measures yield significantly different results asillustrated in Figure 2b (erosion is 0.6), where it can be seenthat there are 2–3 major drainage outlets (with a spacing of�0.5), but where the calculated average channel spacing is0.084 (implying 1/0.084�12 channels). In this case, theempirical relation of Hovius [1996] predicts a spacing of0.5, as observed for major channel outlets (given that thewidth of the slope is 1). An additional complication isimplied by the observation that the channel structure andspacing changes in time (e.g., in Figure 2b compare erosionis 0.02 where there is a single channel spacing frequencywith erosion of 0.6 which shows an additional channelfrequency at larger scale comprising the ‘‘major’’ drainageoutlets). Together these observations highlight the difficultyof resolving the channel spacing problem as currently posedand the importance of devising measures (e.g., analyzingspectra or interface width crossover distances) that can beapplied to large data sets in an objective fashion.[35] It is interesting to note that the characteristic time-

scale for linear hillslope processes based on the relation th =L2/k varies between �1000 and 10,000 Ma for a hillslopediffusivity 103 m2 Ma�1 [e.g., Hanks et al., 1984] andsystems sizes varying between 1 and 10 km, respectively.The enormity of this timescale compared to the time withinwhich landscapes are thought to respond may tempt one toargue that linear hillslope diffusion of regolith or othercohesionless material makes up a negligible contribution tonatural landscapes and can, therefore, be ignored. However,as our simulations demonstrate, both the landscape mor-phology and dynamics are sensitive to variations in De, evenfor large values of De corresponding to small hillslopediffusivities. The inclusion of fluvial sediment transportdrastically decreases the timescale of landscape evolution.For example, increasing De by a factor of 10 decreases thesystem response time by a factor of 6 assuming n = 2 (seesection 4.3). This result implies that in order for landscapesat spatial scales of 1 to 10 km to evolve on timescales of1 Ma, De must be on the order of 1000–10,000 (based on

the hillslope timescale discussed above). A De of thismagnitude appears at least plausible based on the visualappearance of model landscapes (see Figure 4). Note thatthe fluvial transport exponent n will strongly influence therate of topographic evolution, with fast response timescorresponding to large values of n. The surface morphologyand channel spacing are less sensitive to varying n, as longas n > 1.[36] Treating landscapes as a transport-limited system

involving coupling between hillslope and fluvial processesleads naturally to the clear identification of an effective(system) diffusivity ke = k + c qn (see equation (4)), or innon dimensional form ~ke = 1 + De~q

n (see equation (6)). Asnoted above, because ke increases with system size (due todependency of ke on q), the time scale within which finitelandscapes develop is considerably smaller than the casefor systems evolving by pure linear diffusion. In reality, keis a function of the frequency and magnitude of rainfall(only the latter of which has been treated here). Estimationof the effective diffusivity (and therefore of the systemsediment transport function) in natural systems couldpotentially be obtained by measuring (at a variety of lengthscales) system response times related to either climatic ortectonic perturbations.[37] Results of simulations demonstrate that landscapes

undergo important changes in surface and channel networkmorphology as erosion proceeds with time. For example, itwas shown that following the emergence of highly dynamicshallow incisions on an initially unchannelled surface, thechannel network stabilizes and develops by a combinationof headward incision and lateral channel widening. Thesefeatures appear to be reasonably well understood and havebeen observed in many numerical studies [e.g., see Will-goose et al., 1991c; Rigon et al., 1994; Tucker and Sling-erland, 1997]. The general coarsening of natural drainagenetworks with time does not appear to have been widelydiscussed in the literature, although it has been observed[e.g., Collins et al., 1983; Smith et al., 1997b]. Thisbehavior is interpreted to indicate a transition between anearly (linear) development stage where incisive instabilitiesare fast growing and of very short wavelength and laterdevelopment (nonlinear) stages where slower growing,longer wavelength instabilities begin to dominate dueto saturation of earlier instabilities, as theoretical resultssuggest [e.g., Loewenherz-Lawrence, 1994]. If this interpre-tation is correct, one may expect in nature that short-wavelength channel features to have small amplitudes(because they saturate early) and long-wavelength featuresto have relatively large amplitudes. This inference is con-sistent with studies showing that natural landscapes tend tobe self-similar [e.g., Huang and Turcotte, 1989] and withthe observed power law scaling between ~w and ~l (e.g.,Figure 5d). Note, however, that short-wavelength channelsare anticipated to be difficult to observe in humid climateswith abundant soil cover because response times are sorapid at small scales that channels formed in early devel-opment stages are likely to be rapidly obliterated duringperiods of prolonged precipitation. The exception to thismay occur in relatively arid climates during erosion inresponse to short and intense rain bursts. Early stages oftopographic development bearing very short wavelengthincisions have more chance of being preserved due to the

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 13

Page 16: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

sporadic nature of precipitation events that are separated bylong periods within with very little erosion occurs becausean almost complete lack of hillslope diffusion.[38] A characteristic feature demonstrated by the simu-

lations presented is the presence of a two-stage history forlandscape evolution (e.g., see Figure 3). Surface morphol-ogy during the initial (growth) phase reflects the fact thatnew channels are rapidly formed and propagate headwardtoward the upper boundary resulting in growth of relief andsurface roughening. In the later (decay) phase, topographicevolution is dominated by lateral and vertical incisionwithin the already existing channel network that leads toexponential decay of topographic relief and smoothing oftopography toward a flat (steady state) surface. Similarphases of topographic evolution were envisaged by Davis[1954] and have commonly been observed in numericalmodels studying the response to impulse-like tectonicforcing [e.g., Willgoose, 1994; Kooi and Beaumont, 1996;Smith et al., 1997b], although these studies distinguishedthree rather than two main phases. We suggest that thetransition from growth to decay phases of topographicevolution is explained by the growth of spatial correlationson the surface, which eventually reach the system size atwhich time the entire surface becomes correlated and thesurface properties (such as interface width) saturate (seesection 4.1 for more detail). This interpretation is consistentwith the observed dependency of the time required forsaturation (i.e., ~ts) on De and with the general observationthat saturation coincides with the arrival of backwardincision at the upper boundary (e.g., erosion of 0.2 inFigure 2). Kooi and Beaumont [1996] suggest that thetransition from growing to declining sediment yield indi-cates the time when rivers achieve grade at the base level forthe first time, at which time a local dynamic equilibrium isachieved. However, these authors also point out that thisdynamic equilibrium cannot correspond to overall macro-scopic equilibrium since the later coincides with steady statewhen all topography is eroded to base level (i.e., assumingzero base level change following initial block uplift). Aninteresting question that needs to be addressed in futurework is: Does the saturation time related to channelsconnecting with the entire surface correspond with the timeat which steady state is achieved in situations experiencingconstant base level fall?[39] Simulations demonstrate that some important fea-

tures of landscapes form very early in the erosional history,reflecting an important dependence on initial conditions.This result is illustrated by the formation of channels duringthe initial growth phase of topographic evolution, thelocation of which remain relatively constant throughoutsubsequent evolution (Figure 2b) and by the strong influ-ence that the magnitude of initial roughness (relative to theoverall surface slope) has on finite surface and channelnetwork morphologies (Figure 7). Similar interpretationshave been made on the basis of field data. For example, it iswell known that dendritic drainage networks tend to devel-op in regions with gentle initial slopes whereas paralleldrainage networks are characteristic of moderate or steepregional slopes [e.g., Howard, 1967]. Furthermore, Tallinget al. [1997] argue that the drainage spacing may bedetermined primarily during the early stages of channelnetwork growth. We suggest that the dependency of surface

and channel network morphologies on initial noise is relatedto competition between the local roughening effect of noiseand global smoothing effect of diffusive sediment transport.In this sense, the initial topographic noise acts as a form ofquenched noise except that it is progressively decayed intime by diffusion. If the noise is initially large, then it cansignificantly influence the flow directions and the channelnetwork before time has had a chance to eliminate itsinfluence. If, on the other hand, the initial noise is relativelysmall, the surface and channel network morphology isdetermined largely by competition between dispersive andconcentrative sediment transport. Simulations indicate thatthe effect of increasing the magnitude of initial roughnessappears similar to the effect that could be anticipated byincreasing the scale of observation (see section 4.4). Wesuggest that this effect, although not due to a change inscale, may actually reflect one, since casual observation ofnatural examples indicates that, at the scale of a singlehillslope, the ridge height tends to be large relative to themagnitude of roughness (a condition tending to favor theformation of rill-like channels) whereas at larger scales,regionally averaged slopes become progressively smallerrelative to roughness (a condition leading to the formationof more complex dendritic drainage networks). It is sug-gested that structures such as rills are not necessarilyrestricted to small scale (i.e., because of some specificprocess dominance at this scale) but that the initial surfaceconditions favorable for their formation tend to be typical atsmall scale. Similar arguments apply to the formation oflarge-scale dendritic drainage basins. Thus we argue for acontinuous transition between, and a single formation modefor small-scale structures such as rills and large-scaledrainage basins.[40] This paper has also quantified how interface width

scales as a function of space and time. Roughness exponentsderived from simulations, which describe how the interfacewidth changes with changing spatial scale, vary between 0.7and 0.8 at small scales. These exponents progressivelydecrease as scale increases, before eventually droppingclose to zero at the scale defined by the main channelspacing (termed here the crossover length scale). Similarexponents and crossover behavior are reported for naturaldata [e.g., Dodds and Rothman, 2000]. Model roughnessexponents at the smallest scale are shown to be independentof De, which is a surprising result given the strong influenceDe imposes on the overall surface properties. Nevertheless,this result is consistent with the work of Rigon et al. [1994],who showed the scaling of planar properties (such asinterface width) does not appear to be significantly influ-enced by hillslope diffusion, whereas the scaling of surfaceproperties is modified. The roughness exponents are, how-ever, dependent on the relative magnitude of initial noise. Asimilar dependency of roughness exponents on noise hasbeen observed for many other physical systems [Barabasiand Stanley, 1995]. Willgoose et al. [1991b] also showedthat network statistics (such as the length and bifurcationratios) are sensitive to small variations of initial randomness(while holding the magnitude of noise constant). Notehowever, that such variations lead to networks that are‘‘physically similar’’ [see also Howard, 1994], which isnot the case when varying the magnitude of initial noise (seeFigure 7). A surprising result found here is the logarithmic

ETG 7 - 14 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY

Page 17: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

scaling observed between interface width and time. Physi-cally this law implies that the interface width initially growsvery rapidly while later it grows very slowly. As suggestedin section 4.3, the reason for this behavior may lie in theinitially unstable nature of channel incision [see Smith andBretherton, 1972; Loewenherz-Lawrence, 1994] that gener-ates a rapid increase in roughness at small times. It is muchmore normal that surface growth in other physical systemsis controlled by power laws [Barabasi and Stanley, 1995].That fact that growth of surfaces by erosion is not controlledby a power law indicates that the growth exponent is notdefined.

6. Conclusions

[41] This paper presented a detailed investigation of howthe ratio of sediment transport on hillslopes to sedimenttransport in channels influences landscape and channelnetwork morphology and dynamics. This problem has beeninvestigated by numerically solving and investigating sol-utions of equations that are based on the conservation ofsurface water and sediment combined with a simple trans-port law connecting the sediment flux to the local slope andrunoff discharge. The most important points of this studyare summarized as follows:[42] 1. We have identified a critical nondimensional

parameter De that is a function of rainfall, system size, rocktype and hydraulic regime and that is a measure of therelative importance of fluvial and hillslope sediment trans-port. Large values of De imply that sediment transport isdominantly fluvial. De is the only independent parameter(apart from those related to initial conditions) which caninfluence landscape evolution for the set of equationsinvestigated.[43] 2. De has an important and predictable influence on

the surface morphology (e.g., total exposed surface area,surface roughness, local relief), channel network form (e.g.,channel sinuosity), and channel spacing. Fluvial dominatedsystems (i.e., large De) are characterized by rough, high-relief, highly incised surfaces which contain a dense (close-ly spaced) channel network. Lowering De leads to smoothertopography, lower-relief, greater channel spacing and lesssinuous channels.[44] 3. De has an important influence on the dynamics of

erosion, with large values of De greatly increasing the rate atwhich landscapes evolve.[45] 4. Topographic evolution in the absence of base level

change tends to occur in two major phases, an initial growthphase reflecting formation of new channel and creation ofrelief and a later decay phase related to destruction ofpreexisting channels. The transition between growth anddecay phases is suggested to be related to a saturationphenomenon.[46] 5. The scaling behavior of simulated topography

with respect to both time and space is obtained and isshown to be independent of De. Roughness exponents arefound to be in the range 0.7–0.8. The magnitude of initialroughness influences the roughness exponents. The inter-face width is shown to grow and decay as the logarithm oftime. Thus the classical growth exponent is undefined.[47] 6. Surface and channel network morphologies are

shown to be strongly influenced by the overall surface slope

relative to the magnitude of initial topographic roughness(with relatively flat and/or rough initial surfaces favoringdevelopment of highly branching channel networks, where-as relatively steep initial surfaces favor parallel drainagepatterns with regular drainage spacings).[48] Finally, this study indirectly shows that classical

well-established numerical techniques such as the finiteelement method can be used with success to investigatethe coupled runoff-erosion problem.

[49] Acknowledgments. Yuri Podladchikov, Alex Densmore, andMike Ellis are thanked for providing encouragement and for many stimu-lating discussions. Reviews and comments by Clem Chase, Kelin Whipple,and two anonymous reviewers are greatly appreciated. This work wassupported by the Swiss National Science Foundation (grant 620-57863).

ReferencesAhnert, F., Process-response models of denudation at different spatialscales, Catena Suppl., 10, 31–50, 1987.

Andrews, D. J., and R. C. Bucknam, Fitting degradation of shoreline scarpsby a nonlinear diffusion model, J. Geophys. Res., 92, 12,857–12,867,1987.

Barabasi, A.-L., and H. E. Stanley, Fractal Concepts in Surface Growth,366 pp., Cambridge Univ. Press, New York, 1995.

Beaumont, C., P. Fullsack, and J. Hamilton, Erosional control of activecompressional orogens, in Thrust Tectonics, edited by K. R. McClay,pp. 1–18, Chapman and Hall, New York, 1992.

Braun, J., and M. Sambridge, Modeling landscape evolution on geologicaltime scales: A new method based on irregular spatial discretisation, BasinRes., 9, 27–52, 1997.

Carson, M. A., and M. J. Kirkby, Hillslope Form and Process, 475 pp.,Cambridge Univ. Press, New York, 1972.

Chase, C. G., Fluvial landsculpting and the fractal dimension of topogra-phy, Geomorphology, 5, 39–57, 1992.

Collins, B. D., T. Dunne, and A. K. Lehre, Erosion of tephra-coveredhillslopes north of Mt St. Helens, Washington: May 1980–May 1981,Z. Geomorphol., Suppl., 46, 103–121, 1983.

Cordova, J. R., I. Rodriguez-Iturbe, and P. Vaca, On the development ofdrainage networks, in Recent Developments in the Explanation and Pre-diction of Erosion and Sediment Yield, edited by D. E. Walling, IAHSPubl., 137, 239–249, 1982.

Culling, W. E. H., Analytical theory of erosion, J. Geol., 68, 336–344,1960.

Culling, W. E. H., Theory of erosion on soil-covered slopes, J. Geol., 73,230–254, 1965.

Davis, W. M., Geographical Essays, 777 pp., Dover, Mineola, N.Y., 1954.Dodds, P. S., and D. H. Rothman, Scaling, universality and geomorphology,Annu. Rev. Earth Planet. Sci., 28, 571–610, 2000.

Dunne, T., and B. A. Aubry, Evaluation of Horton’s theory of sheetwashand rill erosion on the basis of field experiments, in Hillslope Processes,Binghamton Symp. Geomorphol. Int. Ser., vol. 16, edited by A. D. Abra-hams, pp. 31–53, Allen and Unwin, Concord, Mass., 1986.

Graf, W. H., Hydraulics of Sediment Transport, 513 pp., McGraw-Hill,New York, 1971.

Hanks, T. C., R. C. Bucknam, K. R. Lajoie, and R. F. Wallace, Modificationof wave-cut and faulting-controlled landforms, J. Geophys. Res., 89,5771–5790, 1984.

Hasbargen, L. E., and C. Paola, Landscape instability in an experimentaldrainage basin, Geology, 28, 1067–1070, 2000.

Horton, R. E., Erosional development of streams and their drainage basins:Hydrophysical approach to quantitative morphology, Geol. Soc. Am.Bull., 56, 275–370, 1945.

Hovius, N., Regular spacing of drainage outlets from linear mountain belts,Basin Res., 8, 29–44, 1996.

Howard, A. D., Drainage anaysis in geologic interpretation: A summary,Am. Assoc. Pet. Geol. Bull., 51, 2246–2259, 1967.

Howard, A. D., A detachment-limited model of drainage basin evolution,Water Resour. Res., 30, 2261–2285, 1994.

Huang, J., and D. L. Turcotte, Fractal mapping of digitized images: Appli-cations to the topography of Arizona and comparison with syntheticimages, J. Geophys. Res., 94, 7491–7495, 1989.

Izumi, N., and G. Parker, Inception of channelization and drainage basinformation: Upstream-driven theory, J. Fluid Mech., 283, 341–363,1995.

Izumi, N., and G. Parker, Linear stability analysis of channel inception:Downstream-driven theory, J. Fluid Mech., 419, 239–262, 2000.

SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY ETG 7 - 15

Page 18: Project: Modelling of erosion and landscape formation · Project: Modelling of erosion and landscape formation Anyone who has visited a mountain-belt such as the European alps realizes

Kirkby, M. J., A two-dimensional simulation model for slope and streamevolution, in Hillslope Processes, Binghamton Symp. Geomorphol. Int.Ser., vol. 16, edited by A. D. Abrahams, pp. 203–222, Allen and Unwin,Concord, Mass., 1986.

Kirkby, M. J., Thresholds and instability in stream head hollows: A modelof magnitude and frequency for wash processes, in Process Models andTheoretical Geomorphology, edited by M. J. Kirkby, pp. 295–314, JohnWiley, New York, 1994.

Kooi, H., and C. Beaumont, Escarpment evolution on high-elevation riftedmargins: Insights derived from a surface evolution model that combinesdiffusion, advection and reaction, J. Geophys. Res., 99, 12,191–12,209,1994.

Kooi, H., and C. Beaumont, Large-scale geomorphology: classical conceptsreconciled and integrated with contemporary ideas via a surface processesmodel, J. Geophys. Res., 101, 3361–3386, 1996.

Loewenherz-Lawrence, D. S., Hydrodynamic description for advectivesediment transport processes and rill initiation, Water Resour. Res.,30, 3203–3212, 1994.

Molnar, P., Climate change, flooding in arid environments, and erosionrates, Geology, 29, 1071–1074, 2001.

Montgomery, D. R., and W. E. Dietrich, Channel initiation and the problemof landscape scale, Science, 255, 826–830, 1992.

Rigon, R., A. Rinaldo, and I. Rodriguez-Iturbe, On landscape evolution,J. Geophys. Res., 99, 11,971–11,993, 1994.

Rinaldo, A., W. E. Dietrich, R. Rigon, G. Vogel, and I. Rodriguez-Iturbe,Geomorphological signatures of varying climate, Nature, 374, 632–634,1995.

Roering, J. J., J. W. Kirchner, and W. E. Dietrich, Hillslope evolution bynonlinear, slope-dependent transport: Steady state morphology and equili-brium adjustment timescales, J. Geophys. Res., 106, 16,499–16,513, 2001.

Schumm, S. A., Evolution of drainage systems and slopes in badlands atPerth Amboy, New Jersey, Geol. Soc. Am. Bull., 67, 597–646, 1956.

Smith, T. R., and F. P. Bretherton, Stability and the conservation of mass indrainage basin evolution, Water Resour. Res., 8, 1506–1529, 1972.

Smith, T. R., B. Birnir, and G. E. Merchant, Towards an elementary theoryof drainage basin evolution: 1, The theoretical basis, Comput. Geosci.,23, 811–822, 1997a.

Smith, T. R., G. E. Merchant, and B. Birnir, Towards an elementary theoryof drainage basin evolution: 1, Computational evaluation, Comput. Geos-ci., 23, 823–849, 1997b.

Talling, P. J., M. D. Stewart, C. P. Stark, S. Gupta, and S. J. Vincent,Regular spacing of drainage outlets from linear fault blocks, BasinRes., 9, 275–302, 1997.

Tarboton, D. G., R. L. Bras, and I. Rodriguez-Iturbe, A physical basis fordrainage density, Geomorphology, 5, 59–76, 1992.

Tucker, G. E., and R. L. Bras, Hillslope processes, drainage density, andlandscape morphology, Water Resour. Res., 34, 2751–2764, 1998.

Tucker, G. E., and R. L. Bras, A stochastic approach to modeling the role ofrainfall variability in drainage basin evolution, Water Resour. Res., 36,1953–1964, 2000.

Tucker, G. E., and R. Slingerland, Erosional dynamics, flexural isostacy,and long-lived escarpments: A numerical modeling study, J. Geophys.Res., 99, 12,229–12,243, 1994.

Tucker, G. E., and R. Slingerland, Drainage basin response to climatechange, Water Resour. Res., 33, 2031–2047, 1997.

Tucker, G. E., and K. X. Whipple, Topographic outcomes predicted bystream erosion models: Sensitivity analysis and intermodel comparison,J. Geophys. Res., 107, 2179, doi:10.1029/2001JB000162, 2002.

Willett, S. D., R. Slingerland, and N. Hovius, Uplift, Shortening and SteadyState Topography in Active Mountain Belts, Am. J. Sci., 301, 455–485,2001.

Willgoose, G., A physical explanation for an obsered area-slope-elevationrelationship for catchments with declining relief, Water Resour. Res., 30,151–159, 1994.

Willgoose, G. R., R. L. Bras, and I. Rodriguez-Iturbe, A coupled channelnetwork and hillslope evolution model: 1. Theory,Water Resour. Res., 27,1671–1684, 1991a.

Willgoose, G. R., R. L. Bras, and I. Rodriguez-Iturbe, A coupled channelnetwork and hillslope evolution model: 2. Nondimensionalization andapplications, Water Resour. Res., 27, 1685–1696, 1991b.

Willgoose, G. R., R. L. Bras, and I. Rodriguez-Iturbe, Results from a newmodel of river basin evolution, Earth Surf. Processes Landforms, 16,237–254, 1991c.

Willgoose, G. R., R. L. Bras, and I. Rodriguez-Iturbe, A physical explana-tion of an observed link area-slope relationship, Water Resour. Res., 27,1697–1702, 1991d.

Willgoose, G. R., R. L. Bras, and I. Rodriguez-Iturbe, The relationshipbetween catchment and hillslope properties: Implications of a catchmentevolution model, Geomorphology, 5, 21–37, 1992.

Zienkiewicz, O. C., and R. L. Taylor, The Finite Element Method, vol. 1,The Basis, Butterworth-Heinemann, Woburn, Mass., 2000a.

Zienkiewicz, O. C., and R. L. Taylor, The Finite Element Method, vol. 3,Fluid dynamics, Butterworth-Heinemann, Woburn, Mass., 2000b.

�����������������������F. Schlunegger, Institute of Geological Sciences, University of Bern,

Baltzerstr 1, Bern, CH-3012 Switzerland. ([email protected])G. Simpson, Institut fur Geologie, Sonneggstr. 5, ETH Zentrum, CH-

8092 Zurich, Switzerland. ([email protected])

ETG 7 - 16 SIMPSON AND SCHLUNEGGER: LANDSCAPE EVOLUTION AND MORPHOLOGY