Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-20T09:57:20.970Z Has data issue: false hasContentIssue false

Slow drag in wet-snow avalanche flow

Published online by Cambridge University Press:  08 September 2017

B. Sovilla
Affiliation:
Avalanches, Debris Flows and Rockfall Research Unit, WSL Institute for Snow and Avalanche Research SLF, Flüelastrasse 11, CH-7260 Davos Dorf, Switzerland E-mail: sovilla@slf.ch
M. Kern
Affiliation:
Department of Snow and Avalanches, BFW Institute for Natural Hazards and Alpine Timberline, Hofburg-Rennweg 1, A-6020 Innsbruck, Austria
M. Schaer
Affiliation:
Warning and Prevention Research Unit, WSL Institute for Snow and Avalanche Research SLF, Flüelastrasse 11, CH-7260 Davos Dorf, Switzerland
Rights & Permissions [Opens in a new window]

Abstract

We report impact pressures exerted by three wet-snow avalanches on a pylon equipped with piezoelectric load cells. These pressures were considerably higher than those predicted by conventional avalanche engineering guidelines. The time-averaged pressure linearly increased with the immersion depth of the load cells and it was about eight times larger than the hydrostatic snow pressure. At the same immersion depth, the pressures were very similar for all three avalanches and no dependency between avalanche velocity and pressure was apparent. The pressure time series were characterized by large fluctuations. For all immersion depths and for all avalanches, the standard deviations of the fluctuations were, on average, about 20% of the mean value. We compare our observations with results of slow-drag granular experiments, where similar behavior has been explained by formation and destruction of chain structures due to jamming of granular material around the pylon, and we propose the same mechanism as a possible microscale interpretation of our observations.

Type
Research Article
Copyright
Copyright © International Glaciological Society 2010

1. Introduction

In avalanche engineering, the pressure, p, exerted on an obstacle by an extreme avalanche is assumed to be p = c f ρU 2, i.e. p is proportional to the square of the impact velocity, U, and to the avalanche bulk density, ρ. Shape and rheology effects are taken into account by a drag coefficient, c f. This formulation is also used by the Swiss Guidelines (Reference Salm, Burkard and GublerSalm and others, 1990) and applies to extreme dry avalanches where impact forces are largely correlated with the avalanche kinetic energy (Reference Sovilla, Schaer, Kern and BarteltSovilla and others, 2008a). In practice, as there is no established alternative, the same formula is also used by practitioners for slow, wet-snow avalanches.

Recent measurements indicate that, for avalanches with low Froude numbers, this formula may heavily underestimate the impact pressures (Reference NoremNorem, 1991; Reference GauerGauer and others, 2007; Reference Baroudi and ThibertBaroudi and Thibert, 2009). However, no satisfactory explanation of these discrepancies in terms of possible underlying processes has been given so far.

Reference Sovilla, Schaer, Kern and BarteltSovilla and others (2008a) suggested that the impact pressure of avalanches in the subcritical flow regime may be governed by other processes than those involved in the supercritical flow regime. They reported a preliminary analysis of pressure and velocity data from the Vallée de la Sionne (canton Valais, Switzerland) test site, which shows that the measured impact pressures did not have a significant velocity dependency for Froude numbers less than 1, i.e. for subcritical flows. Furthermore, they observed that the total drag on the pylon strongly depends on the avalanche flow depth. Finally, Reference Sovilla, Schaer, Kern and BarteltSovilla and others (2008a) observed that the amplitude of pressure fluctuations in wet–dense avalanches increases with flow depth, in contrast to pressure fluctuations in dry–dense avalanches, which are larger close to the avalanche surface and smaller lower down in the avalanche body.

Our measurements of drag forces in slow wet avalanche flow may be consistent with theories for granular systems which predict that drag in slow flow varies linearly with flow depth and is independent of fluid velocity (Reference WieghardtWieghardt, 1975; Reference Albert, Pfeifer, Barabási and SchifferAlbert and others, 1999, Reference Albert2000; Reference Chehata, Zenit and WassgrenChehata and others, 2003). Thus, the calculation of avalanche slow drag on structures may require a different analytical approach than the avalanche dynamics models currently used.

In this paper, we present additional data of slow drag in avalanche flow and use new analysis methods to improve the preliminary analysis of Reference Sovilla, Schaer, Kern and BarteltSovilla and others (2008a). A comparison of our results with slow granular flow experiments provides further evidence of the granular flow behavior of wet-snow avalanches.

2. Instrumentation and Methods

We present data from the full-scale avalanche experimental test site Vallée de la Sionne, which is described in detail by Reference Sovilla, Schaer, Kern and BarteltSovilla and others (2008a). Here we analyze large, wet-snow avalanches that started spontaneously in the release zone at ∼2700 m a.s.l. and followed a track of ∼1500 m length down to a zone with artificial obstacles at 1600 m a.s.l. (Fig. 1a). Several obstacles equipped with instruments were installed on a large open slope with an average inclination of about 21° (Reference Sovilla, Schaer and RammerSovilla and others, 2008b). At this location, large wet avalanches have a maximum flow depth of 3.5–5.5 m and flow velocities of 1–10 m s−1 (Table 1). As there is no lateral confinement, the avalanches may expand to a width of up to 100 m. They finally run out in the wide open slope below the obstacle zone, or reach the bottom of the valley and stop at the slope on the other side of the valley. One of the obstacles is an oval-shaped, 20 m high pylon, equipped with pressure and velocity sensors (Fig. 1b).

Fig. 1. (a) The avalanche slope at the Vallée de la Sionne test site. A wet avalanche has just hit the obstacle zone. (b) Pylon and sensor installation details (p is pressure, v is velocity and d is density).

Table 1. Summary of the wet avalanche events measured at the Vallée de la Sionne test site during the winter seasons 2003/04 and 2006/07

Piezoelectric load cells are mounted on the pylon at 0.5–5.5 m above the ground with 1 m vertical spacing (Reference Schaer and IsslerSchaer and Issler, 2001). During winter 2003/04, sensor diameters were 0.1 and 0.25 m. Since winter 2004/05, only sensors with diameters of 0.10 m have been used. Pressures are recorded with a sampling frequency of 7.5 kHz.

Flow velocities are recorded by optoelectronic sensors mounted flush on wedges, with 1 m vertical spacing, up to 6 m above ground. The wedges minimize flow disturbances around the velocity sensors. The velocity is calculated from the cross-correlation between the signals of two reflectivity sensors with 29 mm streamwise spacing. The sampling frequency is 20 kHz. For technical details on the optical velocity measurements and data analysis procedures, see Reference Dent, Burrell, Schmidt, Louge, Adams and JazbutisDent and others (1998) and Reference Kern, Bartelt, Sovilla and BuserKern and others (2009).

We estimated the height of the surface of the flowing avalanche, h(t), at the pylon by using the pressure time series, p(zi , t), from the load cells at different heights, zi , where i indicates the load cells (Fig. 1b). We calculated the centered floating mean, 〈(p(zi , t)〉, of the pressure time series using a time-averaging window of T = 1 s. We then defined h(t), the assumed avalanche surface, as the height where the linear interpolation of the two uppermost pressure measurements reached 5 kPa.

3. Data Analysis

The three avalanches analyzed had very similar features. Rather than describe all the raw data, we concentrate here on avalanche 8448 (1 March 2007), but we also refer to the other avalanches in our analysis. Table 1 contains some of the avalanche parameters. For a more detailed description of the avalanche characteristics see Reference Sovilla, Schaer, Kern and BarteltSovilla and others (2008a) and Reference Kern, Bartelt, Sovilla and BuserKern and others (2009).

A plot of the raw pressure time series, p(zi , t), of avalanche 8448 is given in Figure 2a. Figure 2b shows the time-averaged pressure, 〈p(zi , t)〉, for the same avalanche. Figure 2c gives the corresponding fluctuations, which we characterized by the standard deviation, σp (zi , t), of the pressure time series:

A real avalanche rarely flows in a pure wet- or dry-flow regime. Most of the time, both regimes are present. Typically, in wet avalanches, a plug flow core may be surrounded by dilute flow, particularly if the avalanche was released from altitudes where the snow is still dry, and the snow in the track then becomes wet at lower altitudes.

Fig. 2. Avalanche 8448. (a) Raw pressure data; (b) the average pressure, 〈p〉; (c) fluctuations described by the standard deviation, σp ; and (d) the scaling function, σp /〈p〉. Points not scaling are indicated in gray in (d) and with large dots in (b).

We consider only measurements belonging to the slow-drag avalanche flow regime and therefore we define below a criterion based on the ratio between fluctuations and average pressure, i.e. on the relative fluctuations. This criterion is based on the fact that pressure fluctuations differ strongly between dry–dense and wet–dense flow (Reference Sovilla, Schaer, Kern and BarteltSovilla and others, 2008a).

Figure 2d shows the relative fluctuations of avalanche 8448, σp /〈p〉, where σp stands for σp (zi ,t) and 〈p〉 for〈p(zi , t)〉.

Figure 3 shows a statistical distribution of the relative fluctuations, σp /〈p〉, for each sensor and for each avalanche. On average, σp /〈p〉 is 0.21, with 95% of points being distributed in the interval [0.05, 0.4]. This scaling ratio persists for almost all measurements, at all immersion depths and for all avalanches. In Figure 2d, those points with a scaling ratio lower than 0.05 or higher than 0.4 are marked in grey. In Figure 2b, the same points are marked with large dots. Most of the outliers can be related to measurements from the avalanche surface, the avalanche front or the avalanche tail. These points probably originate from a more dilute layer surrounding the avalanche dense core and are excluded from further analysis.

Fig. 3. Statistics of fluctuations and average pressures ratio, σp /〈p〉, for each sensor of avalanches 8448, 6231 and 6241. Box plots show mean (symbol in box), median (line in box), 25/75% quantiles (box), 5/95% quantiles (whiskers) and 0/100% quantiles (cross).

Figure 4a shows pressure measurements that scale with the corresponding fluctuations (according to Fig. 2d) as a function of flow depth for avalanche 8448. Figure 4b shows a merging of all avalanche pressure profiles.

Fig. 4. (a) Mean pressures and fluctuations (pressure standard deviations) as a function of flow depth for avalanche 8448. Gray dots show measurements that were excluded from the analysis because σp /〈p〉 did not scale. (b) Pressure and fluctuation standard deviations as a function of flow depth for avalanches 8448, 6236 and 6241. Fitting lines for p = ζρgH are inserted.

Figure 5 shows time-averaged velocity profiles measured in the bulk of avalanches 8448, 6236 and 6241. The profile shapes indicate low-shear or even non-shear flow behavior (plug flow).

Fig. 5. Velocity profiles and shear rates, , extracted from the bulk of avalanches 8448, 6236 and 6241. Time periods, 620 s≤ t ≤660 s for avalanche 8448, 530 s≤ t ≤535 s for avalanche 6236, and 205 s≤ t ≤215 s for avalanche 6241 are shown.

4. Discussion

Current avalanche engineering design rules cannot predict correctly the forces exerted by wet-snow avalanches. They predict far smaller forces than those we measured (Reference Sovilla, Schaer, Kern and BarteltSovilla and others, 2008a) and do not describe the depth dependency we found in the experimental data (Fig. 4a and b). Moreover, the scaling between pressure fluctuations, σp , and mean pressure, 〈p〉, (Fig. 2d) cannot be explained by either fluid-dynamic or particle-impact models.

In fact, our experimental data may be better understood using a slow-drag granular model. Reference Geng and BehringerGeng and Behringer (2005) reported the scaling between fluctuations and average pressure as typical for slow drag in granular experiments. In these experiments, the drag force experienced by an object moving through a granular medium has been shown to originate from an apparently random distribution of force chains that depart from the object and disperse into the bulk of the granular medium. The drag fluctuations that typically occur in these experiments arise from the formation and rupture of these chains.

The investigation described in this paper, however, clearly differs from slow-drag granular experiments in that the latter typically use granular material with monodisperse elastic disks while, in our case, avalanches are characterized by a polydisperse granular mixture of plastic, compressible and breakable snow clusters.

Furthermore, slow-drag granular experiments are performed in small containers where boundary conditions normally play an important role (Reference WieghardtWieghardt, 1975; Reference Albert, Pfeifer, Barabási and SchifferAlbert and others, 1999; Reference Geng and BehringerGeng and Behringer, 2005), and the force chains that form during these experiments are typically confined in direction and length. In contrast, our experiments are not influenced by lateral confinement, as the pylon on the study site is on an open slope where an avalanche is normally several hundred meters long, up to 100 m wide and the flow depth can be as high as 5 m.

Nevertheless, none of these differences seems, in principle, to be an obstacle to the formation of force chains. In fact, a polydisperse granular mixture allows for a denser particle packing, and may also favor chain branching and increase the stability of the force chains, even when they are deformed. The absence of lateral confinement may allow the stress chains to develop in any direction, and to cover large areas. Furthermore, small shear rates, and thus small relative displacements between particles, may promote the formation of three-dimensional chains. Our wet avalanches preferably move in a plug-flow regime and are thus characterized by a small shear rate in the core (Reference Sovilla, Schaer, Kern and BarteltSovilla and others, 2008a; Reference Kern, Bartelt, Sovilla and BuserKern and others, 2009). In the bulk of avalanches 8448, 6236 and 6241 (Fig. 5), shear rates, , are 0.24, 0.01 and −0.16 s−1. Thus, from a microscopic point of view, fluctuations in our pressure measurements can be interpreted as an indication of the formation and rupture of stress chains due to snow jamming around the pylon.

At the macroscopic scale, slow-drag laboratory experiments with monodisperse granular material showed that the average drag force on a cylinder is F D = ηρgd c H 2, where H is the immersion depth of the cylinder, the dimensionless parameter, η, characterizes the grain properties, d c is the cylinder diameter, ρ is the flow bulk density and g is the gravitational acceleration (Reference WieghardtWieghardt, 1975; Reference Albert, Pfeifer, Barabási and SchifferAlbert and others, 1999). Our data give similar results to those found by Reference WieghardtWieghardt (1975) and Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999). The straight lines in Figure 4a and b show linear fits of the form p = ζρgH, where ζ is the fitting parameter.

From permittivity measurements performed inside avalanche 8448, 3 m above ground, we estimated the flow density to be p = 400 ± 80 kg m−3 (Reference Louge, Steiner, Keast, Decker, Dent and SchneebeliLouge and others, 1997). We used this value to indicate the average density for all avalanches. The best fits resulted for ζ = 7.2 ± 2.1, 8.1 ± 1.6 and 7.6 ± 1.7, for avalanches 8448, 6236 and 6241, respectively. The linear relation, p = ζρgH, is compatible with the drag force relation, F D = ηρgd c H 2, found by Reference WieghardtWieghardt (1975) and Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999). Indeed, Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999) observed that in general F D was linearly proportional to d c. But they also found an important deviation from linearity, when the particle diameter, d g, was large compared with d c.

Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999) explained the increased drag for low d c/d g by the fact that even when d c → 0, the cylinder must push grains aside as it moves relative to the medium and must therefore approach a threshold value of F D for dc → 0. Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999) reported this phenomenon from experiments in the approximative range 0.6 < d c/d g <3. In our experiments, assuming d c = 0.60 m and d g varying in the range 0.07–0.12 m (from granulometry observations in wet avalanche deposits by Reference Bartelt and McArdellBartelt and McArdell (2009)), we obtain 5 < d c/d g < 8. Therefore, since d c/d g for the avalanches we analyzed is in the same order of magnitude as found by Reference Albert, Pfeifer, Barabási and SchifferAlbert and others (1999), we expect that increasing the pylon diameter will result in a lower coefficient ζ.

Another striking characteristic of our experimental data is shown in Figure 4b where all avalanche pressure profiles are merged: the average pressures measured at a given immersion depth are in the same range for all avalanches. In the framework of slow-drag granular mechanics, this can be explained by assuming that the three avalanches had a similar granulometry and density. Slow-drag experiments performed by Reference Geng and BehringerGeng and Behringer (2005) showed that density strongly influences the avalanche drag. For densities above a minimum threshold value, force chains evolve and average pressure increases rapidly with density (Reference Geng and BehringerGeng and Behringer, 2005). Thus, if the investigated avalanches had a different flow density, this should have led to varying drag. The assumption that the wet avalanches we analyzed had a similar density is also in agreement with recent granulo-metric investigations of avalanche deposits that show a very similar particle size distribution for a variety of wet and dry avalanches. In particular, Reference Bartelt and McArdellBartelt and McArdell (2009) analyzed wet-snow avalanche deposits and observed a log-normal granular size distribution with a median particle size of approximately 0.07–0.12 m. Snow boulders of >0.5 m diameter were found on the avalanche surface, and the smallest grain size was in the order of 0.02 m.

Finally, we analyzed the influence of the avalanche velocities, v, which for the avalanches we studied (Table 1) were in the range 1–10 m s−1. No apparent velocity dependency can be seen in the merged pressure profiles of Figure 4b. Such a dependency would be immediately visible, as it would corrupt the fit of the profiles. Figure 6 shows the dependency of the mean pressure, 〈p〉, of avalanche 8448 on its velocity. The velocity dependency is very weak and almost non-perceptible. This weak dependency has also been reported in slow-drag granular experiments (Reference WieghardtWieghardt, 1975; Reference Albert, Pfeifer, Barabási and SchifferAlbert and others, 1999; Reference Geng and BehringerGeng and Behringer, 2005).

Fig. 6. Pressure as a function of velocity for avalanche 8448. We indicate measurements from the same range of immersion depth by color.

5. Conclusions

Analyzing pressure signals from wet-snow avalanches, which are characterized by both plug- and dilute-flow regimes, we found that relative fluctuations of the signal are a good criterion for distinguishing between the two regimes. We observed that, for wet-snow avalanches moving in a plug-flow regime, the impact pressure on a pylon increases linearly with flow depth and the pressure is independent of the avalanche velocity. Furthermore, the measured average impact pressure is about eight times larger than the hydrostatic snow pressure. Even if we account for the effects of the small pylon width, average pressures are up to an order of magnitude larger than those predicted by standard engineering rules (Reference Sovilla, Schaer, Kern and BarteltSovilla and others, 2008a).

It is possible that, so far, the drag forces arising from slow avalanche flow have been considerably underestimated since such avalanches have usually been assumed to have fluid-like behavior and to involve viscous and collisional particle interaction as underlying processes (Reference Salm, Burkard and GublerSalm and others, 1990). Thus, our results have direct relevance for estimates of drag forces in design procedures for avalanche-prone structures.

A sound theoretical basis for calculating the force-chainrelated pressures on obstacles is missing, but simple fit templates (e.g. Reference Albert, Pfeifer, Barabási and SchifferAlbert and others, 1999) may still help to predict the drag exerted by a slow-moving avalanche on a narrow obstacle better than current avalanche dynamics formulas do. Alternatively, Coulomb’s theory for passive failure in a Mohr–Coulomb material may also explain the linear dependency of pressure with flow depth and may be a future solution for the calculation of slow drag in avalanche flow (Reference WieghardtWieghardt, 1975). Better predictions should help to prevent incorrect dimensioning of avalanche protection measures which might have fatal consequences.

We also found that the avalanche pressure fluctuations scaled to the average pressure and the average pressures measured at a given immersion depth are in the same range for all avalanches. These features can be explained in the framework of slow drag in granular media. Accordingly, we interpret the pressure fluctuations as indicating the formation and rupture of such force chains.

With our current level of knowledge, we are far from a proof of the existence of force chains in dense avalanche flow and from a sound theoretical description of their formation and geometry. Nevertheless, we have been able to provide various indications that the force chains observed in granular experiments may also be present in natural gravity flows of polydisperse inelastic granular mixtures, such as wet-snow avalanches, and that these can also be studied in situations where their flow is not laterally confined.

Acknowledgements

Part of this work was funded by the Swiss National Foundation under grant No. 206021-113069/1. We thank F. Dufour for organizing the field campaigns, and canton Valais, Switzerland, for providing valuable financial support. Special thanks are due to the mechanical and electronic staff of SLF for all their invaluable help.

References

Albert, I. and 8 others. 2000. Jamming and fluctuations in granular drag. Phys. Rev. Lett., 84(22), 51225125.Google Scholar
Albert, R., Pfeifer, M.A., Barabási, A.-L. and Schiffer, P.. 1999. Slow drag in a granular medium. Phys. Rev. Lett., 82(1), 205208.Google Scholar
Baroudi, D. and Thibert, E.. 2009. An instrumented structure to measure avalanche impact pressure: error analysis from Monte Carlo simulations. Cold Reg. Sci. Technol., 59(2–3), 242250.CrossRefGoogle Scholar
Bartelt, P. and McArdell, B.W.. 2009. Granulometric investigations of snow avalanches. J. Glaciol., 55(193), 829833.Google Scholar
Chehata, D., Zenit, R. and Wassgren, C.R.. 2003. Dense granular flow around an immersed cylinder. Phys. Fluids, 15(6), 16221631.Google Scholar
Dent, J.D., Burrell, K.J., Schmidt, D.S., Louge, M.Y., Adams, E.E. and Jazbutis, T.G.. 1998. Density, velocity and friction measurements in a dry-snow avalanche. Ann. Glaciol., 26, 247252.CrossRefGoogle Scholar
Gauer, P. and 7 others. 2007. On full-scale avalanche measurements at the Ryggfonn test site, Norway. Cold Reg. Sci. Technol., 49(1), 3953.Google Scholar
Geng, J. and Behringer, R.P.. 2005. Slow drag in two-dimensional granular media. Phys. Rev. E, 71(1), 011302. (10.1103/PhysRevE.71.011302.)Google Scholar
Kern, M., Bartelt, P., Sovilla, B. and Buser, O.. 2009. Measured shear rates in large dry and wet snow avalanches. J. Glaciol., 55(190), 327338.CrossRefGoogle Scholar
Louge, M.Y., Steiner, R., Keast, S.C., Decker, R., Dent, J.D. and Schneebeli, M.. 1997. Application of capacitance instrumentation to the measurement of density and velocity of flowing snow. Cold Reg. Sci. Technol., 25(1), 4763.CrossRefGoogle Scholar
Norem, H. 1991. Estimating snow avalanche impact pressure on towers. Eidg. Inst. Schnee- und Lawinenforsch. Mitt. 48, 4256.Google Scholar
Salm, B., Burkard, A. and Gubler, H.. 1990. Berechnung von Fliesslawinen: eine Anleitung für Praktiker mit Beispielen. Eidg. Inst. Schnee- und Lawinenforsch. Mitt. 47.Google Scholar
Schaer, M. and Issler, D.. 2001. Particle densities, velocities and size distributions in large avalanches from impact-sensor measurements. Ann. Glaciol., 32, 321327.CrossRefGoogle Scholar
Sovilla, B., Schaer, M., Kern, M. and Bartelt, P.. 2008a. Impact pressures and flow regimes in dense snow avalanches observed at the Vallée de la Sionne test site. J. Geophys. Res., 113(F1), F01010. (10.1029/2006JF000688.)Google Scholar
Sovilla, B., Schaer, M. and Rammer, L.. 2008b. Measurements and analysis of full-scale avalanche impact pressure at the Vallée de la Sionne test site. Cold Reg. Sci. Technol., 51(2–3), 122137.CrossRefGoogle Scholar
Wieghardt, K. 1975. Experiments in granular flow. Annu. Rev. Fluid Mech., 7, 89114.CrossRefGoogle Scholar
Figure 0

Fig. 1. (a) The avalanche slope at the Vallée de la Sionne test site. A wet avalanche has just hit the obstacle zone. (b) Pylon and sensor installation details (p is pressure, v is velocity and d is density).

Figure 1

Table 1. Summary of the wet avalanche events measured at the Vallée de la Sionne test site during the winter seasons 2003/04 and 2006/07

Figure 2

Fig. 2. Avalanche 8448. (a) Raw pressure data; (b) the average pressure, 〈p〉; (c) fluctuations described by the standard deviation, σp; and (d) the scaling function, σp/〈p〉. Points not scaling are indicated in gray in (d) and with large dots in (b).

Figure 3

Fig. 3. Statistics of fluctuations and average pressures ratio, σp/〈p〉, for each sensor of avalanches 8448, 6231 and 6241. Box plots show mean (symbol in box), median (line in box), 25/75% quantiles (box), 5/95% quantiles (whiskers) and 0/100% quantiles (cross).

Figure 4

Fig. 4. (a) Mean pressures and fluctuations (pressure standard deviations) as a function of flow depth for avalanche 8448. Gray dots show measurements that were excluded from the analysis because σp/〈p〉 did not scale. (b) Pressure and fluctuation standard deviations as a function of flow depth for avalanches 8448, 6236 and 6241. Fitting lines for p = ζρgH are inserted.

Figure 5

Fig. 5. Velocity profiles and shear rates, , extracted from the bulk of avalanches 8448, 6236 and 6241. Time periods, 620 s≤ t ≤660 s for avalanche 8448, 530 s≤ t ≤535 s for avalanche 6236, and 205 s≤ t ≤215 s for avalanche 6241 are shown.

Figure 6

Fig. 6. Pressure as a function of velocity for avalanche 8448. We indicate measurements from the same range of immersion depth by color.