You are on page 1of 13

142: Debris Flow

ARONNE ARMANINI, LUIGI FRACCAROLLO AND MICHELE LARCHER


CUDAM and Department of Civil and Environmental Engineering, University of Trento, Trento, Italy

The principal features of debris ow are described. One section is devoted to debris-ow triggering, accounting for the geomechanical criteria by Takahashi (1978) and for most recent hydrological distributed models. In a second section, the rheology of debris ows is described: the theory of dispersive stresses by Bagnold, the kinetic theories of dense gases, and some recent experimental observations about debris ows in equilibrium. A section is devoted to mathematical models and the nal section discusses debris ows countermeasures.

INTRODUCTION
Debris ow is a motion of widely sorted debris (from a few millimeters to some meters) inside a watery matrix or also in the presence of clayish mud. One of the most evident facts is the oatation of huge boulders on the surface of the debris ow. Taking up the original denition proposed by Takahashi (1981), debris ows are massive sediment transport phenomena that manifest themselves in mountain streams characterized by a steep slope, where the motion of the granular phase is induced directly by gravity. Here, the ratio between the liquid and the solid transport rates is relatively low, and can be zero in case of dry granular mixtures (e.g. dry landslides). On the contrary, in the case of ordinary sediment transport (bed load and suspended load), sediments are not driven directly by gravity, but by the hydrodynamic actions induced by the uid, and liquidsolid transport ratio is relatively high. There is a wide range of two-phase sediment-laden motions for which this distinction is not so sharp. Other situations outside simple schemes concern cases where, although the solid transport rate is in the debris-ow range, the nest fractions of the grain distribution change the rheology of the interstitial uid. In fact, the role of ne sediments, because of their size, and in some cases, of their electrochemical properties, is completely different from the coarser ones: they can mix with water, forming a homogeneous interstitial uid (slurry), typically characterized by high viscosity and in some cases by cohesive behavior. Its composition is strongly connected to the ow dynamics, making it very

difcult to dene and distinguish the role of the two phases. Further, complications take place when a third phase, constituted by air and other gases, is present inside and affects the owing mixture, such as in the stout front of advancing debris-ow waves. The picture presented above is mirrored by manifold classications found in literature, where massive phenomena involving the liquid and solid phase are reported with different terminology: for example, debris ows, mud ows, hyperconcentrated ows, and so on, any of them possibly being preceded by adjectives such as turbulent, laminar, and so on. Recently, the knowledge about granular ows has progressed a lot, and nowadays (2003) there is a wide literature regarding both the rheological features of these kinds of ows and the applicative aspects concerning design criteria for defence and protection works. As a rule, there are at least two fundamental elements that must be present for the triggering of a debris ow: an opportune succession of rain events and the availability of solid material. As better explained afterwards, the weather event has to present absolutely particular characteristics: generally it must be very intense, but in order to be able to move the bulk it must also be preceded by an event long enough to take the sediments to saturation. The concomitance of these happenings gives little predictability to debris-ow events that are relatively rare, but present a remarkable destructive power; they concern above all small basins and in particular alluvial fans that have often undergone a recent urbanization. Basically, they are nonstationary phenomena developing in particularly

Encyclopedia of Hydrological Sciences. Edited by M G Anderson. 2005 John Wiley & Sons, Ltd.

2174

OPEN-CHANNEL FLOW

short times. Because of their unexpected character, their destructive power is often undervalued: small creeks that are affected by very modest solid or liquid events for decades, are sometimes affected by debris ows of huge intensity. The recurrence of such events, in fact, is very difcult to determine; systematic observations show that the debris ows have return times of the order of 50100 years, a period of time during which the phenomenon is likely not to manifest at all. Unlike oods in water streams, in which some extreme events happen almost regularly every year, even though with a different intensity, for debris ows it is exactly the sudden and often unexpected triggering that makes the phenomenon insidious.

DEBRIS-FLOW TRIGGERING
The assessment of debris-ow risk is very difcult as three main factors participate in the triggering of debris ows: the rainfall intensity, the initial state of the basin soil moisture, and the presence of enough sediment. The analysis of many different events has shown that generally the debris ows manifest themselves after an extreme rainfall event following a rainfall of long duration. In a certain area, in fact, debris ows usually occur when the antecedent rainfall intensity exceeds a certain critical value. Yet, it is necessary that in a suitable previous period (of the order of 7 to 10 days) a congruous volume of rain has fallen, making at least part of the valley saturated. A more physically based approach consists in using a steady-state shallow subsurface ow model (e.g. TOPMODEL by Beven and Kirkby, 1979) to evaluate the saturated areas and their expansion during the storm. However, unlike landslides generated in hillslopes by subsurface ow, debris-ow initiation is mainly due to surface runoff and a complete rainfall-runoff model is required. If design of defence works is implied, a special kind of rainfall representation could be useful in the form of intensity-durationfrequency (IDF) curves that relate the rainfall intensity and duration to a prescribed return time. If just the peak discharge is required, some simplied models are available and are sufcient for this goal (Rigon et al., 2005). According to the theory presented below, once the runoff creates a ow depth greater than a suitable threshold, the slope collapses and the debris-ow run-out starts. This critical value depends obviously on the geometric, geologic, and morphologic characteristics of the debris deposit and of the sublayer on which it lies. This approach alone is, however, not sufcient to get the total of sediment mass moved. Besides the weather events, there must also be a deep-enough debris storage so that the frequency of debris ows does not coincide with that of intense rainfall events. Hampel (1968) distinguishes between watercourses with a rocky channel bed and those which ow on erodible alluvial deposits. The former require

a gradual debris accumulation between an event and the following ones and are temporarily stable after the event; the latter can be destabilized by an event and this can bring about a period of activity with rather high frequencies, followed by the return to a state of rest. The use of stochastic models to assess the contemporary presence of all the conditions required is presently (2003) being studied, but it is still not sufciently tested to be utilized in eld applications (Iida, 1999). A further possibility of debris-ow generation is obviously due to the dynamic action of landslide in their run-out from hillslopes into channels or the collapse of retention structures existing along the mainstream. The total quantity of debris that rests after the event is usually due both to the size of the source of material (especially if it derives from a landslide) and to the dynamics of the debris ow, which erodes the bed on which it is moving. Takahashi (1978) rst introduced the stability theory regarding loose materials for the study of debris-ow triggering. This theory refers to a noncohesive and uniform granular body and is based on the balance of forces acting in different imbibition conditions. The Takahashi instability condition is written as follows: tan tan C0 1 + C0 1 h0 + nD (1)

where is the inclination angle of the deposit, is the friction angle of the material, C0 is the volume concentration of the particles in a state of rest, h0 is the ow depth of the water owing over the deposit (in uniform-ow condition), = (s )/ is the relative density of the immersed material, D the mean diameter of the material, and n is a nondimensional coefcient of the order of the unit depending on the shape and distribution of the grains. Note that if h0 = 0, corresponding to a completely saturated body with no surface runoff, the stability condition becomes: tan tan C0 1 + C0 (2)

Once in movement, the body is soon transformed into a debris ow. In this situation, the dynamic friction can also move the material lying underneath the layer initially made unstable; this phenomenon is known as mass entrainment. Moreover, according to Takahashi (1978) the values of the ratio (h0 /nD) able to generate a real debris-ow range between 0 and 1.33. For values below 0, there is a partially dry deposit which, when it becomes unstable owing to big enough slopes, gives rise to a landslide. For values over 1.33, there is a movement more similar to bed load rather than to debris ow. In short, according to Takahashi, the

DEBRIS FLOW

2175

slopes for which there is debris-ow range between the two following extremes: tan C0 C0 tan tan 1 + C0 + 1.33 1 + C0 (3)

Body

Tail

Front

If, for example, for the above-cited parameters we assume the following values that are characteristic of stony, noncohesive material, C0 = 0.7, = 1.65, tan = 0.8, we obtain: 15 9 23 5 (4) Once the movement is triggered, the debris ows can ow even with slopes smaller than the limit on the right of relation (4). Anyway, with slopes lesser than 3 the debris ows do stop. Consequently, the basins that are more frequently affected by debris ows are those in which the presence of hillslopes or torrents with slopes comprised between the limits listed above are statistically more important. The analysis of the historical events in the different regions suggests that the debris-ow phenomena regard essentially small catchments, generally from 2 to 10 km2 . Other elements that are important for the debris-ow formation are the geology and the vegetative cover. It is evident that the presence of vegetation inhibits strongly the debris-ow formation; therefore, debris ows are more likely to form above the vegetation limit. After all, it is at higher elevations that there are hillslopes with a slope ranging between the limits cited. Other mechanisms generating debris ows are given by landslides depositing in the channel bed, or by the collapse of a natural dam which was formed temporarily in the channel bed because of the stop of vegetation transported by the current, or else by the collapse of some retention structures (in particular check dams) built along the mainstream of the river affected by an initial debris ow.

Figure 1 Scheme of the longitudinal section of a debris ow: in the central part (body) the motion is quasi-uniform

conditions. The study of debris-ow dynamics requires the knowledge of the interaction between particles and interstitial uid, between the contour, the particles, and the interstitial uid, and among the particles themselves. Therefore, the quantitative description of debris-ow dynamics is very complex. It is often convenient to consider the debris ow as a continuous medium, to which is assigned a suitable rheological law able to simulate the different interactions between particles, uid, and wall. In fact, a granular material subjected to deformation can determine different types of interactions among grains and, therefore, it can generate stresses by different mechanisms. Individual particles may interact with one another in rigid particle clusters, generating a network of contact forces through sustained rolling or sliding contacts, or by nearly instantaneous collisions, during which momentum is exchanged and energy is dissipated because of inelasticity and friction. The relative importance of these mechanisms may be used as the characteristic dening the various ow regimes (Savage, 1984). Basically, the main interactions can be listed as follows: deformation of the mean uid eld by the particles; collisions among the particles; friction among the particles during long contact periods; deformation of the turbulence structure generated by the wall.

DEBRIS-FLOW RHEOLOGY AND DYNAMICS


Once started, debris ows tend to assume a typical layout, formed by a round-shaped front in which the biggest boulders tend to accumulate, followed by a body in which the free surface is nearly parallel to the undisturbed debris bed. In this part, the motion can be assumed to be close to the uniform motion. The central part with uniform motion is followed by a tail, in which the ow depth becomes thinner and where the zone of erosion is exposed (see Figure 1). It is possible to study the behavior of the motion in the different parts of the debris ow separately. It is, above all, the part with quasi-uniform motion that becomes signicant, since the hypothesis of uniform ow makes it possible to give a detailed determination of kinematic and dynamic

Depending on particle concentration and hydrodynamic conditions, some of the mechanisms described above can prevail with respect to the others.
Theory of Dispersive Stresses

Theoretical and experimental ndings by Bagnold (1954), regarding the rheology of a gravity-free dispersion of large solid spheres in a Newtonian uid under shear, probably represent the most extensively cited research over the last 50 years among scientists involved in the study of debris ows, snow avalanches, and of the ow of granular materials in general. Bagnolds theory presents some conceptual limits that have been analyzed in detail in the recent literature, among others by Hunt et al. (2002). Yet, it has the merit of being simple and of being the rst

2176

OPEN-CHANNEL FLOW

one that indicated the physics of the phenomenon clearly; moreover, it has been successfully applied to debris ows, above all by Takahashi (1978). The theory is based on the observation that when particles collide among themselves or on the containment wall, collisions manifest themselves as an increase in pressure called dispersive pressure. According to Bagnold (1954), two extreme situations can occur. The rst one occurs in the presence of particles of small size being well dispersed within the interstitial uid: the debris ow assumes then a macroviscous behavior and the viscous stresses prevail with respect to the dispersive ones because of the intergranular collisions. On the contrary, an opposed situation occurs when the particles have high concentrations and the debris-ow speed is high, and then the collisions are much more frequent. In this regime, called grain-inertia, the collisions among grains are the determining factors as to the debrisow resistance and the effect due to the viscosity of the interstitial uid is negligible. Bagnold distinguished grain-inertia and macroviscous ow regimes on the basis of the nondimensional parameter Ba presented in equation (5), subsequently termed Bagnold number, representing the ratio between stresses due to inertia and those due to viscosity. Ba = D 2 (du/dz) 1/2 1 with = (C0 /C)1/3 1

in the macroviscous region : = 2.253/2 = tan

du dz (6)

and in the grain-inertia region : = 0.042 = tan

du D dz

cos (7)

(5)

It should be noted that both in the macroviscous and in the grain-inertia regime, the normal stress is proportional to the shear stress in the form = tan , where represents a dynamic friction angle, depending on collision conditions. Such behavior is reminiscent of the Coulomb criterion used to describe the stresses in cohesionless soils under conditions of limited equilibrium. According to Brown and Richards (1970), typical values for obtained during quasistatic yielding at low stress levels are close to the angle of repose, that is, about 24 for spherical glass beads and 38 for angular sand grains. Bagnold proposed a dynamic friction angle 37 for the macroviscous regime and = 18 for the grain-inertia regime. = As a completion of Bagnolds theory, nowadays the macroviscous regime is considered to be practically absent in debris ows, while where the shear rate and the velocity of the uid are small, a regime called frictional or quasistatic takes place, in which the contact among particles is quasi-permanent and the ratio between normal and shear stress can be considered roughly constant.
Kinetic Theories

where D is the grain diameter, is Bagnold linear concentration, C is the volume concentration of the solid fraction, C0 is the maximum concentration possible (concentration at rest), du/dz is the shear rate, and are the dynamic viscosity and the mass density of the interstitial uid. Bagnold called macroviscous the regime characterized by small Bagnold numbers (Ba < 40), where the shear stresses behave like a Newtonian uid with a viscosity corrected by the presence of the particles, and grain-inertia the other regime, is characterized by Ba > 450, where the stresses were independent of the uid viscosity and proportional to the square of the shear rate and to the square of the linear concentration . The intermediate range of Bagnold number (40 < Ba < 450) occupies a transitional region. Bagnold derived simple analyses to explain the rheological behavior in the two limiting regimes. In the viscositydominated macroviscous regime, shear and normal stresses are linear functions of the shear rate du/dz. Bagnold identied the presence of a normal stress in radial direction, termed dispersive pressure, and attributed it to a statistically preferred anisotropy in the spatial particle distributions. He proposed the following relations for the shear stress and the normal stress :

Within a granular ow, the velocity of each particle may be decomposed into the sum of a mean velocity and a random component, taking into account the relative motion of the particle compared to the time-averaged value. Ogawa (1978) introduced rst the concept of granular temperature Ts , where 3Ts is the mean square of particle velocity uctuations as expressed by equation (8), and Savage and Jeffrey (1981) made the rst attempt to make more substantial use of the ideas contained in the previous theoretical work that had dealt with dense gases, for example, Chapman and Cowling (1970). Ts = 1 2 u +v2+w2 3 (8)

In analogy with thermodynamic temperature, granular temperature plays similar roles in generating pressures and in governing the internal transport rates of mass, momentum, and energy. Granular temperature can be generated with two distinct mechanisms (Campbell, 1990). The rst, the so-called collisional temperature generation, is a by-product of interparticle collisions, in the sense that two colliding particles

DEBRIS FLOW

2177

will have resultant velocities depending not only on their initial velocities but also on the type of collision they experienced; therefore, they will contain apparently random velocity components. The second mode of temperature generation, the so-called streaming temperature generation, is itself a by-product of the random particle velocities. Following its random path, a particle moving parallel to the local velocity gradient will acquire an apparently random velocity that is proportional to the difference in mean velocity between its present location and the point of its last collision. It should be noted that in both mechanisms of granular temperature generation, the magnitude of the generated random velocities is proportional to the local velocity gradient. However, unlike the collisional temperature generation, the streaming mechanism can generate only the component of random velocity lying in the direction perpendicular to the mean velocity gradient, therefore the generated granular temperature will be anisotropic. Campbell (1990) describes how the physical similarity between rapid granular ows and kinetic-theory view of gases has led to a great deal of work on creating similar models for granular materials on the basis of the idea of deriving a set of continuum equations (typically mass, momentum, and energy conservation) entirely from microscopic models of individual particle interactions. All the models are based on the assumption that particles interact by instantaneous collisions, implying that only binary or two-particle collisions need to be considered. Particles are usually modeled in a simple way, ignoring surface friction or any other particle interactions tangential to the contact-point, and considering a constant coefcient of restitution to represent the energy dissipated by the impact normal to the point of contact between the particles, even though Lun and Savage (1986) and other researchers showed a strong dependence of the coefcient of restitution on the relative impact velocity. Furthermore, molecular chaos is generally assumed, this implying that the random velocities of particles are independently distributed. Jenkins and Hanes (1998) apply kinetic theories to a sheet ow in which particles are driven by turbulent uid and supported by their collisional interactions rather than by the velocity uctuations of the interstitial uid. Azanza et al. (1999) adapted kinetic theories to a channel ow in which the particles are driven by gravity. Armanini et al. (2005) accounted for the interstitial uid interaction by means of an added mass coefcient. The constitutive relation for the particle pressures is therefore: 1 + 2C 2(1 C) (9) The function g0 (C) describes the variation of the particle collision rate with concentration, and was derived by Carnahan and Starling (1969) from considerations about the = Cs 1 + r s (1 + 4Cg0 )Ts where r =

nearly geometric form of the virial series for nonattracting rigid spheres. (2 C) g0 (C) = (10) 2(1 C)3 Again, the constitutive relation for the particle shear stress is taken to be the one for a dense molecular gas, in the form: = 8D(1 + r/s )s C 2 g0 Ts 1/2 5 1/2 5 1+ 12 8Cg0
2

1+

du dz

(11)

The balance of particle uctuation energy is equal to that for the energy of the velocity uctuations of the molecules of a dense gas. For inelastic particles, the gradient of the vertical component Q of the uctuation energy ux, expressed by equation (13), is required to balance rst the net rate of uctuation energy production per unit of volume of the mixture, and secondly the rate of collisional dissipation d : du dQ (12) = d dz dz where the rst term (on the left-hand side) represents the energy diffusion, the second one the net rate of production (the rate of working of the particle shear stress through the mean shear rate) and the last one, d , is the rate of collisional dissipation. Particles are driven into collisions by the mean motion, creating uctuation energy, while the inelasticity of the collisions dissipates uctuation energy into real thermal energy. The constitutive relation for the ux of particle uctuation energy is taken to be the one for a dense molecular gas, in the form given by Chapman and Cowling (1970), while the rate of collisional dissipation per unit of volume may be calculated using the Maxwellian velocity distribution function, in the form given by Jenkins and Savage (1983): Q = 4 1 + 9 32 1+ 5 12Cg0
2

C 2 g0 r 1+ 0.5 s s d = 24(1 e)

DTs 1/2

dTs dz

(13) (14)

C 2 g0 s (1 + r/s )Ts3/2 D 0.5

Savage (1998) developed a theory for slow, dense ows of cohesionless granular materials for the case of planar deformations, employing the notion of granular temperature. The conservation equations for mass, momentum, and particle uctuation energy are employed. At

2178

OPEN-CHANNEL FLOW

Frictiona
Collisional
Frictional
Collisiona l

l
Fricti onal

Co

llisi

Frictiona

ona

(a)

(b)

(c)

(d) Channel slope Froude number

Increasing direction for flow slope and transport concentration Increasing direction for Froude number

Figure 2 Typology of the ows examined: (a) loose bed, immature (or oversaturated) debris ow; (b) loose bed, mature debris ow; (c) loose bed, plug (or undersaturated) debris ow; (d) rigid bed debris ow

low deformation rates, the apparent form of the constitutive behavior is similar to that of a liquid, in the sense that the actual viscosity decreases as the granular temperature augments, contrary to rapid granular ows, in which viscosity increases as granular temperature augments. Stresses are constituted by two parts: a rate-independent, dry friction contribution, and a rate-dependent viscous part, having a quadratic dependence on the shear rate, obtained from the high shear rate granular ow kinetic theories in the form of equations (9) and (11). The magnitude of the rate-independent contribution was chosen so that the sum of the two parts satised the overall momentum balance perpendicular to the ow direction. Savage and Jeffrey (1981) introduced a parameter R, involving the particle diameter D, the square root of the granular temperature and the shear rate, written in the form: D du (15) R = 1/2 Ts dz As discussed in Savage (1998), in both analyses and computer simulations, typically the parameter R is found to be of the order of the unit for granular ows, ranging from purely collisional to slow, predominantly frictional ows.
Experimental Analysis

In contrast with previous experimental studies dealing with solidliquid ows in annular shear cells, closed ducts, or nonrecirculatory chutes, the experimental analysis carried out in a steady uniform channel ow has shown that the dynamics of a free-surface debris ow constituted of sedimentable material is more articulated than what was predicted by Bagnold for gravity-less granular ows. Contrary to what was presupposed by Takahashi (1978) in adapting Bagnolds theory to debris ows, in the experiments it has been observed that a debris ow can present a series of layers one on top of the other, each governed by a different rheology. The experiments by Armanini et al. (2005) permit

to distinguish four main regimes of interest (Figure 2): (a) loose bed, immature debris ow; (b) loose bed, mature debris ow; (c) loose bed, plug debris ow; and (d) rigid bed ow. The rst three regimes (a) (c) exhibit loose-bed equilibrium conditions. The immature debris-ow regime (a) is characterized by the ow of a clear water layer over a uid-driven sheet of granular material supported by contacts with the stagnant bed. For mature debris ow (b), the entire moving layer is composed of a mixture of liquid and grains. In the plug debris ow (c), a partially emerged, quasi-static assembly of grains translates over a liquid-granular shear layer. The possibility for the owing mixture to have an underlying bed formed by the same constituents (movable bed ) leads to the formation of an equilibrium condition, since the slope of the bed is a dependent variable dynamically coupled with the ow. For nonequilibrium conditions, the mixture has to ow over a nonerodible bed, and in uniform-ow conditions, the free surface is parallel to the rigid bed of the ume; therefore the presence of the solid phase results somehow in a variable independent of the bed slope. In fact, in this case the solid discharge is smaller than the effective transport capacity of the current. The distinction between equilibrium and nonequilibrium conditions has of course a strong inuence on the rheological processes throughout the cross section of the uniform ow. In fact, considering the proles of the variables of interest in the stress formation (mean velocity, velocity uctuations, and solid concentration), one can observe that they change enormously from cases with or without equilibrium. The experimental results made it possible to create a picture of the ow rheology that is rather simple: throughout the ow depth there are layers dominated by collisions among grains, whereas the complementary domains are essentially frictional. When the two rheological mechanisms coexist, the separation between collisional and frictional layers is represented as a rather sharp interface, correspondent to Stokes numbers, dened by equation (16), being equal to 5 10. Large Stokes numbers characterize collisional regions, while small ones

DEBRIS FLOW

2179

characterize frictional layers. St = 1 s D 2 du/dz 18

ow depth h. 2 h (C + 1) g 5 D s a sin (18) Bagnold suggests to assume a = 0.042, while, on the basis of debris-ow laboratory data, Takahashi (1978) proposed a greater coefcient a = 0.35. This value corresponds in case of real debris ows (e.g. for D = 5 cm, C = 0.56, = 36 , = 1.65, h = 3 m, C0 = 0.70, = 13) to a Chezy coefcient DF = 11 m1/2 s1 . The debris-ow concentration, if assumed to be equal to that of incipient movement in saturated conditions, can be expressed by the following relation as a function of the channel slope and of the relative density of the material: with DF = C= tan (tan tan ) (19) U = DF h sin

(16)

In equation (16), is the kinematic viscosity of the interstitial uid and D(du/dz) represents the relative velocity between adjacent sheared rows of particles. Moreover, the experimental vertical proles of energy production, diffusion, and dissipation (Armanini et al., 2003) lead to the conclusion that in the collisional layers energy production balances roughly energy dissipation, so that in equation (12) the diffusive term can be neglected. Exploiting this assumption, normal and shear stress described by equations (9) and (11) depend on the square of the shear rate, exactly like in Bagnolds graininertia theory, even if the ratio between shear and normal stresses is perfectly constant only in Bagnolds conjectures.
Global Relations for Debris Flows in Equilibrium

According to the descriptions above, a succession of frictional and collisional layers, one on top of the other, constitutes debris ows. However, considering the debris ow throughout its depth, if the frictional layers, characterized by small mean velocities, give a negligible contribution to the total discharge of the ow, uniform-ow formulas can be obtained referring only to the collisional/graininertia layers. For this purpose, it can be assumed (Takahashi, 1978) that the ow is steady and uniform and that the tangential projection of the entire burden of the ow (water + grains) is charged over the solid phase only, therefore assuming that the shear stress in the liquid phase is negligibly small. Moreover, the vertical component of momentum balance for the particle phase requires that the gradient in the particle pressure balances the buoyant weight of a unit volume of particles. d (z) d = C g cos = (C dz dz + 1)g sin (17)

Takahashi (1991) proposed empirical formulas similar to equation (18) when the regime is clearly not dominated by collisional particle interactions. In case of viscous debris ow or mudows, the problem uid is often treated as a single-phase visco-plastic uid, characterized by a HerschelBulkley rheology (Coussot, 1997). This kind of approach is particularly suitable when the presence of silt and clay is relevant. For low solid concentrations, the ows behave like a Newtonian uid, while by increasing the concentration the mixture presents a yield stress c and a nonlinear relation between the shear stress and the shear rate, as expressed by equation (20): = c + k du dz
n

du dz

if > c (20)

= 0 if < c

Under these hypotheses and considering the concentration C to be constant (which corresponds to assume the dynamic friction angle to be constant), it is possible to integrate Bagnolds grain-inertia rheological equation (7) combined with equation (17) throughout the ow depth, thus obtaining a relation between the mean velocity and the slope of the current in uniform motion in the same form of Gaukler Strickler or Chezy formula. In this case, Chezy friction coefcient DF is a function of the solid fraction, of the sediment size, of the ratio between uid and solid density, of the friction angle of the material, and of the

where k and n are parameters depending on the type of uid and on its concentration, to be determined by means of laboratory tests and by the utilization of a rheometer. Like equation (7), equation (20) expresses a rheological link between shear stress and shear rate, and therefore it can be integrated throughout the ow depth in order to obtain the mean velocity value in uniform-ow conditions. U= n (n + 1)h h m g sin k
1/n

c m g sin

1+(1/n)

n c h 2n + 1 m g sin

(21)

where m represents the density of the homogeneous mixture. If we assume n = 1 and k = , relations (20) and (21) are referred to the case of a Bingham uid.

2180

OPEN-CHANNEL FLOW

Debris-ow Peak Discharge

Takahashi (1978) proposed a scheme for the estimation of debris-ow peak discharge based on the hypothesis that, after a short time, the debris-ow front assumes a self-similar prole. In this way, it is possible to make some assessments regarding the ratio between liquid and solid discharge. The debris-ow front is supposed to move with velocity Uf and also the tail goes downhill with uniform speed Ut , smaller than Uf ; in order that debris-ow prole becomes longer without becoming thicker. With reference to Figure 3, the balances of liquid and solid phase masses give: U0 h0 = Uf (1 C)h Ut (1 C)h Ut (1 C0 )sa + Ut h0 (22) Uf Ch = Ut Ch Ut C0 a (23)

discharge. Supposing that the front speed Uf coincides with the speed of the incoming water, then relation (24) becomes the following: C0 (25) QDF = Q0 C0 C where QDF is the debris-ow peak discharge and Q0 is the liquid peak discharge. As previously observed, in general, we can assume C0 0.65 for natural debris. = According to Takahashi (1978), for sufciently high slopes (i > 20 ), debris-ow concentration can be assumed to be equal to 90% of the maximum concentration. By doing so, equation (25) becomes QDF = 10Q0 . In the case of milder slopes (i < 20 ), debris-ow concentration is assumed to be equal to that of incipient movement in saturated conditions expressed by equation (19). In practice, debris-ow concentration ranges between the following limits: 0.3 < C < 0.9 1.43Q0 < QDF < 10Q0 C0 (26)

where h is the front depth, a is the depth of the excavation produced by the passing of the tail, s is the degree of saturation of the pores under the debris ow. The liquid feeding from upriver gives rise to a speed U0 and to a ow depth h0 and then to a discharge per length unit q0 = U0 h0 . From equation (23), we infer that the front speed must be lesser than the tail speed (Ut < Uf ), and then that the debris ow itself tends to become longer while going downhill. From equations (22) and (23), which admit that, as the debris ow passes by, the material in the channel bed is completely saturated (s = 1), we derive the following relation between the debris-ow discharge and the incoming liquid discharge: C0 Uf h = U0 h U0 h0 C0 C + C 1 a Uf (24)

Liquid peak discharge Q0 can be calculated with common hydrologic methods. An alternative method used to determine debris-ow peak discharge consists in assimilating the debris-ow front to that which forms after the collapse of a natural dam. The expression obtained is a function of the torrent depth hm , to be evaluated upstream of the dam, and of the torrent width B. 8 (27) Bhm ghm QDF = 27

MATHEMATICAL MODELING
Mathematical models are tools suitable to describe the general features of initiation, motion, and deposition of a gravity-driven mixture of debris and water and, lastly, the zoning of areas where damage is likely to occur. Owing to the high complexity that characterizes a real event, simplied assumptions in the mathematical modeling have to be introduced. At present in the applications the grain-size composition in the mixture is generally not accounted for, and only the solid concentration, which may change in time and space, remains to characterize the owing sediments. Therefore, the possibility to represent the variations of the sediment size throughout the debris ow (from its steep front, mainly constituted of boulders, to the tail) is generally very limited. General three-dimensional models require an extremely high computing time and level of hardware to determine the relevant solutions by means of numerical techniques, and are applicable, at present, to local situations only. The representations in a reduced two- or one-dimensional frame correspond to depth- or section-integrated models.

Equation (24), relative to the propagation of a quasistationary debris-ow front, can be utilized to derive the relation between debris-ow discharge and incoming liquid
Ut Uf
t + t a

t
a t Q0 h0

t
C h (C0, s)

Ut

Uf

Figure 3 Scheme debris-ow front

of

quasi-uniform

propagation

of

DEBRIS FLOW

2181

This averaging process requires hypotheses about the proles of the eld variables (velocities, pressure, concentration) through the ow depth. These assumptions make it possible to derive models where the ow domain is the basal topography. The characterization of the uid mixture takes into account the fact that the uid is not strictly homogeneous, but made up of different phases. The liquid phase refers to the interstitial slurry that is made up of water and of the nest grain fractions. The solid phase represents the coarser grain fractions. Some models admit that the two phases have different velocity distributions. The coupling between the two phases is present both in the mass conservations and in the momentum equations (due to the solidliquid stress interaction). When we apply the hypothesis that there is no relative velocity between the two phases, the resulting integrated models present only one momentum equation, referring to a uid having the bulk mass density. Since attention is focused on averaged equation models, the rheological properties of the owing mixture is lumped in the denition of algebraic expressions for the tangential stresses acting on the boundaries of the ow domain, especially on the bottom. Therefore, the boundary stresses are a function of the eld variables of the model (such as ow depth, velocity, slope, sediment size) and show up as source terms in the momentum equations. Mathematical and numerical models are thought of with proper boundary and initial conditions. Debris ows run-out and depositions over fans clearly depend on the incoming hydrograph. Initial conditions may also play a signicant role, as in cases where the ow is triggered by an abrupt collapse of barrages (dam breakinduced debris ows; Fraccarollo and Capart, 2002). Motion equations are based on conservation laws. Their number, along with the number of variables, depends on the renement of the model. The options mainly concern the number of phases and the number of momentum equations, one for each direction and phase-velocity eld. In case of a nonhomogeneous uid, the bed-level changes in time and the morphological evolutions are part of the solution. For sake of simplicity, one-dimensional models are referred to a ow section of unit width and rectangular shape, assuming no frictional inuences from the sidewalls. Modeling of a one-phase system in the one-dimensional framework is herein reported: t (h) + x (U h) = 0 2 (U h) + U 2 h + g h + gh zb = m t x x 2 (28) where h is the ow depth, U is the ow-depth-averaged velocity, m = (C + 1) is the density of the homogeneous mixture, zb is bed elevation, /m represents the only

source term, providing the shear stress at the bottom, is the momentum-ux coefcient that takes into account the nonuniform velocity distribution through the ow depth. In this context, the main point characterizing the uid is the formulation of the wall friction force, to be deduced from rheological assumptions. Many rheological models have been tested in the past by different authors. Most of them can be classied in the following categories: (i) models for muddy uid (e.g. Bingham models); (ii) models for mixtures in the inertial regime (e.g. equation 11); (iii) a combination of different dissipation models, including strain-rate independent contributions, to simulate grainto-grain static prolonged contact, and various strain-rate dependent ones, such as Newtonian, turbulent, or collisional behaviors. Under the same hypotheses, and moreover assuming = 1, the two-dimensional extent of the mass and momentum balances are: t (h) + x (hU ) + y (hV ) = 0 (hU ) + 1 gh2 + hU 2 + (hU V ) + gh zb t x 2 y x x = m (hV ) + (hU V ) + 1 gh2 + hV 2 + gh zb t x y 2 y y = m (29) where the inner shear stresses have been also neglected, as in most models of this kind, and where x and y are the components of the bottom shear stress and U and V are the components of the velocity. One of the answers expected from model applications for mapping the risk concerns the description of debris deposition phenomena at the alluvial fans. In case of a homogeneous uid, the stopping of the ow can be induced by a visco-plastic constitutive law (i.e. Bingham, HerschelBulkley models): when the yield stress balances or exceeds the local acting forces, the uid comes to a local stop throughout the ow depth. On the contrary, in situations where the debris deposition comes together with a reduction of the water content because of a velocity reduction of the ow over the alluvial fan, it is necessary to exploit models dealing with a nonhomogeneous two-phase uid. The presence of solid and liquid phases is accounted for by considering the volumetric solid concentration (depth averaged) in both mass and momentum conservations. With a full two-phase model, two mass and two momentum equations should be considered. However, assuming that the velocities of the two phases are correlated, it comes out that a single momentum equation (and a single velocity eld), associated to the bulk uid-density, remains. Here, it is chosen to represent the formally simpler case, assuming that there is no velocity-phase difference. The

2182

OPEN-CHANNEL FLOW

following model is obtained in the one-dimensional case: t (h + zb ) + x (U h) = 0 (Ch + C z ) + (CU h) = 0 t b b x [(C t + 1)U h] + x (C +(C + 1) U 2 h + kg h 2
2

laws devoted to the massive sediment transport under consideration. Under the same hypotheses, assuming, moreover, that the and k coefcients are equal to unity, the two-dimensional extension of the mass and momentum balances can be derived, in which, as in the case of homogeneous uid, the inner shear stresses can be neglected.

(30) where C is the depth-averaged solid concentration, is the water density, k is the active/passive coefcient employed in soil mechanics, determined by the depth-averaged ratio between the vertical and the longitudinal normal stresses (Savage and Hutter, 1991; Iverson, 1997); k is often assumed equal to one, as for pure uids. In two-phase models (equation 30) one more closureassumption, relevant to the volumetric solid concentration C, is needed with respect to the homogeneous uid model (equation 28). This relation can be derived from assuming a transport-capacity formulation, under the hypothesis that the processes of debris transport rapidly t the dynamic changes (hypothesis referred to as equilibrium). Regarding the transport-capacity formulation to be used, equation (19) is an example concerning granular debris ows. Many other empirical laws can be considered, spanning from renements of bed-load formulations often employed in torrential oods (Armanini, 2005), to more specic

+ 1)gh zb = x

MITIGATION AND RISK REDUCTION MEASURES


The mitigation remedial and defence techniques against debris ows are in their present form relatively recent tools: in the past, in fact, people simply tried not to build in the areas where there had been previous debrisow events. Only recently, as a consequence of the frantic development of the tourism settlements, risk areas, frequently located along the alluvial fans, have started to be urbanized. In general, the defence works against the debris ows can be classied in two categories: the active countermeasures and the passive countermeasures. The former consist basically of interventions aiming at reducing the risk of debris-ow triggering. Then they are meant to give stability to the debris deposits in the torrents beds or in the hillslopes. The latter instead are works built to defend directly the settlements or the zones subject to

Consolidation of sediment deposits

Consolidation of bed torrent Check dam and debris flow breakers Walls and embankments

Walls against slope failures

Over pass or tunnel Channel

Deposition basin

Figure 4 Scheme of the different types of active defence works against debris ows

DEBRIS FLOW

2183

debris-ow risk or the single structures (Figure 4). There are of course works responding to both criteria.
Consolidation of the Debris Deposits

The stabilization of the debris deposits is obtained either by reducing the slope or by hindering or reducing the possibility of saturation. This objective can be reached by consolidation works transversal to the deposits (endowed, where possible, with effective draining systems), by piling and anchors and, where possible, by vegetation. However, the bulkiest sources of the material mobilized by the debris-ow events are often located in places hardly reachable by the mechanical means; moreover, it is often counterproductive to move the deposits themselves in order to build drainages, as the movement of the material may increase its instability. As a result, the cases of applicability of these devices are often limited.
Consolidation of Riverbed and Lateral Slopes

In case the debris deposit affects directly the channel bed, the more limited the extent of the works is, the easier is the consolidation. The works constructed in these cases are sills or traditional closed check dams designed to reduce and to stabilize the riverbed slope and therefore to reduce the ow velocity. Often, in fact, the channel bed instability initially gives rise to an intense bed load that can degenerate into a debris ow owing to the erosion of the riverbanks or to the accumulations caused by the geometric variations of the sections (natural dam break). A system that augments the channel bed stability is represented by the chains of consolidation dams. The check dams built for this purpose are absolutely similar to the check dams used in case of solid ordinary transport. A phenomenon that characterizes the passing of the debris ow and that must be considered in the dimensioning of the check dam foundations is the amplication of ow depth from bed uidication that accompanies the passing of the debris ow (J ggi and a Pellandini, 1997).
Debris-ow Breakers

Figure 5 Slit check dam with a series of debris-ow breaker on Torrent Chieppena Trentino Italy (Provincia Autonoma di Trento, 2002). A color version of this image is available at http://www.mrw.interscience.wiley.com/ehs

More often, open check dams are built to intercept the solid material. However, debris ows usually occur in streams with a steep slope, where there is often lack of space where the material can be stored. In this case, the objective of the check dam is to diminish the ow velocity in order to reduce its destructive power in case of a dynamic impact. Therefore, these structures must resist above all the dynamic impact and they should also intercept the huge boulders. Various devices have been proposed for this purpose. The most common structures are slit check dams with

quite a large opening often protected by one or more stout buttresses that are dimensioned so that they can resist the dynamic impact (Figure 5). The hydrodynamic working of such structures is not clear and practical criteria suggested by experience are used for their design. The most effective results are obtained by water separation because the reduction of water concentration increases considerably the energy dissipation inside the ow and consequently its velocity. Yet it is important that the volume upstream of the check dam remains free after each event, therefore the opening must allow the removal of the material accumulated. Accordingly, it can be useful to dimension the opening, so that the excavators meant to remove the material accumulated can pass through it. The upstream side of the buttresses is inclined with respect to the vertical so as to reduce the dynamic impact of the debris ow that depends on the normal component of velocity. The inclination adopted varies between 45% and 60%. The overpressure p generated in the dynamic impact of the debris-ow front colliding against a vertical wall is obtained by applying the mass balance and the motion quantity balance, thus obtaining (Armanini and Scotton, 1993) the following expression: p = m U (U aw ) (31)

2184 OPEN-CHANNEL FLOW

where m is the debris-ow density, U is the mean velocity of the front and aw is the velocity of propagation of the gravitational wave reected by the wall. Very often relation (31) is corrected with an opportune coefcient p in order to consider possible secondary effects. The impact coefcient p often includes the effect of the reected wave: p = p s U 2 (32)

The coefcient p varies from 2 for quite slow and dense debris ows to 0.7 for faster and more liquid debris ows.
Debris Flows Articial Channels and Retention Basins

It has been already pointed out that often there is not sufcient space to intercept all the debris-ow volume upstream urbanized areas. Provided that debris-ow breakers and check dams have reduced the ow velocity and intercepted part of the largest boulders, it is necessary to canalize the torrent downstream the retaining works, often by building articial channels able to deviate the debris ow. These channels usually cross the villages and the alluvial fans where the space available is limited. Yet, it is useful to distinguish between a channel designed for the water discharge only and one affected by remarkable solid discharges or by debris ows. Often, it is the second necessity that is more urgent. Then the channel is supposed to resist to high velocities and strong tangential stresses, and must offer a very stable bed. Therefore, it is often necessary to resort to channels covered and strengthened with concrete. In this case, the covering surface is often smoothed to reduce the wall roughness. This is not the case of the channel for debris ow (Figure 6) where the resistance does not depend on surface roughness, but on particle collisions.

Bends with strong curvatures must be avoided in order to prevent depositions induced by secondary effects due to the supercritical nature of the ow, and in any case freesurface elevation must be properly accounted for, in order to avoid lateral debris-ow ooding. In other cases, it can be useful to deviate the debris ow on the side to safeguard some sites. This deection is obtained by the same techniques used for the snow avalanches, that is, by diverting walls or dikes. The topography of the torrent often does not offer spaces granting a sufcient volume for the deposition of the solid material. In this case, it is necessary to create such storages. The debris retention basins are created through lateral training dikes, protected downstream by a check dam, which, if necessary, is inserted into the body of an articial banking. To optimize the volume available, it is necessary to remember that the widening cone of an overcritical current depends on the Froude number of the incoming current.
Acknowledgment

The authors thank Prof. Riccardo Rigon for his help in preparing the section about the triggering of debris ow.

REFERENCES
Armanini A. (2005) Mountain streams. Encyclopedia of Hydrological Sciences, John Wiley and Sons. Armanini A., Capart H., Fraccarollo L. and Larcher M. (2005) Rheological stratication in experimental free-surface ows of granular-liquid mixtures. Journal of Fluid Mechanics, 532, 269 319. Armanini A., Fraccarollo L. and Larcher M. (2003) Dynamics and energy balances in uniform liquid-granular ows, Proceedings of FLOWS 2003, International Workshop on Occurrence and Mechanisms of Flows in Natural Slopes and Earthlls, Sorrento, 14 16 May 2003, L. Picarelli editor. Patron editore: pp. 131 137. Armanini A. and Scotton P. (1993) On the dynamic impact of a debris ow on structures. Proceedings of the XXV IAHR Congress, Tokyo, vol. B, paper no. 1221. Azanza E., Chevoir F. and Moucheront P. (1999) Experimental study of collisional granular ows down an inclined plane. Journal of Fluid Mechanics, 400, 199 227. Bagnold R.A. (1954) Experiments on a gravity-free dispersion of large solid spheres in a Newtonian uid under shear. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 225, 49 63. Beven K. and Kirkby M.J. (1979) A physically-based, variable contributing area model of basin hydrology. Hydrological Sciences Bulletin, 24, 43 69. Brown R.L. and Richards J.C. (1970) Principles of Powder Mechanics, Pergamon Press: London. Campbell C.S. (1990) Rapid granular ows. Annual Review of Fluid Mechanics, 22, 57 92.

Figure 6 Articial channel on Torrent Gola Trentino Italy (Provincia Autonoma di Trento, 1991). A color version of this image is available at http://www.mrw.interscience. wiley.com/ehs

DEBRIS FLOW

2185

Carnahan N.F. and Starling K. (1969) Equations of state for non-attracting rigid spheres. Journal of Chemical Physics, 51, 635 636. Chapman S. and Cowling T.G. (1970) The Mathematical Theory of Non-Uniform Gases, Third Edition, Cambridge University Press. Coussot P. (1997) Mudow Rheology and Dynamics, A. A. Balkema: Rotterdam. Fraccarollo L. and Capart H. (2002) Riemann wave description of erosional dam-break ows. Journal of Fluid Mechanics, 461, 183 228. Hampel R. (1968) Geschiebeablagerungen in Wildb chen, a Dargestellt am Modellversuchen Teil 1 und 2 . Wildbach- und Lawinenverbau, Nr. 1 2, 32nd Year. Hunt M.L., Zenit R., Campbell C.S. and Brennen C.E. (2002) Revisiting the 1954 suspension experiments of R. A. Bagnold. Journal of Fluid Mechanics, 452, 1 24. Iida T.A. (1999) Stochastic hydro-geomorphological model for shallow landsliding due to rainstorm. Catena, 34, 293 313. Iverson R.M. (1997) The physics of debris ows. Reviews of Geophysics, 35(3), 245 296. J ggi M.N.R. and Pellandini S. (1997) Torrent Check Dams a as a Control Measure for Debris Flows. Lecture Notes on Earth Sciences, Vol. 64, Recent Developments of Debris Flows, Armanini A. and Michiue M. (Eds.), Springer-Verlag: pp. 186 205. Jenkins J.T. and Hanes D.M. (1998) Collisional sheet-ow of sediment driven by a turbulent uid. Journal of Fluid Mechanics, 370, 29 52. Jenkins J.T. and Savage S.B. (1983) A theory for the rapid ow of identical, smooth, nearly elastic particles. Journal of Fluid Mechanics, 130, 187 207.

Lun C.K. and Savage S.B. (1986) The effects of an impact dependent coefcient of restitution on stresses developed by sheared granular materials. Acta Mechanica, 63, 15 44. Ogawa S. (1978) Multitemperature theory of granular materials. Proceedings of the US-Japan Seminar on Continuum Mechanics and Statistical Approaches to Mechanics of Granular Mater, Gukujutsu Bunken Fukyukai: Tokyo, pp. 208 217. Provincia Autonoma di Trento (1991) Per Una Difesa Del Territorio, Edizioni Arco Trento. Provincia Autonoma di Trento (2002) http://www.sistemazio
nemontana.provincia.tn.it/

Rigon R., DOdorico P. and Bertoldi G. (2005) The peak ows and their geomorphic structure. Submitted to Water Resources Research. Savage S.B. (1984) The Mechanics of Rapid Granular Flows. Advances in Applied Mechanics, 24, 289 366. Savage S.B. (1998) Analyses of slow high-concentration ows of granular materials. Journal of Fluid Mechanics, 377, 1 26. Savage S.B. and Hutter K. (1991) The dynamics of avalanches of granular materials from initiation to runout. Part I: analysis. Acta Mechanica, 86, 201 223. Savage S.B. and Jeffrey D.J. (1981) The stress tensor in a granular ow at high shear rates. Journal of Fluid Mechanics, 110, 255 272. Takahashi T. (1978) Mechanical characteristics of debris ow. Journal of Hydraulics Division, ASCE, 104(8), 1153 1169. Takahashi T. (1981) Debris ow. Annual Review of Fluid Mechanics, 13, 57 77. Takahashi T. (1991) Debris Flow, IAHR Monograph, A. A. Balkema: Rotterdam.

You might also like