17
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 104, NO. B4, PAGES 7359-7375, APRIL 10, 1999 Self-organization and scaling relationships of evolving river networks Jon D. Pelletlet Division of Geological and PlanetaryScience, California Institute of Technology, Pasadena Abstract. The powerspectra$ of linear transects of Earth's topography is often observed to be a power law function of wave number k with exponent closeto -2: S(k) oc k -2. In addition, rivernetworks are fractal trees that satisfy several power law relationships between their morphologic components. A modelequation for the evolution of Earth's topographyby transport-limited erosional processes whichproduces fractal tepography and fractal river networks is presented, and its solutions are compared in detail to real topography.The model is the diffusion equation for sediment transporton hillslopes and channels with the diffusivity constanton hillslopes and proportionalto the three-halves power of discharge in channels.The dependence of diffusivityon discharge is consistent with sediment ratingcurves. We study the model in twoways. In the firstanalysis the diffusivity is parameterized asa function of elevation, anda Taylor expansion procedure is carried out to obtain a differential equation for the landform elevation which includes the spatiallyvariablediffusivity to first order in the elevation. The solution to this equationis a self-affine or fractal surface with linear transects that have power spectra $(k) oc k -•'s, independent of the age of the topography, consistent with observations of real topography. The hypsometry produced by the model equation is skewed such that lowlands makeup a largerfractionof the total area than highlands as observed in real topography. In the second analysis we include river networks explicitly in a numericalsimulation by calculating the discharge at every point. We characterize the morphology of real river basins with five independent scaling relations between six morphometric variables.Scaling exponents are calculated for seven river networksfrom a variety of tectonic environments using high-quality digitalelevation models. River networks formed in the model match the observed scaling laws and satisfy Tokunagaside-branching statistics. 1. Introduction A remarkable feature of Earth's surface is its scale invariance.Objectssuch as a hammeror a person of- ten need to be included in landscapephotographs be- cause features such as variations in height of a topo- graphic profile have no characteristic scale. Similarly, a scale on a map is oftennecessary to determine whether the map detailsfeatures at the scale of I km or hun- dredsof kilometers. The scale invariance of topography can be quantified with techniques of time series anal- ysis. Variations in the height of a topographic profile can be characterized with the probability density func- tion and the power spectraldensity or power spectrum. The power spectrum is proportional to the square of the coefficients in a Fourier seriesrepresentation of the Copyright 1999by the American Geophysical Union. Paper number 1998JB900110. 0148-0227/99/1998JB900110509.00 transect. The powerspectra S of linear transects of to- pographyhave a power law dependence on wave num- ber with an exponentclose to -2 over a wide range of scales: S(k) oc k-2 [Vening Meinesz, 1951; Mandelbrot, 1975;Sayles and Thomas, 1978;Newmanand Turcotte, 1990]. Culling and Datko[1987] have obtained equiv- alent results with the rescaled-range technique. Mat- sushita and Ouchi [1989] have computed the Hausdorff measureHa defined by the relationshipbetween the standard deviation and the length L of the transect, a c• LHa/2, for several topographic transects. They obtained Ha • 0.55 which implies a power spectral ex- ponent of fl = -2.1 from the relation fl = 2Ha + 1 [e.g.,Turcotte, 1992].Ahnert [19s4] obtained a similar value. Turcotte [1987] and Balmino [1993] have com- puted the power spectrumfrom a spherical harmonic representation of Earth's topography and bathymetry. They observed a power spectrum S(k) crk -2 at scales i 10,000 km and an approximately constantspectrum at largerscales. Similar scale invariance has beenidenti- fiedin Earth'sbathymetry [Bell, 1975], the topography 7359

Self-organization and scaling relationships

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Self-organization and scaling relationships

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 104, NO. B4, PAGES 7359-7375, APRIL 10, 1999

Self-organization and scaling relationships of evolving river networks

Jon D. Pelletlet

Division of Geological and Planetary Science, California Institute of Technology, Pasadena

Abstract. The power spectra $ of linear transects of Earth's topography is often observed to be a power law function of wave number k with exponent close to -2: S(k) oc k -2. In addition, river networks are fractal trees that satisfy several power law relationships between their morphologic components. A model equation for the evolution of Earth's topography by transport-limited erosional processes which produces fractal tepography and fractal river networks is presented, and its solutions are compared in detail to real topography. The model is the diffusion equation for sediment transport on hillslopes and channels with the diffusivity constant on hillslopes and proportional to the three-halves power of discharge in channels. The dependence of diffusivity on discharge is consistent with sediment rating curves. We study the model in two ways. In the first analysis the diffusivity is parameterized as a function of elevation, and a Taylor expansion procedure is carried out to obtain a differential equation for the landform elevation which includes the spatially variable diffusivity to first order in the elevation. The solution to this equation is a self-affine or fractal surface with linear transects that have power spectra $(k) oc k -•'s, independent of the age of the topography, consistent with observations of real topography. The hypsometry produced by the model equation is skewed such that lowlands make up a larger fraction of the total area than highlands as observed in real topography. In the second analysis we include river networks explicitly in a numerical simulation by calculating the discharge at every point. We characterize the morphology of real river basins with five independent scaling relations between six morphometric variables. Scaling exponents are calculated for seven river networks from a variety of tectonic environments using high-quality digital elevation models. River networks formed in the model match the observed scaling laws and satisfy Tokunaga side-branching statistics.

1. Introduction

A remarkable feature of Earth's surface is its scale

invariance. Objects such as a hammer or a person of- ten need to be included in landscape photographs be- cause features such as variations in height of a topo- graphic profile have no characteristic scale. Similarly, a scale on a map is often necessary to determine whether the map details features at the scale of I km or hun- dreds of kilometers. The scale invariance of topography can be quantified with techniques of time series anal- ysis. Variations in the height of a topographic profile can be characterized with the probability density func- tion and the power spectral density or power spectrum. The power spectrum is proportional to the square of the coefficients in a Fourier series representation of the

Copyright 1999 by the American Geophysical Union.

Paper number 1998JB900110. 0148-0227/99/1998JB900110509.00

transect. The power spectra S of linear transects of to- pography have a power law dependence on wave num- ber with an exponent close to -2 over a wide range of scales: S(k) oc k -2 [Vening Meinesz, 1951; Mandelbrot, 1975; Sayles and Thomas, 1978; Newman and Turcotte, 1990]. Culling and Datko [1987] have obtained equiv- alent results with the rescaled-range technique. Mat- sushita and Ouchi [1989] have computed the Hausdorff measure Ha defined by the relationship between the standard deviation and the length L of the transect, a c• L Ha/2, for several topographic transects. They obtained Ha • 0.55 which implies a power spectral ex- ponent of fl = -2.1 from the relation fl = 2Ha + 1 [e.g., Turcotte, 1992]. Ahnert [19s4] obtained a similar value. Turcotte [1987] and Balmino [1993] have com- puted the power spectrum from a spherical harmonic representation of Earth's topography and bathymetry. They observed a power spectrum S(k) cr k -2 at scales i 10,000 km and an approximately constant spectrum at larger scales. Similar scale invariance has been identi- fied in Earth's bathymetry [Bell, 1975], the topography

7359

Page 2: Self-organization and scaling relationships

7360 PELLETIER: FRACTAL RIVER NETWORKS

of natural rock surfaces [Brown and $cholz, 1985], and the topography of Venus [Kucinskas et al., 1992]. The observation of scale-invariant topography on Venus in- dicates that fractal topography can be formed without erosion. A power law power spectrum is indicative of scale invariance since the power law function has no length scale. Turcotte [1992] has shown that these ob- servations define topography to be a self-affine fractal with a fractal dimension close to 2.5. Synthetic topog- raphy which assumes random phases in the Fourier co- efficients can be generated which has S(k) oc k -2 as observed in real topography. Images of synthetic to- pography with $(k) oc k -2 produced with techniques described by Voss [1988] and presented by Mandelbrot [1983] resemble natural topography. However, the lack of any river networks in these images indicates that the power spectrum is not a complete representation of Earth's topography [Weissel et al., 1994]. Gilbert [1989] and Evans and McClean [1995] have documented deviations from scale invariance in Earth's topography.

Besides the power spectrum, the distribution or hyp- sometry of topography is an important statistical mea- sure. The topography of Earth's continents is highly skewed such that a much larger percentage of Earth's topography is lowlands (topography with an elevation below the median elevation for a region) and there is a positive correlation between elevation and slope (i.e., as one drives up a mountain, the steepness of the climb in- creases). This is not predicted by a model with a Gaus- sian distribution such as a Brownian walk [Weissel et al., 1994] . This discrepancy between Gaussian models of topography and observed topography is consistent with the observation of Mandelbrot [1983], who found that when he transformed his Gaussian synthetic to- pography with a cubic function, the topography looked more realistic.

Many studies have attempted to model the evolu- tion of drainage networks. Recent papers have em- phasized their fractal properties. Some authors de- scribe discretized models that follow the flow of dis-

crete units of runoff down the steepest slope and erode the hillslope with an assumed dependence of denuda- tion on slope (and possibly other factors) [Willgoose et al., 1991; Chase, 1992; Kramer and Marder, 1992; Leheny and Nagel, 1993; Inaoka and Takayasu, 1993; Howard, 1994; Kooi and Beaumont, 1996]. Willgo- ose et al. [1991], Leheny and Nagel [1993], Kramer and Marder [19921, and Inaoka and Takayasu [1993] have shown that •heir models produce drainage ne•- works consistent with the observed scaling behavior of real drainage networks. •hase [1992] has presented a model that can produce fractal topography with frac- ml dimensions broadly consistent with observed values. Models have also been proposed which do no• explicifiy model the landform topography but only •he growth of the drainage network. Stark [1991] has proposed a model based on self-avoiding percolation clusters. Kon- doh and Matsushita [1986], Meakin et al. [1991], Masek

and Turcotte [1993], and Stark [1994] have presented models based on diffusion-limited aggregation (DLA) and variants of DLA. Masek and Turcotte [1993] in- troduced random walkers randomly into the landscape rather than at the boundaries (as in traditional DLA) to better model the effect of storms producing runoff randomly in space and time. Other approaches include that of Newman and Turcotte [1990], who propose a cascade model similar to Kolmogorov's model of the turbulent cascade in which the variance at each scale

is dependent only on that scale and the next largest one. Several studies have proposed continuum growth equations. $ornette and Zhang [1993] have advocated a model equation known as the Kardar-Parisi-Zhang (KPZ) equation, originally introduced in the physics literature to model the growth of atomic surfaces by ion deposition [Kardar et al., 1986], as a model for the evolution of topography by geomorphic processes. New- man and Turcotte [1990] and $ornette and Zhang [1993] have stressed the necessity of nonlinear terms in order to produce scale-invariant topography. The KPZ equa- tion is

Oh A (Vh)•. + r/(x, y t) (1) Ot = DV2h + • ' where r/(x, y, t) is a Gaussian white noise. The first term on the right-hand side represents the classic Culling [1960, 1963] diffusion model of slope erosion. Culling hypothesized that the horizontal flux of eroded mate- rial was proportional to the slope. With conservation of mass this yields the diffusion equation. Solutions to the diffusion equation have been successfully applied to modeling the evolution of alluvial fans, incised chan- nels, prograding deltas, and eroding fault scarps [Wal- lace, 1977; Nash, 1980; Begin et al., 1981; Gill, 1983a,b; Hanks et al., 1984; Hanks and Wallace, 1985; Kenyon and Turcotte, 1985; Phillips and Sutherland, 1986]. The last term in (1) represents spatial variations in erosion intensity produced by the intermittent nature of runoff production and mass movements, spatial variations in the erodibility of the soil and rock of the landscape, and episodic tectonic uplift. The inclusion of a stochastic term such as this is a universal feature of models consid-

ered in the physics literature which produce self-affine or fractal surfaces. $ornette and Zhang [1993] have ar- gued that the nonlinear term in (1) is essential for the formation of a scale-invariant surface. The nonlinear

term is equivalent to assuming that the erosion rate is proportional to the local exposed landscape surface. $omfai and Sander [1997] have also employed the KPZ equation and have produced landscapes which obey the observed scaling between drainage area, main channel length, and Strahler order with simulations incorporat- ing only the nonlinear term in the the equation for the local erosion rate of the surface. Smith and Bretherton

[1972] presented a generalized model of hillslope evolu- tion which included the KPZ nonlinear term as a special case. Giacometti et al. [1995] have proposed a contin-

Page 3: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7561

uum growth equation that includes higher-order terms not present in the KPZ equation.

It should be emphasized that a great variety of mod- els, including most of those listed, obey Horton's laws. Horton's laws state that stream number, average stream length, and average drainage area decrease as a power law with stream order defined by the Strahler order- ing scheme. In this scheme, a stream with a chan- nel head is defined as a first-order stream. When two

like-order streams combine, they form a downstream segment one order higher than the order of the trib- utary streams. Since so many different models obey Horton's laws, agreement with Horton's laws does not appear to be sufficient to verify that a model is an accurate representation of drainage network formation [Willgoose, 1994]. In fact, Kirchner [1993] has argued that Horton's laws are satisfied by virtually all possible branching networks. Costa-Cabral and Burges [1997] have reevaluated the hypothesis of Kirchner [1993] and find that Horton's laws are not necessarily satisfied by all possible networks. Either way, Kirchner's work has reminded us that similar measures of river basin mor-

phology between model and real basins may not enable us to validate our model unless the measures are diag- nostic. Tokunaga [1984] and Peckham [1995] improved the classification of stream order by introducing a more stringent test for models.

In this paper we investigate a model of landform evo- lution with topographic evolution by geomorphic pro- cesses parameterized with the diffusion equation. The model we propose reproduces the observed statistical behavior of both topography and river networks. The diffusion equation models the relaxation of topographic variations by transport-limited overland and channel flow. In section 2 we show that diffusivity is propor- tional to river discharge to the three-halves power. It is this dependence on discharge, itself dependent on the basin morphology, that gives rise to a complex unstable behavior from such a simple model. Our first analysis of the model will parameterize the diffusivity as a func- tion of elevation and perform a Taylor expansion retain- ing the spatial variability of diffusivity to first order in the elevation. A nonlinea• partial differential equation is derived and solved which reproduces the hypsome- try and power spectral behavior of real topography. In section 4, numerical simulations are performed which explicitly include river networks and calculate the dis- charge at every point. Realistic river networks and to- pography are generated.

2. Dependence of Diffusivity on Discharge

Our model consists of applying Culling's [1960, 1963] classic diffusion model of hillslope evolution to the hill- slopes in our model. For channels, we will also as- sume that sediment flux is proportional to slope as in

Culling's model, but now we will allow the diffusiv- ity to be a function of discharge, consistent with the fact that large rivers carry more sediment than small rivers. A diffusion model with a discharge-dependent diffusivity is consistent with the laboratory measure- ments of aggradation and degradation in channels by Gill [1983a, 1983b] and Phillips and Sutherland [1986]. To determine the dependence of the diffusivity on dis- charge, we have calculated suspended sediment rating curves for all of the stations in the U.S. Geological Survey (USGS) National Stream Water-Quality Mon- itoring Network [U.S. Geological Survey, 1996] with at least 30 measurements of instantaneous sediment flux

and discharge. The relationship between instantaneous suspended sediment discharge and instantaneous water discharge is often observed to be a power-law function: L - aQ b, where a and b are constants [Leopold et al., 1964]. The values of a and b define the sediment rat- ing curve for that location. The probability distribution function for the values of b from the U.S. Geological Sur- vey data is plotted in Figure 1. The distribution has a mean value of 2.03, although there is considerable vari- ation with a skewed distribution about this mean value.

For our purposes, we will adopt the mean value of 2.0 and assume that sediment flux in rivers is proportional to the square of the discharge.

The dependence of sediment flux on discharge is not the same as the dependence of diffusivity on discharge, however, because it is the sediment flux per unit chan- nel width that determines the diffusivity to be used in the the landform evolution equation. Since the width of a river is proportional to the square root of the dis- charge [Leopold et al., 1964], the diffusivity of a channel is proportional to Q2/w oc Q2/Q« oc Q•. Therefore, in our model we will have a constant diffusivity on hill- slopes and a diffusivity proportional to the three-halves power of discharge in channels:

0.14

0.12

0.10

0.08 -

0.06

0.04

0.02

0.00

0.5 1.0 1.5 2.0 2.5 3.0 3.5 b

Figure 1. Probability density function of the sedi- ment rating curve exponents b for all 470 stations with more than 30 measurements from the USGS National

Stream Water-Quality Monitoring Networks [U.S. Ge- ological Survey, 1996]. The mean value of b is 2.03.

Page 4: Self-organization and scaling relationships

7362 PELLETIER: FRACTAL RIVER NETWORKS

lO 8

10 6

10 4

10 2

Qav •A

' I

10 ̧ 10 -2 10 10 4 10 5

I I I

-1 10 ̧ 101 10 2 10 3 A (km 2)

Figure 2. Relationship between average discharge and drainage area for 672 hydrologic monitoring stations in the Mississippi River basin. Data from Slack and Landwehr [1992].

Oh

O-• = V. [o(q)v•] (2) where D is constant if the site is a hillslope and D oc Q • if the site is a channel.

In order to solve (3) it is necessary to parameterize discharge in terms of the morphology of the basin. Dis- charge is principally a function of drainage basin area. It is often assumed that discharge and area are propor- tional [Rodriguez-Iturbe et al., 1992]. In order to test this assumption we have plotted data for discharge and contributing area for 672 hydrologic monitoring stations in the Mississppi River basin of the U.S. Geological Sur- vey using the Hydro-Climatic Data Network [Slack and Landwehr, 1992]. The results are presented in Figure 2. The data indicate that Q oc A. A different rela- tion was found in the data set of Mulder and Syvitski [1996] using river basins from around the world. They found Qa,oc A ø'74. The variation from linearity that Mulder and Syvitski [1996] observed may be a result of including data from both arid and humid regions. In a uniform humid, midlatitude climate such as in the Mississippi River basin, a proportional relationship is obtained. The model equation as a function of area is then

Oh

= v. (3) where D - Dh if the site is a hillslope and D oc Q« - DcA] if the site is a channel and where Dh and Dc are positive constants. This model is equivalent to that of Willgoose et al. [1991] for the particular values mx - 3/2 and n• - 1 in their equations.

One simple way to model growth of the network is to start a simulation with one or more small chan-

nels draining to the border of the lattice and to ex- tend the channel from these locations at a given rate when the area drained by the grid point, a proxy for discharge, exceeds a given threshold. Such a model of

headward growth is supported by field studies [Patton and Schumm, 1975; Begin and Schumm, 1979]. It was also used in the model of Willgoose et al. [1991]. Hills- lope grid points become channels if they have a neigh- boring point that is a channel. The rate of channel advancement has been observed to be proportional to A«. In our model, therefore, channels grow with a rate proportional to A« when the drainage area exceeds a threshold value.

3. Approximate Solution of the Model Equation

Discharge within a single basin is a function of basin relief, the difference in elevation between the highest point in the basin and the point where the discharge is measured. In first-order streams the discharge is very small compared to lowland rivers. In this section we parameterize the discharge as a function of elevation to obtain a single differential equation for the local ele- vation in space and time, h(x, y,t). This approach is only approximate, however, since there is not a one-to- one correspondence between discharge and relief, only a general trend for rivers of large discharge to be in low- lands. An exact solution to the model equation (3) will require that the contributing area be calculated for each point in the basin exactly. This will be done in section 4. The treatment in this section allows us to make a

connection between (3) and an equation introduced in the physics literature for which the statistics of the so- lutions are well known. The solutions to this equation are found to have a hypsometry and power spectrum in agreement with real topography dominated by fluvial erosional processes.

In section 4 we will present the results of morpho- metric analyses of river networks. We will show the drainage area to be a power function with exponents between 3 and 5 of the basin relief, h- hmax, where h is the elevation of the outlet and hmax is the maximum elevation of the basin. Since the diffusivity is defined to be proportional to drainage area to the three-halves power, the dependence of the diffusivity on relief is also a power law with exponent a:

D oc (h- hmax)" (4)

We will make a linear approximation to this power law relationship: D • Do- Dlh, where Do and Dx are pos- itive constants, to keep only the leading-order nonlinear term in the resulting erosion equation.

The diffusion equation with spatially variable diffu- sivity is

Oh = v. ot

OD

= (n)vn + (vn) The diffusivity must be kept inside the gradient term since it is not a constant.

Page 5: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7363

Substituting (4) into (5) and keeping only first-order terms in h gives

Oh = _ ot

The two terms on the right side of (6) have a markedly different effect on the evolution of the hypsometry of the topography [Barabasi and Stanley, 1995]. The diffusion term maintains the form of the hypsometric curve as the topography is eroded. In contrast, the nonlinear term erodes material where the slope is large leading to a skewed hypsometry with a larger fraction of topography in lowlands. The effect of the two terms are illustrated

by Figures 5.4 and 6.2 of Barabasi and Stanley [1995]. In order to model episodic erosion and deposition,

we will add a Gaussian white noise term to (6). The result is the KPZ equation (1). There is abundant empirical evidence that spattally and temporally vari- able erosion rates are universal over a wide range of timescales and length scales. Much of the sediment carried away in rivers is carried away in intermittent storms whose occurrence can best be described statis- tically. An example of this intermittency is the time series of sediment load in the Santa Clara River. The

sediment load of the Santa Clara was carefully moni- tored for a period of 18 years. Over half of the total sediment yield carried by the river was transported in only three large floods totaling 7 days [Milliman and $yvitski, 1992]. Similar large bursts of sediment trans- port are evident in the sediment transport time series of Plotnick and Prestegaard [1993] on timescales of hours. Many examples of large floods which have resulted in major landscape modification have been documented in the geologic record [Bretz, 1969; Meyer and Nash, 1983; Ager, 1993]. Spattally and temporally variable erosion rates have been documented over a variety of spatial and temporal scales and geomorphological set- tings [Luk, 1982; Omni, 1982; Schmidt, 1985]. A steady state competition between episodic erosion and depo- sition and subsequent relaxation of the landscape has been argued to be the essential dynamic of landscape evolution [ Wolman and Getson, 1978]. A model of !and- scape evolution as a sequence of epsodic events is consis- tent with the age distributions of rock avalanches and sedimentary sequences [Grifflths, 1993]. These obser- vations suggest that any model of landform evolution must include stochastic spatial and temporal variabil- ity in erosion rates.

The solution to the KPZ equation is a self-affine or fractal surface with linear transects that have power law power spectra with an exponent of-1.8' S(k) oc k -•'s [Amar and Family, 1989]. In their simulations, Amar and Family [1989] nondimensionalized the KPZ equation to obtain

Oh = + + y, t) (z) ot

where the noise term, •/(x, y, t), now has a variance of

one, e - erDa/Do is a nondimensional parameter that quantifies the strength of the nonlinear term, and er is the standard deviation of the original noise term in (1). The surface generated by this equation reaches a steady state balance between the smoothing effects of diffusion and the toughening effects of the nonlinear and noise terms. In that steady state condition, the surface has linear transects that have power law power spec- tra with an exponent of-1.8. This can be concluded by relating the Hansdorff measure reported by Amar and Family [1989] for the KPZ model, Ha = 0.40, to the one-dimensional power spectral exponent/3 through the relation/3 - 2Ha + 1 [e.g., Turcotte, 1992]. It is not clear whether there is a value of e low enough for this result to be no longer valid since it takes longer for the surface to acheive a steady state condition as e goes to zero. However, Amar and Family [1989] found no evi- dence for a transition as they lowered e and found self- alline behavior with/3 = 1.8 even for small values of e; e must still be finite, however, for the resulting surface to be self-alline. Equation (1) without the nonlinear term is known as the Edwards-Wilkinson equation [Ed- wards and Wilkinson, 1982]. The Edwards-Wilkinson equation is not self-affine in three dimensions. A re-

view of scale-invariant models of surface growth models has been given by Krug [1997].

A cellular automaton model, the restricted solid-on- solid (RSOS) model, has been developed by Kim and Kosterlitz [1989] with rules that mimic the terms in the KPZ equation. We have performed simulations of this model in order to verify the finite difference simulations of Amar and Family [1989] and to obtain the hypsom- etry of the KPZ equation. The reason why this ap- proach is preferable to solving the KPZ equation di- rectly by finite differencing is that it is much quicker to simulate since in the finite difference scheme, very small time steps are required to maintain numerical stability. For this reason, the RSOS model and similar discrete models have been developed to mimic the terms in the KPZ equation. The RSOS model is particularly useful for solving the KPZ equation in three dimensions since this is the dimension where the strength of the diffu- sion term is closest to that of the roughening terms and the equation must be solved very carefully to obtain accurate results using finite differencing [Newman and Bray, 1996]. Park and Kahng [1995] have shown that the RSOS model is equivalent to the KPZ equation in the continuum limit. In the RSOS model, a site on a two-dimensional lattice of points is chosen at random. The height of the surface at that point is incremented by one if the elevation at that point is greater than or equal to the elevations of all of its four nearest neigh- bors. If this restriction is not satisfied, nothing happens. This rule is repeated until a surface with a height equal to or greater than the linear dimension of the lattice is generated. This model corresponds to the KPZ equa- tion with a particular value of e. Periodic boundary conditions are used. Here we report the results of sim-

Page 6: Self-organization and scaling relationships

7364 PELLETIER: FRACTAL RIVER NETWORKS

Figure 3. Shaded relief image of topography produced with the restricted solid-on-solid (RSOS) model.

ulations of this model performed on a 256 x 256 lattice. Other lattice sizes yielded similar results. A shaded re- lief image of an example of a surface generated with this model is shown in Figure 3. The average power spec- trum, estimated as the square of the coefficients of the fast Fourier transform, of linear transects of this sur- face is presented in Figure 4. The power spectrum of each row of the lattice was computed and then averaged with the power spectra of all other rows in order to ob- tain the average power spectrum. A good match with the power spectrum $(k) or k -•'8•, indicated by the straight line, is obtained. This power spectrum agrees with the finite difference calculation of the KPZ equa- tion of Amar and Family [1989]. This power spectrum is close to the spectrum $(k) or k -2 observed in erosional topography by Huang and Turcotte [1989] and others. The power spectral behavior observed in real topogra- phy is independent of the age of the topography (time since significant uplift occurred) and the initial relief following tectonic uplift. For instance, young, rough mountain ranges, such as the Rocky Mountains, exhibit the same power spectral exponent or fractal dimension as smooth mountain ranges, such as the Appalachians. Similarly, in the RSOS model a steady state condition is acheived once a rough surface is produced. In the steady state condition the smoothing effects of the dif- fusion term are balanced by the roughening effects, and the power spectral behavior is independent of time. A rougher landscape, defined as a larger variance per unit wavelength, can be produced by increasing the ratio of the variance of the stochastic term to the diffusion con-

stant. Rougher topography has a larger power spectral density. However, the power spectral exponent, which quantifies the relative amplitude of topography at dif- ferent wavelengths, is the same for rough or smooth topography. Changing the diffusion constant does not affect the fractal character of the topography or the -1.8 exponent. The steady state condition is achieved independent of the initial conditions of the simulation. For instance, the simulation can be initiated with a self- affine surface with $(k) or k-•'8. The erosion of the surface by the rules of the RSOS model will then main-

tain the fractal character of the surface from that time

onward.

The topography of Earth's surface is skewed such that a much larger percentage of Earth's topography is low- lands and there is a positive correlation between eleva- tion and slope. This skew can be associated with the nonlinear term in (3). If only the diffusion term were present, the resulting topography would have a Gaus- sian distribution. This is because any linear transfor- mation of a function with a Gaussian distribution, such as the noise term in (1), results in a function with a Gaussian distribution. The probability density func- tion (pdf) of elevations produced by the RSOS model is presented in Figure 5a. The pdf was computed using the surfaces generated from 10 simulation runs. The pdf's for individual runs were the same except for small random variations resulting from the finite length of the simulations. The observed pdf is not Gaussian but is skewed such that the most probable elevation is be- low the mean elevation of the landscape; that is, more of the total landscape area is represented by lowlands than highlands. The distribution is not smooth be- cause of the discrete nature of the model. The prob- ability distributions for the Kentucky and Mississippi River basins are given in Figures 5b and 5c, respectively. The ETOP05 data set [Loughridge, 1986] was used to compute the hypsometry of the Mississippi River basin and the USGS 1 ø digital elevation models (DEMs) [U.S. Geological Survey, 1990] were used to compute the hyp- sometry of the Kentucky River basin. Both hypsome- tries exhibit skew toward lower elevations. This skew is

directly comparable to the skew in the RSOS topogra- phy and can be associated with the nonlinear term.

Figures 6a-6c shows the cumulative percentage of area larger than a given area computed by Harrison et al. [1983] using the ETOP05 data set [Loughridge, 1986] for the continents of Africa, North America, and South

lO 3

lO 2

1 lO

lO 0

10-1 1 10 100

k

Figure 4. Average power spectrum $ as a function of wave number k for one-dimensional transects of the sur-

face generated with the RSOS model. A least squares fit to the logarithms of the ordinate and abscissa yield a slope of -1.81, indicating that $(k) or k -•'8•.

Page 7: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7365

0.20

,- 0.15

-o 0.10 N

o_

E o 0.05

0.00

-10

,

-5 0 5

difference from mean

10

......... I ......... i ......... i ......... i ......... i ....... [•

lOO 200 500 400 500 600 700

difference from mean (m)

E 2.5

"• 2.0 o

o_

'- 1.5

o_

c 1.0

:-_- 0.5

o ,• 0.0

metric curves appear to approach a skewed lognormal distribution as the age since significant uplift increases. Of the three continents, Africa has experienced the least neotectonic uplift. South America has experienced the most. The hyposmetry of Africa matches a lognormal distribution more closely than South America where the

1.0

0.6 _

0.4 _

0ø2

0.0

0.0

1.0

.c 0.8

i i i

0.2 0.4 0.6 0.8

normalized elevation h

1.0

0.6

0.4

0.2

0.0

0.0

1.0

-200 0 200 400 600 800 1000 1200

difference from mean (m)

Figure 5. (a) Probability density function of elevations of the surface (hypsometry) produced by the RSOS model. (b) Probability density function of the Ken- tucky River basin. (c) Probability density function of the Mississippi River basin. All three exhibit a signif- 0.0 icant skew such that lowlands (topography below the median elevation) make up a larger fraction of the total 0.0 area than highlands.

0.2 0.4 0.6 0.8 1.0

normalized elevation h

America, with a least squares fit to a lognormal distri- bution. The hypsometric curves presented in the inde- pendent study of Cogley [1985] are similar to those pre- sented by Harrison et al. [1983]. The calculated hypso-

0.8

e 06 o

0.4

0.2

South America

0.2 0.4 0.6 0.8 1.0 normalized elevation h

Figure 6. Cumulative hypsometric curves (dots), the fraction of area above an elevation normalized by the maximum elevation, for (a) Africa, (b) North America, and (c) South America, from the data of Harrison et al. [1983]. The line accompanying each plot is the least squares fit to a lognormal distribution for each conti- nent.

Page 8: Self-organization and scaling relationships

7366 PELLETIER: FRACTAL RIVER NETWORKS

Figure 7. Drainage networks analyzed: (a) Kumaun, (b) loess plateau, Shanxi Province, (c) Schoharie Creek, (d) Nepal, (e) Kentucky River, (f) Mississippi River, and (g)Bhutan.

presence of the relatively young Andes mountains re- sults in a significant deviation from a lognormal dis- tribution. This suggests that continental hypsometric curves approach a lognormal distribution as erosion has more time to act on the landscape.

4. River Networks

The objective of this section is to incorporate river networks into (3). In section 3, low elevations had a greater diffusivity than higher elevations but an equal value for equal elevations. In this section the diffusivity is calculated exactly for each point by computing the flow draining into each pixel in the landscape. With

this improvement, river networks can be incorporated in the simulations. There are several power law rela- tionships between the morphologic components of river basins. Some of these components are based on the Strahler ordering scheme. A partial list of the observed morphological relations in natural river networks is as follows:

1. Horton [1945] defined three ratios Rs, RL, and RA to be the ratio between the number, average length, and average drainage area from the streams of one order to those of the next lowest order. He found this ratio to be

a constant for all stream orders. The fractal dimension

of river networks is defined as D = log Rs / log RL. 2. An improved classification scheme leading to a

Page 9: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7367

relation similar to Horton's laws has been developed by Tokunaga [1984] and Peckham [1995]. They defined matrix elements To,• as the number of side tributaries of order k of streams of order o. Natural river networks

satisfy the constraint that To,o-• is constant for all o. The classic DLA growth model satisfies this property [ Ossadnik, 1992].

3. Hack [1957] found that the length of a main chan- nel scales with the drainage area according to power law: L cx: A q. He reported values of q • 0.6 for two river basins. Gray [1961] obtained a value of 0.57, and Mueller [1973] obtained 0.55. Other morphometric analyses based on hundreds of river basins, however, have found no significant deviation from 0.5 [Mont- gomery and Dietrich, 1992; Mulder and Syvitski, 1996].

4. Along-channel slope is a power law function of --0 5 discharge with an exponent close to -0.5: $ o• Qav'

[Carlston, 1968]. Together with the assumption that Q o• A, this yields S o• A-«. The relationship S o• A-« has been reported in studies by Rodriguez-Iturbe et al. [1992].

The fractal properties of river networks have been reviewed by many authors including La Barbera and Rosso [1989], Beer and Borgas [1993], Nikora [1994], and Maritan et al. [1996]. Moussa [1997] has argued that river basins are a fractal Sierpinski space.

In order to better characterize the morphometric re- lations between drainage basin components we have car- ried out river network extraction and analyses on seven basins from a variety of tectonic environments using high-quality digital elevation models. Four river net- works were chosen from the composite DEM of Field- ing et al. [1994]. This data set has 80 m resolution and does not rely on the interpolation of contour lines. Such interpolation, as is done in the USGS 1 ø DEMs, may lead to biased slope estimates. Three of these basins are located along the Himalayan front in Nepal, Ku- maun, and Bhutan. The fourth is located in the Shanxi Province, China, and is formed in loess. This basin has an unusally ordered shape characterized by a high de- gree of symmetry and unusually straight valleys. The remaining three networks are located in North America. The Kentucky River basin and Schoharie Creek basin were extracted from USGS 1 ø DEMs [U.S. Geological

Table 1. Summary of Results of River Network Analyses

Relation Observed Value Model Value

N o• A p p • -0.99 to -0.95 p • -0.95 L o• A q q • 0.5-0.55 q • 0.5-0.6 5' o• A r r • -0.37 r • -0.37 R o• A s s • 0.20 to 0.35 s • 0.30-0.35

A cx: R• RA • 10 ø'aT • 4.6 R•4 • 10 ø'aT • 4.6 To•u k u• 2.5 u•2.5

RB 4.6 4.6 R•; 2.2 2.3

lO 5

z

lO ̧

-5 lO

lO

i I I i

-0.97

-2 10 ̧ 10 2 10 4 10 6 A (kin 2)

Figure 8. Plot of average number of streams N of a given Strahler order as a function of the average drainage area A for that Strahler order for each of the seven river basins. The plots are offset from or.e an- other so that they may be placed on the same graph. The plots correspond, from top to bottom, to the river basins in Figures 7a-7g, respectively. The data indicate that N • A -•.

Survey, 1990]. The Mississippi River basin was chosen so that a large basin was represented. The Mississippi was extracted from the ETOP05 data set [Loughridge, 1986]. The seven basins are plotted in Figure 7. The river network extraction and analyses were carried out down to the pixel size of the DEM and only rivers with Strahler orders larger than three were plotted so that the network can be identified. The river network ex-

traction and analyses were carried out with RiverTools 1.01 (S.D. Peckham, electronic software available at / / cires. color ado. edu/people/peckham.scott / RT. html).

The analyses we carried out enable us to identify five independent morphologic relationships between six components. The results are shown in Figures 8-12 and are summarized in Table 1. The six morpho- metric components are stream number of a particular Strahler order N, Strahler order o, main channel length L, drainage area A, basin relief R, and along-channel slope $. The relationships between the variables are defined as

N o• A p (8) L cx: A q (9) $ o,: .A" R o• A • (11) A o• R• (12)

where p, q, r, and s are arbitrary powers. The exponent o in (13) is the Strahler order.

Figure 8 presents the total number of streams of a given Strahler order as a function of the average basin area for that Strahler order for each of the seven basins.

The order of plots in Figures 8-13 is, from top to bot-

Page 10: Self-organization and scaling relationships

7368 PELLETIER: FRACTAL RIVER NETWORKS

lO 6

lO 4

102

10 ̧

10-2

-4 lO

lO -2 lO ̧ lO 2 lO 4 lO 6 A (kin 2)

Figure 9. Plot of average main channel length L of a given Strahler order as a function of the average drainage area A for that Strahler order for each of the seven river basins. The plots are offset from one an- other so that they may be placed on the same graph. The plots correspond, from top to bottom, to the river basins in Figures 7a-7g, respectively. The data indicate that L • A q with q .•. 0.5-0.55.

104 - _ _.._._-------• 0.23

102 _ •:•.•_.._.• v • 0.25

10 -2 •0.24 - -4 10 -

• I • I • I • I

10-2 100 102 2 104 106 A(km )

Figure 11. Plot of average basin relief R for a given Strahler order as a function of the average drainage area A for that Strahler order for each of the seven river

basins. The plots are offset from one another so that they may be placed on the same graph. The plots corre- spond, from top to bottom, to the river basins in Figures 7a-7g, respectively. The data indicate that R • A -s with s ranging from 0.20 to 0.35.

tom, the Kumaun basin, the Loess plateau of the Shanxi Province, Schoharie Creek, the Nepal basin, the Ken- tucky River basin, the Mississippi River basin, and the Bhutan river basin. The plots are offset so that they may be placed on the same graph. Figure 8 indicates that N is approximately proportional to A -•, indicat- ing that p .•. -1.

Figure 9 presents the relationship between L and A. The plots indicate that q • 0.5. There has been consid-

lO 4

10 ¸ cn -2

10 i - 0 -4

10 -6 • -0.22 -0.34

I I I • I

10 -2 10 ¸ 102 104 10 6 ^ 2)

Figure 10. Plot of average along-channel slope $ of a given Strahler order as a function of the average drainage area A for that Strahler order for each of the seven river basins. The plots are offset from one an- other so that they may be placed on the same graph. The plots correspond, from top to bottom, to the river basins in Figures 7a-7g, respectively. The data indicate that $ • A -ø'a7.

erable debate about whether q approaches 0.5 exactly or whether there is a significant deviation. The reason for the debate is that if q ?• 1/2, then this may represent a deviation from self-similarity [Ijjasz-Vasquez et al., 1993]. Our analyses exhibit a variation in q from basin to basin that is roughly equal to the previously reported deviations from 1/2. Therefore we cannot definitively conclude whether q differs from 1/2 in river basins as a general rule.

Figure 10 indicates that r • -0.37. This is inconsis- tent with previous studies that have reported $ oc A-«

lO lO

lO 5

lO ̧

-5 10

..... 0.'65_

0.67 0.70

0.69

0.67

0.64

0.65

• , ,

o 2 4 6 8 10 1_•

Figure 12. Plot of average basin area A of a given Strahler order as a function of the Strahler order o for each of the seven river basins. Note the log-linear scale. The plots are offset from one another so that they may be placed on the same graph. The data indicate that RA is constant and equal to • 100'67 • 4.6.

Page 11: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7369

lO lO

105

100

10 -5

0.39

0.39

0.33

0.48 0.34

0.39

1 3 5 7 9 k

Figure 13. Plot of the Tokunaga ratio Tk a function of the Strahler order k for each of the seven river basins. Note the log-linear scale. The plots are offset from one another so that they may be placed on the same graph. The data indicate that T• o( u n with u • 10 ø'4ø • 2.5.

[Tarboton et al., 1989]. Tarboton et al. [1989] carried out analyses on two river basins with 3 orders of mag- nitude of area in the analyses. Given that our analy- ses were carried out on several basins with 5 orders of

'magnitude of area with a high-quality DEM, we pro- pose that r • -0.37 is a more reliable estimate. This conclusion appears to be consistent with the data of Montgomery and Dietrich [1988], who presented similar

slope-area relationships. Although no exponents were obtained, the trqnds of their data follow a power law relationship with exponent significantly greater than - 1/2 (i.e., closer to -0.37).

Plotted in Figure 11 are the relationships between re- lief R and area A. A range of values is observed between s .• 0.20 and s • 0.35.

The relationship between basin area and Strahler or- der o is plotted in Figure 12. In Figure 12, area is plotted on a logarithmic (base 10) scale, while Strahler order is plotted on a linear scale. The slopes are ob- served to be • 0.67. Thus the Horton ratio R• is equal to 10 ø'47 • 4.6. Using the other morphometric relations, we can estimate R• a•d Rs as 2.2 a•d 4.6, respectively.

The result of Tokunaga side-branching statistical anal- yses for the seven drainage basins is presented in Figure 13. In this ordering scheme, matrix elements To,• are defined to be the number of side tributaries of Strahler

order k of streams of Strahler order o. Natural river

networks satisfy the constraint that To,o-• is constant for all o. We have determined T• by averaging the vales of To,o-• over o:

n-k

o--1

Figure 13 plots T• on a logarithmic (base 10) scale as a function of k on a linear scale. The plots indicate that T• o( u • where u is estimated to be 10 0'40 • 2.5.

Figure 14. Grayscale plot of the elevation of the model river network when (a) 1/8, (b) 1/4, (c) 1/2, and (d) all of the grid points of the 64 x 64 lattice has become channelized. The elevations are mapped to brightness scale with a gamma function with a coefficient of 2.0.

Page 12: Self-organization and scaling relationships

7370 PELLETIER: FRACTAL RIVER NETWORKS

Figure 15. Model river networks produced after all of the grid points have become channels. The width of the river is proportional to its order. (a) The fully deterministic model. (b) A model run where the diffusivity is allowed to have a stochastic variation with a standard deviation 10% of the mean.

We now consider the numerical simulation of (3). The resulting model river networks will be analyzed in the same manner as the real river networks analyzed above. Equation (4) was simulated on a square lattice of 64 x

64 grid points and the equation was discretized in space and time. Integration in time utilized the predictor- corrector method which varies the time step to ensure

stability. The initial condition was an elevation of 1.0

Page 13: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7371

104

105

102 z

1 lO

lO ̧ _

lo-li lO ̧

ß ..... I I i

lO 1 lO 2 lO 5 A

-0.95-

-0.95

-0.94-

Figure 16. Plot of average number of streams N of a given Strahler order as a function of the average drainage area A for that Strahler order for three model river basins. The plots are offset from one another so that they may be placed on the same graph. The plots correspond, from top to bottom, to the model river basins produced with a deterministic 64 x 64 model run until all of the grid points were channels, a determinis- tic 64 x 64 model run until 50% of the grid points were channels, and a 64 x 64 model with small (10%) stochas- tic varaitions in the diffusivity run until all of the grid points were channels. The data indicate that N • A -z, similar to that observed for real river networks.

plus a small random noise on every point of lattice ex- cept for the upper left corner which was moved down to 0.0 at t-0 and fixed at zero for the entire calculation.

The curvature at all of the edges was fixed to be zero ex- cept for the upper left corner which was unconstrained. The contributing area at each point was continuously updated. Each grid point drained to the lowest of its

lO 3

0.60

102 O.56 0.59

101

10 ¸

lO ¸ 1• 1 1• 2 1• 5 ' ,, A

Figure 17. Plot of average main channel length L of a given Strahler order as a function of the average drainage area A for that Strahler order for each of the three model river basins. The plots are in the same order as Figure 16. The data indicate that L • A q with q .• 0.5-0.6, similar to that observed for real river networks.

lO -1 - o

-0.58',

10-2 5t -0.5

I ..................... ,'% ,-0_.•6 I

10 ̧ 101 102 105 A

Figure 18. Plot of average along-channel slope $ of a given Strahler order as a function of the average drainage area A for that Strahler order for each of the three model river basins. The plots are in the same or- der as Figure 17. The data indicate that $ • A -ø's?, the same relationship observed in real river basins.

nearest neighbors. If a grid point was moved down less than its neighbors such that it no longer was the low- est neighboring grid point for one of its neighbors, the contributing area of each grid point downhill of the old drainage path was decremented by one and the con- tributing area of each grid point downhill of the new drainage path was incremented by one. The thresh- old drainage area for channel growth was zero so that eventually all of the grid points in the model become channelized.

Figures 14a-14d plots the elevations of the model sur- face at four times. The surface elevations are mapped with a brightness scale with a gamma function with a coefficient of 2.0. The four times correspond to those where 1/8 (Figure 14a), 1/4 (Figure 14b), 1/2 (Figure

_ ß 0.29

100 o 0.55

ß ß 0.29

10 -1

i ........ I , , , , , i,iI ....

10 ̧ 101 102 105 A

Figure 19. Plot of average basin relief R for a given Strahler order as a function of the average drainage area A for that Strahler order for each of the three model river basins. The data indicate that R .• A -s with

s •0.3, consistent with observations.

Page 14: Self-organization and scaling relationships

7372 PELLETIER: FRACTAL RIVER NETWORKS

14c), and all of the lattice points (Figure 14d) have be- come channelized. The erosion front is observed to be

unstable, resulting in an irregular shape which grows over time.

In Figure 15a the river network corresponding to the surface of Figure 14d is presented. In this plot the river width is made equal to the Strahler order so that thicker rivers indicate those that drain more area. In Figure 15b, the river network created with the same param- eters as Figure 15a is given but with some stochastic variability included in the sediment transport. For each pixel and each time step a factor (1 + 0.1•) was multi- plied times Ah, where V is a Gaussian white noise with mean zero and standard deviation of one. As previously argued, spatial and temporal variations in erodibility are a universal feature of landscape evolution. The purpose of including this stochastic variability was to determine its effect, if any, on the morphology of the basin.

Figures 16-21 are plots of the morphologic relation- ships corresponding to those for real river basins pre- sented in Figure 8-13. In these plots, three river basins are analyzed. The results for the deterministic simula- tion given in Figure 14d are presented as the top plot. The middle line represents the results of morphometric analyses on the partially developed river basin of Figure 14c. The results of the stochastic river basin illustrated

in Figure 15b are presented as the bottom line. The results of the morphometric analyses on the real and model river basins are summarized in Table 1.

The same scaling relations are observed for the three model river basins, suggesting that the scaling laws we have identified are satisfied by river basins continuously as they evolve by headward growth and that they are satisfied even in the presence of heterogeneous erodibil- ity and/or runoff. Close agreement between model and real river basins was observed for all of the morphome- tric relationships.

103 ., ......... , ......... , ......... , ......

102

•.x 101

10 ̧

10-1

0.39

0.37

0.38

, i ......... i , , i • i , • i • i , i , , ..... i ......

1 2 3 4 k

Figure 21. Plot of the Tokunaga ratio Tk a function of the Strahler order k for each of the three model river

basins. Note the log-linear scale. The data indicate that Tk oc u • with u • 100'4 • 2.5.

The probability density function of the model basin of Figure 15a is presented in Figure 22. As observed in the Kentucky and Mississippi River basins, the hypsometry is skewed such that the most probable elevation is lower than the median elevation. The evolution of the hypso- metric curve (cumulative distribution function) for the model basins of Figure 14a-14d is shown in Figure 23 as the top to bottom lines, respectively. A close sim- ilarity exists between this sequence and the evolution observed in real basins. For example, the hypsometric curves observed over time at Perth Amboy, New Jersey, by $chumm [1956] are quite similar to those of Figure 23. Both the real and model basins begin with a nearly constant surface equal to the maximum elevation of the basin. As the channels dissect the basin, the hypsomet- ric curve smooths out. The basins reach equilibrium with a skewed pdf as evidenced by the fact that the

104 0.67 0.63

10 3 o 0.68

<: 102 _

101

10 ̧ -1

10 -. ............. , ..........

1 2 3 4 5 o

Figure 20. Plot of average basin area A of a given Strahler order as a function of the Strahler order o for

each of the three model river basins. Note the log-linear scale. The data indicate that RA is constant and equal to approximately 10 ø'75 • 4.6, the same value observed in real networks.

2.0 i

1.5

0.5

0.0

0.0 0.2 0.4 0.6

normolized elevotion

0.8 1.0

Figure 22. Probability density function of the model river basin of Figure 17. The skew in this pdf is directly analagous to that of the Kentucky and Mississippi River basins of Figures 5b and 5c, respectively, and can be as- sociated with the dependence of diffusivity on elevation.

Page 15: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7373

1.0

0.8

0.6

0.2

0.0

0.0 0.2 0.4 0.6 0.8 1.0

normalized elevation h

Figure 23. Evolution of the hypsometric curves for the model river basin for the four instants of time illustrated

in Figure 14. Increasing time results in a smoother hyp- sometric curve.

vation that channel and hillslope adjustment by sedi- ment transport can often be modeled by the diffusion equation with a diffusivity dependent on discharge. The dependence of diffusivity on discharge introduces a non- linear term in the partial differential equation for land- scape elevation which is responsible for the fractal na- ture of topographic transects and results in a charac- teristic hypsometric curve. The morphometric relation- ships for seven river basins have been computed and have been found to be remarkably universal. Model drainage basins have morphologies which closely ap- proximate those observed in nature.

Acknowledgments. I wish to thank Chris Duncan and Don Turcotte for helpful conversations. Special thanks is extended to Scott Peckham for making RiverTools 1.01, his remarkable river network and extraction program, available over the Internet. Most of this work would not have been possible without Scott's program.

model and real hypsometric curves fall below 1/2 at a normalized elevation of 1/2. In contrast, a Gaussian or any other symmetric distribution would have a median elevation of h/hmax = 1/2.

In previously studied deterministic models of alluvial landform evolution, sediment flux has often been pa- rameterized as a power law function of drainage area and slope: J or Am(Oh/Ox) '•, where J is the sediment flux and m and n are free parameters. In this paper we have argued that m = 3/2 and n = I are the most ap- propriate values to model alluvial landform evolution.

One of the most striking results of the analyses of real river basins carried out in this paper is the simi- larity of scaling relationships of the seven basins ana- lyzed despite the wide range of tectonic environments the basins are taken from. The four basins from the

Himilayan front are incised into bedrock, while the re- maining basins are alluvial. It is generally accepted that for bedrock channels the erosion rate, Oh/Or, not the sediment flux, is a power law function of area and slope [$eidl et al., 1994]:

Oh A• Oh at •c (•xx)- (14) where the exponents m and n may or may not be the same as the alluvial case considered in this paper. An important direction of future research will be to perform forward modeling studies of (14) and to analyze the resultant river networks in an attempt to explain the surprising similarity of alluvial and bedrock drainage networks.

5. Conclusions

We have presented a model of landscape evolution by overland and channel flow which reproduces many of the basic statistical features of topography and river basin morphology. The model is based on the obser-

References

Ager, D.V., The New Catastrophism, Cambridge Univ. Press, New York, 1993.

Ahnert, F., Local relief and the height limits of mountain ranges, Am. J. $ci., 28•, 1035-1055, 1984.

Amar, J.G., and F. Family, Numerical solution of a con- tinuum equation for interface growth in 2+1 dimensions, Phys. Rev. A, •1, 3399-3402, 1989.

Balmino, G., The spectra of the topography of Earth, Venus, and Mars, Geophys. Res. Left., 20, 1063-1066, 1993.

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

Beer, T., and M. Borgas, Horton's laws and the fractal na- ture of streams, Water Resour. Res., 29, 1475-1487, 1993.

Begin, Z.B., and S.A. Schumm, Instability of alluvial valley floors: A method for its assessment, Trans. ASAE, 22, 347-350, 1979.

Begin, Z.B., D.F. Meyer, and S.A. Schumm, Development of longitudinal profiles of alluvial channels in response to base level lowering, Earth Surf. Processes, 6, 49-68, 1981.

Bell, T.H., Statistical features of sea-floor topography, Deep Sea Res., 22, 883-892, 1975.

Bretz, J.H., The Lake Missoula floods and the channeled scablands, J. Geol., 77, 505-543, 1969.

Brown, S.R., and C.H. Scholz, Broad bandwidth study of the topography of natural rock surfaces, J. Geophys. Res., 90, 12575-12582, 1985.

Carlston, C.W., Slope-discharge relations for eight rivers in the United States, U.S. Geol. Surv. Prof. Pap., 600, 45- 47, 1968.

Chase, C.G., Fluvial landsculpting and the fractal dimen- sion of topography, Geomorphology, 5, 39-57, 1992.

Cogley, J.G., Hypsometry of the continents, Z. Geomorph. Suppl., 53, 1-48, 1985.

Costa-Cabral, M.C., and S.J. Burges, Sensitivity of channel network planform laws and the question of topological randomness, Water Resour. Res., 33, 2179-2197, 1997.

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

Culling, W.E.H., Soil creep and the development of hillside slopes, J. Geol., 71, 127-161, 1963.

Culling, W.E.H., and M. Datko, The fractal geometry of the soil-covered landscape, Earth Planet. $ci. Left., 12, 369, 1987.

Edwards, S.F., and D.R. Wilkinson, The surface statistics of

Page 16: Self-organization and scaling relationships

7374 PELLETIER: FRACTAL RIVER NETWORKS

a granular aggregate, Proc. R. Soc. London, Set. A, 381, 17-31, 1982.

Evans, I.S., and C.J. McClean, The land surface is not unifractal: Variograms, cirque scale, and allometry, Z. Geomorph. Suppl., 101, 127-147, 1995.

Fielding, E., B. Isacks, M. Barazangi, and C.C. Duncan, How fiat is Tibet?, Geology, 22, 163-167, 1994.

Giacometti, A., A. Maritan, and J.R. Banavar, Continuum model for river networks, Phys. Rev. Lett., 75, 577-580, 1995.

Gilbert, L.E., Are topographic datasets fractal?, Pure Appl. Geophys., 131, 241-254, 1989.

Gill, M.A., Diffusion model for aggrading channels, J. Hy- draul. Res., 21, 355-367, 1983a.

Gill, M.A., Diffusion model for degrading channels, J. Hy- draul. Res., 21, 369-378, 1983b.

Gray, D.M., Interrelationships of watershed characteristics, J. Geophys. Res., 66, 1215-1223, 1961.

Griffiths, G.A., Estimation of landform life expectancy, Ge- ology, 21, 403-406, 1993.

Hack, J.T., Studies of longitudinal profiles in Virginia and Maryland, U.S. Geol. Surv. Proj •. Pap., 29•B, 1957.

Hanks, T.C., and R.E. Wallace, Morphological analysis of the Lake Lahontan shoreline and beachfront fault scarps, Pushing County, Nevada, Bull. Seismol. Soc. Am., 7'5, 835-846, 1985.

Hanks, T.C., R.C. Buckman, K.R. Lajoie, and R.E. Wallace, Modification of wave cut and faulting-controlled land- forms, J. Geophys. Res, 89, 5771-5790, 1984.

Harrison, C.G.A., K.J. Miskell, G.W. Brass, E.S. Saltz- mann, and J.L. Sloan II, Continental hypsometry, Tec- tonics, 2, 357-378, 1983.

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

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: Applications to the topography of Arizona and comparisons with synthetic images, J. Geophys. Res., 9•, 7491-7495, 1989.

Ijjasz-Vasquez, E.J., R.L. Bras, and I. Rodriguez-Iturbe, Hack's relation and optimal channel networks: The elon- gation of river basins as a consequence of energy mini- mization, Geophys. Res. Lett.; 20, 1583-1586, 1993.

Inaoka, H., and H. Takayasu, Water erosion as a fractal growth process, Phys. Rev. E, •7, 899-910, 1993.

Kardar, M., G. Parisi, and Y.-C. Zhang, Dynamic scaling of growing interfaces, Phys. Rev. Lett., 56, 889-892, 1986.

Kenyon, P.M., and D.L. Turcotte, Morphology of a delta prograding by bulk sediment transport, Geol. Soc. Am. Bull., 96, 1457-1465, 1985.

Kim, J.M., and J.M. Kosterlitz, Growth in a restricted solid- on-solid model, Phys. Rev. Lett., 62, 2289-2292, 1989.

Kirchner, J.W., Statistical inevitability of Horton's laws and the apparent randomness of stream channel networks, Ge- ology, 21, 591-594, 1993.

Kondoh, H., and M. Matsushita, Diffusion-limited aggrega- tions with anisotropic sticking probability: A tentative model for river networks, J. Phys. Soc. Jpn., 55, 3289- 3292, 1986.

Kooi, H., and C. Beaumont, Large-scale geomorphology: Classical concepts reconciled and integrated with contem- porary ideas via a surface processes model, J. Geophys. Res., 101, 3361-3386, 1996.

Kramer, S., and M. Marder, Evolution of river networks, Phys. Rev. Lett., 68, 205-208, 1992.

Krug, J., Origins of scale-invariance in growth processes, Adv. Phys., •6, 139-282, 1997.

Kucinskas, A.B., D.L. Turcotte, J. Huang, and P.G. Ford, Fractal analysis of Venus topography in Tinatin Planatia and Ovda Regio, J. Geophys. Res., 97, 13635-13641, 1992.

La Barbera, P., and R. Rosso, On the fractal dimension of stream networks, Water Resour. Res., 25, 735-741, 1989.

Leheny, R.L., and S.R. Nagel, Model for the evolution of river networks, Phys. Rev. Lett., 71, 1470-1473, 1993.

Leopold, L.B., M.G. Wolman, and J.P. Miller, Fluvial Pro- cesses in Geomorphology, W.H. Freeman, New York, 1964.

Loughridge, M.S., Relief map of Earth's surface, Eos Trans. A GU, 67, 121, 1986.

Luk, S.-H., Variability of rainwash erosion within small sam- ple areas, in Space and Time in Geomorphology, edited by C. E. Thorn, pp. 243-268, Allen and Unwin, Winchester, Mass., 1982.

Mandelbrot, B., Stochastic models for the Earth's relief, the shape and the fractal dimension of coastlines, and the number-area rule for islands, Proc. Natl. Acad. Sci. U.S.A., 72, 3825-3828, 1975.

Mandelbrot, B., The Fractal Geometry of Nature, W. H. Freeman, New York, 1983.

Maritan, A., A. Rinaldo, R. Rigon, A. Giacometti, and I. Rodriguez-Iturbe, Scaling laws for river networks, Phys. Rev. E, 53, 1510-1515, 1996.

Masek, J.G., and D.L. Turcotte, A diffusion-limited aggre- gation model for the evolution of drainage networks, Earth Planet. Sci. Left., 119, 379-386, 1993.

Matsushita, M., and S. Ouchi, On the self-affinity of various curves, Physica D, 38, 246-251, 1989.

Meakin, P., J. Feder, and T. Jossang, Simple statistical mod- els for river networks, Physica A, 176, 409-429, 1991.

Meyer, L., and D. Nash (Eds.), Catastrophic Flooding, Blackwell Sci., Cambridge, Mass., 1983.

Milliman, J.D., and J.P.M. Syvitski, Geomorphic/tectonic control of sediment discharge to the ocean: The impor- tance of small mountainous rivers, J. Geol., 100, 525-544, 1992.

Montgomery, D.R., and W.E. Dietrich, Where do channels begin?, Nature, 336, 232-234, 1988.

Montgomery, D.R., and W.E. Dietrich, Channel initiation and the problem of landscape scale, Science, 255, 826- 830, 1992.

Moussa, R.T., Is the drainage network a fractal Sierpinski space?, Water Resour. Res., 33, 2399-2408, 1997.

Mueller, J.E., Re-evaluation of the relationship of master streams and drainage basins: Reply, Geol. Soc. Am. Bull., 8•, 3127-3130, 1973.

Mulder, T., and J.P.M. Syvitski, Climatic and morphologic relationships of rivers: Implications of sea-level fluctua- tions on river loads, J. Geol., 10•, 509-523, 1996.

Nash, D.B., Forms of bluffs degraded for different lengths of time in Emmet County, Michigan, USA, Earth Surf. Processes, 5, 331-345, 1980.

Newman, T.J., and A.J. Bray, Strong-coupling behavior in discrete Kardar-Parisi-Zhang equations, J. Phys. A, 29, 7917-7928, 1996.

Newman, W.I., and D.L. Turcotte, Cascade model for fluvial geomorphology, Geophys. J. Int., 100, 433-439, 1990.

Nikora, V.I., On self-similarity and self-affinity of drainage basins, Water Resour. Res., 30, 133-137, 1994.

Ormi• A.R., Temporal variability of a summer shorezone, in Space and Time in Geomorphology, edited by C. E. Thorn, pp. 285-313, Allen and Unwin, Winchester, Mass., 1982.

Ossadnik, P., Branch order and ramification analysis of large diffusion-limited-aggregation clusters, Phys. Rev. A, •5, 1058-1066, 1992.

Park, K., and B. Kahng, Exact derivation of the Kardar- Parisi-Zhang equation for the restricted solid-on-solid model, Phys. Rev. E, 51, 796-798, 1995.

Page 17: Self-organization and scaling relationships

PELLETIER: FRACTAL RIVER NETWORKS 7375

Patton, P.C., and S.A. Schumm, Gully erosion, Northwest- ern Colorado: A threshold phenomenon, Geology, 3, 88- 90, 1975.

Peckham, S.D., New results for self-similar trees with appli- cations to river networks, Water Resour. Res., 31, 1023- 1029, 1995.

Phillips, B.C., and A.J. Sutherland, Diffusion models ap- plied to channel degradation, J. Hydraul. Res., 25, 179- 191, 1986.

Pierce, K.L., and S.M. Colman, Effect of height and ori- entation (microclimate) on geomorphic degradation rates and processes, late-glacial terrace scarps in central Idaho, Geol. Soc. Am. Bull., 97, 869-885, 1986.

Plotnick, R.E., and K. Prestegaard, Fractal analysis of geo- logic time series, in Fractals in Geography, edited by L. De Cola and S. Lam, pp. 193-210, Prentice-Hall, Englewood Cliffs, N.J., 1993.

Rodriguez-Iturbe, I., E. Ijjasz-Vasquez, R.L. Bras, and D.G. Tarboton, Power law distributions of discharge mass and energy in river basins, Water Resour. Res., 28, 1089-1093, 1992.

Sadlet, P.M., Sediment accumulation rates and the com- pleteness of stratigraphic sections, J. Geol., 8g, 569-584, 1981.

Sayles, R.S., and T.R. Thomas, Surface topography as a non-stationary random process, Nature, 271, 43!-434, 1978.

Schmidt, K.-H., Regional variation of mechanical and chem- ical denudation, upper Colorado River basin, U.S.A., Earth Surf. Processes, 10, 497-508, 1985.

Schumm, S.A., Evolution of drainage systems and slopes on badlands at Perth Amboy, New Jersey, Geol. Soc. Am. Bull., 67, 597-696, 1956.

Seidl, M.A., W.E. Dietrich, and J.W. Kirchner, Longitu- dinal profile development into bedrock: An analysis of Hawaiian channels, J. Geol., 102, 457-474, 1994.

Slack, J.R., and J.M. Landwehr, Hydro-climatic data net- work: A U.S. Geological Survey streamflow data set for the United States for the study of climatic variations: 1874-1988, U.S. Geol. Surv. Open File Rep., 92-129, 1992.

Smith, T.R., and F.P. Bretherton, Stability and the conser- vation of mass in drainage basin evolution, Water Resour. Res., 8, 1506-1529, 1972.

Somfai, E., and L.M. Sander, Scaling and river networks - A Landau theory for erosion, Phys. Rev. E, 56, R5-R8, 1997.

Sornette, D., and Y.-C. Zhang, Non-linear Langevin model of geomorphic erosion processes, Geophys. J. Int., 113, 382-386, 1993.

Stark, C.P., An invasion percolation model of drainage net- work evolution, Nature, 352, 432-425, 1991.

Stark, C.P., Cluster growth modeling of plateau erosion, J. Geophys. Res., 99, 13957-13969, 1994.

Tarboton, D.G., R.L. Bras, and I. Rodriguez-Iturbe, The fractal nature of river networks, Water Resour. Res., 24, 1317-1322, 1988.

Tarboton, D.G., R.L. Bras, and I. Rodriguez-Iturbe, Scaling and elevation in river networks, Water Resour. Res., 25, 2037-2051, 1989.

Tokunaga, E., Ordering of divide segments and law of divide segment numbers, Trans. Jpn. GeomorThol. Union, 5, 71- 77, 1984.

Turcotte, D.L., A fractal interpretation of topography and geoid spectra on the earth, Moon, Venus, and Mars, J. Geophys. Res., 92, 597-601, 1987.

Turcotte, D.L., Fractals and Chaos in Geology and Geo- physics, Cambridge Univ. Press, New York, 1992.

U.S. Geological Survey, Digital elevation models, rev. ed., National Mapping Program technical instructions, Data users guide 5, Reston, Va., 1990.

U.S. Geological Survey, Data from selected U.S. Geological Survey National Stream Water-Quality Networks (WQN), Digital Data Series DDS-37, Reston, Va., 1996.

Vening Meinesz, F.A., A remarkable feature of the Earth's topography, Proc. K. Ned. Akad. Wet., Set. B, 5•, 212- 228, 1951.

Voss, R., Fractals in nature: From characterization to sim- ulation, in The Science of Fractal Images, edited by H. Peitgen and D. Saupe, pp. 22-70, Springer-Verlag, New York, 1988.

Wallace, R.E., Profiles and ages of young fault scarps, north- central Nevada, Geol. Soc. Am. Bull., 88, 1267-1281, 1977.

Weissel, J.K., L.F. Pratson, and A. Malinverno, The length- scaling properties of topography, J. Geophys. Res., 99, 13997-14012, 1994.

Willgoose, G., A statistic for testing the elevation charac- teristics of landscape evolution models, J. Geophys. Res., 99, 13987-13996, 1994.

Willgoose, G., R.L. Bras, and I. Rodriguez-Iturbe, A cou- pled channel network growth and hillslope evolution model, 1, Theory, Water Resour. Res., 27, 1671-1684, 1991.

Wolman, M.G., and R. Gerson, Relative scales of time and effectiveness of climate in watershed geomorphology, Earth Surf. Processes, 3, 189-208, 1978.

J. D. Pelletier, Division of Geological and Planetary Science, MC 150-21, California Institute of Technology, Pasadena, CA 91125. ([email protected])

(Received November 24, 1998; accepted December 4, 1998.)