Accepted Manuscript A generalized quasi two-phase bulk mixture model for mass flow Puskar R. Pokhrel, Khim B. Khattri, Bhadra Man Tuladhar, Shiva P. Pudasaini
PII: DOI: Reference:
S0020-7462(17)30635-2 https://doi.org/10.1016/j.ijnonlinmec.2017.12.003 NLM 2946
To appear in:
International Journal of Non-Linear Mechanics
Received date : 15 September 2017 Revised date : 18 September 2017 Accepted date : 4 December 2017 Please cite this article as: P.R. Pokhrel, K.B. Khattri, B.M. Tuladhar, S.P. Pudasaini, A generalized quasi two-phase bulk mixture model for mass flow, International Journal of Non-Linear Mechanics (2017), https://doi.org/10.1016/j.ijnonlinmec.2017.12.003 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1
A Generalized Quasi Two-phase Bulk Mixture Model for Mass Flow
2
Puskar R. Pokhrel1,2 , Khim B. Khattri1 , Bhadra Man Tuladhar1 , Shiva P. Pudasaini3 1
3 2
4 3
5
School of Science, Kathmandu University, Dhulikhel, Kavre, Nepal
Department of Mathematics, R. R. Campus, Tribhuvan University, Kathmandu, Nepal
Department of Geophysics, Steinmann Institute, University of Bonn, Meckenheimer Allee 176, D-53115 Bonn, Germany
6
Correspondence to: Puskar R. Pokhrel, Email:
[email protected]
7
Abstract: We employ full dimensional two-phase mass flow equations (Pudasaini, 2012) to develop a gen-
8
eralized quasi two-phase bulk model for a rapid flow of a debris mixture consisting of viscous fluid and solid
9
particles down a channel. The emerging model, as a set of coupled partial differential equations, is charac-
10
terized by some new mechanical and dynamical aspects of generalized bulk and shear viscosities, pressure,
11
velocities and effective friction for the mixture where all these are evolving as functions of several dynamical
12
variables, physical parameters, inertial and dynamical coefficients and drift factors. These coefficients and
13
factors are uniquely constructed, contain the underlying physics of the system, and reveal strong coupling
14
between the phases. The new model is mechanically described by an extended pressure and rate-dependent
15
Coulumb-viscoplastic rheology of mixture. The model is expected to simulate the velocities and pressure of
16
the bulk mixture much faster than the two-phase mass flow model. Furthermore, the introduction of the ve-
17
locity and pressure drifts makes it possible to reconstruct the two-phase dynamics. This shows the application
18
potential of the new model presented here.
19
Keywords: Mixture flow, Quasi-two-phase flow, Generalized bulk and shear viscosities, Generalized mixture
20
velocities and pressure, Velocity and pressure drifts
21
1
22
Geophysical mass flows such as debris flows, avalanches and landslides generally contain more than one mate-
23
rial and often form a mixture of soil and rock particles with a significant to dominant quantity of interstitial
24
fluid (Johnson, 1970; O’Brien et al., 1993; Iverson and Denlinger, 2001; Pitman and Le, 2005; Pudasaini et al.,
25
2005; Takahashi, 2007). Debris flows as a typical example of geophysical mass flows, are effectively two-phase
26
and gravity driven flows caused by intense rainfall or a sudden surge of water. Such surges may occur due to
Introduction
1
27
the breaking off of a saturated landmass or dam, or by a landslide event that impacts a lake, ocean or river
28
causing over flow that may result in catastrophic dam collapse (Pitman et al., 2013; Mergili et al., 2017a).
29
As the flood surge rushes down the slope, it may transform into a debris flow by entraining the bed material
30
(Hungr et al., 2005; Iverson, 2012; Pudasaini and Fischer, 2016; Mergili et al., 2017c).
31
Debris flows, which generally occur in mountainous areas throughout the world, are extremely destructive
32
and dangerous natural events. A debris flow differs from other types of mass flows including rock and snow
33
avalanches because it contains solid particles and viscous fluid both of which generate fundamentally different
34
forces that significantly influence the motion. During these events, the mixture material undergoes rapid
35
motion and large deformations. The fast motion together with the evolving mixture density provides the
36
debris flow with potentially huge energy and destructive power. Especially the interactions between solid and
37
fluid forces can cause exceptionally long run out (Iverson, 1997; Pitman and Le, 2005; Pudasaini and Miller,
38
2013; Pudasaini and Fischer, 2016).
39
In simulations, debris flows are often treated as a single phase mass flows (e.g., Chen, 1988; O’Brien et
40
al., 1993; Takahashi, 2007), mixture flows (Iverson and Denlinger, 2001; Bohorquez, 2008), a two-fluid flow
41
(Pitman and Le, 2005), and a two-layered model (Fernandez-Nieto et al., 2008). With a major advance in
42
two-phase debris flow modeling, Pudasaini (2012) proposed a comprehensive theory and simulation technique.
43
The model accounts for strong interactions between the solid and the fluid. The model includes buoyancy
44
and other three important physical aspects of the flow: enhanced non-Newtonian viscous stress, virtual mass
45
and the generalized drag. The depth-averaged version of this general two-phase mass flow model has been
46
widely applied in different flow scenarios: rock-ice avalanche dynamic simulation (Pudasaini and Krautblatter,
47
2014), tsunami generated by debris impact on lakes and submarine sediment transport (Kafle et al., 2016),
48
and glacial lake outburst flood (Kattel et al., 2016). Furthermore, employing the general two-phase mass
49
flow model (Pudasaini, 2012) in the computational technique, r.avaflow, Mergili et al. (2017a) presented the
50
simulation results for the interaction and propagation of mass flow in real mountain topographies. Using the
51
same model and tool, Mergili et al. (2017b, 2017c) studied the complex hydro-geomorphic process chains
52
in a multi-lake outburst flood in the Santa Cruz Valley (Cordillera Blanca, Peru). These are the first ever
53
explicit simulations for the evolution of the solid and fluid phase in subaerial and submarine two-phase debris
2
54
flows. Numerical results indicate that the model can adequately describe the complex dynamics of subaerial
55
two-phase debris flows, particle-laden flows, sediment transport, submarine debris flows and associated phe-
56
nomena. These results highlight the applicability of the two-phase model to a wide range of geophysical mass
57
flows.
58
Here, by reducing the general three-dimensional full two-phase model (Pudasaini, 2012), we generate a gener-
59
alized quasi two-phase full (non-depth averaged) two-dimensional model for rapid bulk mixture flow down a
60
channel. To do so, we combine the solid and fluid phase mass balances, and similarly the momentum balance
61
equations to obtain a quasi-two phase bulk mixture model. This is achieved by introducing the drift factors
62
for fluid velocities and the fluid pressure in terms of the solid velocities and solid pressure, respectively. We
63
constitute some structurally new concepts of dynamically evolving drift induced generalized barycentric veloc-
64
ities and pressure for the bulk mixture flow. Similarly, two further structures of bulk mixture viscosities, and
65
the effective bulk and shear viscosities lead to the emergence of the representative complex mixture viscosity.
66
The emerging model for bulk mixture flow consists of the full (non-depth averaged) two-dimensional quasi
67
two-phase mass and momentum conservation equations for the generalized mixture velocities and pressure.
68
The derived mass and momentum equations appear in non-conventional form due to the inertial coefficients,
69
and the form of the mixture viscosities. This model structure results in a novel, dynamically evolving effective
70
mixture friction coefficient, which made it possible to extend the pressure- and rate-dependent Coulomb-
71
viscoplastic granular flow rheology (Domnik and Pudasaini, 2012; Domnik et al., 2013) to the flow of debris
72
mixtures. This is an important mechanical aspect of the derived mixture model.
73
Several important research activities have been carried in the past in the mixture theory of continuum mechan-
74
ics (Truesdell, 1957; Atkin and Craine, 1976a, 1976b; Bowen, 1976; Bedford and Drumheller, 1983; Truesdell,
75
1984; Rajagopal and Tao, 1995; Massoudi, 2003, 2010). They have focused on the modeling of a mixture of
76
two fluids and fluid mixed with solid particles so that each component is considered as a single continuum.
77
Iverson (1997), Iverson and Denlinger (2001) and Pudasaini et al. (2005) progressed significantly in modelling
78
effectively single-phase mixture flows taking into account the pore fluid pressure. Furthermore, Fernandez-
79
Neito et al. (2008) presented two-layered model whereas Pitman and Le (2005) proposed a two-fluid model
80
for debris flow. However, our modeling approach first develops the mixture pressure, velocity and mixture
3
81
shear and bulk viscosities to define the mixture stress tensor. The two-phase dynamics of solid and fluid
82
can then be re-constructed from the definition of the mixture quantities and drift coefficients based on the
83
known velocity and pressure of the mixture. This makes our model mechanically and structurally novel. On
84
the one hand, the simulations based on the new model should be computationally substantially faster than
85
the full two-phase simulations. On the other hand, such simulations are expected to be more accurate than
86
the effectively single-phase models (Iverson and Denlinger, 2001; Pudasaini et al., 2005). This highlights the
87
application potential of the proposed model.
88
2
89
In a two-phase model, the phases have distinct material properties. In general, the fluid phase is characterized
90
by an isotropic stress distribution and a viscosity ηf and density ρf . Similarly, the solid phase is characterized
91
by an anisotropic stress distribution, as a function of an internal frictional angle φ and the basal friction angle
92
δ and the material density ρs (Savage and Hutter, 1989; Pitman and Le, 2005). The balance equations for
93
the three-dimensional, two-phase debris flow are given by (Pudasaini, 2012):
Three Dimensional Model for Two-Phase Mass Flow
∂αs + ∇ · (αs us ) = 0, ∂t
(1)
∂αf + ∇.(αf uf ) = 0, ∂t
(2)
∂ (αs ρs us ) + ∇.(αs ρs us ⊗ us ) = αs ρs f − ∇.αs Ts + ps ∇αs + Ms , ∂t
(3)
∂ (αf ρf uf ) + ∇.(αf ρf uf ⊗ uf ) = αf ρf f − αf ∇pf + ∇.αf τ f + Mf , ∂t
(4)
94
where (1) and (2) are the mass balances for the solid and fluid, and (3) and (4) are the momentum balances
95
for the solid and the fluid fractions, respectively, with Ms =
τf
αs αf (ρs − ρf )g (uf − us )|uf − us |J−1 [UT {PF(Rep ) + (1 − P)G(Rep )}]J ∂uf ∂us +αs ρf CV M + uf .∇uf − + us .∇us = −Mf , ∂t ∂t
A(αf ) = ηf ∇uf + (∇uf )T − ηf [(∇αs )(uf − us ) + (uf − us )(∇αs )] . αf
(5) (6)
96
In these model equations, t is time, αs is the solid volume fraction, αf (= 1 − αs ) is the fluid volume fraction.
97
Similarly, us = (us , vs , ws ) is the velocity vector for solid and uf = (uf , vf , wf ) is that for fluid, ⊗ is the tensor 4
98
product, f is the body force density; Ts is the solid stress tensor; ps and pf are solid and the fluid pressures;
99
Ms and Mf (= −Ms ) are interfacial momentum transfers for solid and fluid respectively. Furthermore, τ f
100
denotes the extra stress for fluid. The exponent J can have the values 1 (for linear drag) and 2 (for quadratic
101
drag). The first term of Ms is the drag force induced by the differences of the phase velocities, and the
102
coefficient of drag CDG =
103
G and fluid-like F drag contributions to flow resistance. F and G are functions of densities, volume fractions,
104
and the particle Reynolds number Rep given by F = (ρf /ρs )(αf /αs )3 Rep /180 and G = αf
105
is a parameter depending on Rep (Richardson and Zaki, 1954; Pitman and Le, 2005). The behavior of drag
106
depends on the interpolation parameter P ∈ [0, 1] which combines G and F. Setting P = 0 is more suitable
107
when the solid particles are moving through a fluid and P = 1 for flows of fluids through dense packing of
108
grains (Pitman and Le, 2005; Pudasaini, 2012). The drag coefficient depends on the solid volume fraction,
109
and increases with increasing values of αs and P. The second term of Ms is the virtual mass force (Ishii, 1975;
110
Ishii and Zuber, 1979; Drew, 1983; Drew and Lahey, 1987; Rivero et al., 1991; Maxey and Riley, 1993) induced
111
by the difference in the acceleration between phases with CV M as virtual mass coefficient. With virtual mass
112
force, the solid particles may bring more fluid mass with them. This results in loss of some inertia of solid
113
mass that induces the kinetic energy of the fluid (Pudasaini, 2012). The first term of τ f is the Newtonian
114
viscous stress with T as the transpose operator, and the second term of τ f is an enhanced Non-Newtonian
115
viscous stress that depends on the solid volume-fraction gradient ∇αs (also see, Ishii, 1975; Drew, 1983; Ishii
116
and Hibiki, 2006; Pudasaini, 2012). A is the mobility of fluid at the particle-fluid interface. There is a strong
117
coupling between the solid and the fluid momentum transfer both through the interfacial momentum transfer,
118
which includes the viscous drag, the virtual mass force; and the enhanced viscous fluid stress.
119
3
120
Here, by reducing the three-dimensional, full two-phase model (1)-(4), we generate a generalized quasi two-
121
phase, full two-dimensional model for rapid bulk mixture flow down a channel. For simplicity, we assume
122
that the difference between the phase accelerations is negligible, i.e., the virtual mass coefficient is close to
123
zero, and the non-Newtonian term in the fluid extra stress tensor is insignificant. However, we note that in
αs αf (ρs − ρf )g is modeled by a linear combination of solid-like [UT {PF(Rep ) + (1 − P)G(Rep )}]J M (Rep )−1
, where M
A Quasi Two-phase Generalized Bulk Mixture Model
5
Figure 1: Side view of the rapid flow of debris material along an inclined chute. ζ is the inclination angle, (usin , ufin ) and (hsin , hfin ) are solid and fluid phase inflow velocities and heights (modified from Domnik and Pudasaini, 2012).
124
the following model formulations, these terms could be retained and the model can be extended for three-
125
dimensional unconfined flows. Let us = (us , ws ), uf = (uf , wf ) be the respective velocity components for
126
solid and fluid in the down-slope (x) and perpendicular to the channel surface (z) directions as shown in Fig.
127
(1). Since αs + αf = 1, adding the mass balance equations (1) and (2), we get the mixture mass balance ∇.(αs us + αf uf ) = 0.
(7)
128
In order to develop a reduced, but generalized quasi two-phase bulk mixture model we introduce the solid and
129
fluid velocity drift coefficients (Zuber and Findlay, 1965; Wallis, 1969; Manninen et al., 1996; Krasnopolsky
130
et al., 2016; Pudasaini and Fischer, 2016), (λu , λw ) and the pressure drift coefficient λp , as
131
uf = λu us , wf = λw ws ; pf = λp ps .
(8)
i i ∂ h ∂ h (αs + λu αf )us + (αs + λw αf )ws = 0. ∂x ∂z
(9)
Then (7) reduces to
6
132
This mass balance is explicitly written in terms of the solid phase (velocities). However, the fluid-phase is also
133
represented through the velocity drift parameters (λu , λw ). Interestingly, it reduces to a traditional single-
134
phase bulk mass balance when λu = λw = 1, which does not distinguish between the solid and fluid velocity,
135
or the phase dynamics (Iverson and Denlinger, 2001; Pudasaini et al., 2005; Fernandez-Nieto et al., 2008).
136
As the solid and fluid phases separately are considered to be incompressible (Mills, 1966), ρs and ρf are both
137
constants. Dividing by ρs and ρf on both sides of (3) and (4) respectively, we get ∂ αs Ms αs , (αs us ) + ∇.(αs us ⊗ us ) = αs f − ∇ps + ∇. τ s + ∂t ρs ρs ρs
(10)
αf αf Mf ∂ (αf uf ) + ∇.(αf uf ⊗ uf ) = αf f − ∇pf + ∇. τ f + , ∂t ρf ρf ρf
(11)
138
where, the solid and fluid Cauchy stresses are now written in terms of the solid and fluid pressures (ps , pf )
139
and deviatoric stresses (τ s , τ f ), respectively (Drew, 1983). Following Domnik and Pudasaini (2012) and
140
Domnik et al. (2013), we will model the internal deformation of the solid component in debris bulk as a
141
Coulomb-viscoplastic material with yield strength (Bingham, 1922; Jop et al., 2006; Ancy, 2007; Balmforth
142
and Frigaard, 2007; Forterre and Pouliquen, 2008; Pailha and Pouliquen, 2009). The extra fluid stress tensor
143
τ f for fluid, and τ s for solid are (Goddard, 1996; Tardos, 1997; Domnik et al., 2013; Schaefer, 2014) Ds τ f = ηf ∇uf + (∇uf )T , τ s = 2ηs Ds + 2τys . ||Ds ||
(12)
144
Here, τ s is the deviatoric stress tensor for a viscoplastic material (granular fluid). With the yield stress
145
τys = τc + τp ps , where τc is cohesion, it becomes pressure- and rate-dependent Coulomb-viscoplastic law for
146
solid-phase (Domnik and Pudasaini, 2012; Domnik et al., 2013; von Boetticher et al., 2016, 2017). As in the
147
fluid, the symmetric part of the velocity gradient Ds =
148
is the dynamic viscosity. The norm of Ds is defined by ||Ds || =
149
Pudasaini, 2012). The effective kinematic viscosity for solid takes the form: νse = νs +
τys , ||Ds ||
1 2 [∇us
+ (∇us )T ] is the strain-rate tensor, and ηs p
2tr(D2s ) (Jop et al., 2006; Domnik and
(13)
150
where νs = ηs /ρs and τys := τys /ρs are the solid kinematic viscosity and yield stress normalized by the solid
151
density. During the mass flow, with τys = 0, the material behaves as a Newtonian fluid for which νse = νs .
152
Now, τ s in (12) can be written as τ s = 2νse Ds , where τ s := τ s /ρs . In what follows, for simplicity, we replace
153
νes by νs . 7
154
Next, we construct a generalized single momentum balance equation for a bulk mixture flow. Adding (10)
155
and (11), we get: ∂ Ms Mf (αs us + αf uf ) + ∇ · [(αs us ⊗ us ) + (αf uf ⊗ uf )] = f + ∇ · (αs τ s + αf τ f ) − (αs ∇ps + αf ∇pf ) + + , (14) ∂t ρs ρf
156
where, pressures (ps , pf ) and stresses (τ s , τ f ) are normalized with respective densities, and the viscosities in
157
(τ s , τ f ) are kinematic viscosities. With uf = λu us and wf = λw ws , the local time derivative of the ‘mixture
158
velocity’ in (14) becomes i ∂ ∂h (αs us + αf uf ) = (αs + λu αf )us , (αs + λw αf )ws . ∂t ∂t
159
(15)
The convective terms in (14) can be written as ∂ ∂ [(αs + λ2u αf )u2s ] + [(αs + λu λw αf )us ws ], ∂x ∂z ∂ ∂ [(αs + λu λw αf )us ws ] + [(αs + λ2w αf )ws2 ] . ∂x ∂z
∇ · [(αs us ⊗ us ) + (αf uf ⊗ uf )] =
160
(16)
The divergence of the rate-of-strain tensors in (14) become ∇ · [αf τ f + αs τ s ] h∂ ∂uf ∂uf ∂wf ∂ ∂ ∂us ∂ ∂us ∂ws 2νf αf + νf αf + + 2νs αs + νs αs + , ∂x ∂x ∂z ∂z ∂x ∂x ∂x ∂z ∂z ∂x ∂uf ∂wf ∂wf ∂ ∂ ∂ ∂us ∂ws ∂ ∂ws i P P pppp νf αf + + 2νf αf + νs αs + + 2νs αs ∂x ∂z ∂x ∂z ∂z ∂x ∂z ∂x ∂z ∂z h ∂ ∂us ∂ ∂us ∂ws PPp = 2 (νs αs + λu νf αf ) + (νs αs + λu νf αf ) + (νs αs + λw νf αf ) , ∂x ∂x ∂z ∂z ∂x ∂ ∂us ∂ws ∂ ∂ws i P P pppp (νs αs + λu νf αf ) + (νs αs + λw νf αf ) +2 (νs αs + λw νf αf ) . (17) ∂x ∂z ∂x ∂z ∂z PPp =
161
With the definition pf = λp ps , the bulk mixture pressure gradients in (14) become ∂ps ∂ps − (αs ∇ps + αf ∇pf ) = − (αs + λp αf ) , (αs + λp αf . ∂x ∂z
(18)
162
For simplicity, in the above formulations, we have assumed that variations of λu , λw , λp with x, and z are
163
negligible. Finally, we deal with the total interfacial momentum exchange in (14): αs αf g Ms Mf + = (1 − γ) (uf − us )|uf − us |J−1 ρs ρf [UT {PF(Rep ) + (1 − P)G(Rep )}]J αs αf g 1 P ppppppP P − −1 (uf − us )|uf − us |J−1 . γ [UT {PF(Rep ) + (1 − P)G(Rep )}]J 8
164
165
Since uf − us = (λu − 1)us , (λw − 1)ws , we have h i αs αf g (λu − 1)us , (λw − 1)ws Ms Mf 1 + = − (1 − γ)2 ρs ρf γ [UT {PF(Rep ) + (1 − P)G(Rep )}]J which become zero if γ → 1 or λu → 1, λw → 1 .
J−1 (λ − 1)u , (λ − 1)w , u s w s
(19)
166
So, with f = (fx , fz ) from (15)-(19), (14) yields the following x- and z- directional momentum balances for
167
the bulk mixture: ∂ ∂ ∂ ∂ps [(αs + λu αf )us ] + (αs + λ2u αf )u2s + [(αs + λu λw αf )us ws ] = fx − (αs + λp αf ) ∂t ∂x ∂z ∂x ∂us ∂ ∂us ∂ ∂ws (νs αs + λu νf αf ) + (νs αs + λu νf αf ) , ppppP P + 2 + (νs αs + λw νf αf ) ∂x ∂x ∂z ∂z ∂x
(20)
∂ ∂ ∂ ∂ps [(αs + λw αf )ws )] + [(αs + λu λw αf )us ws ] + [(αs + λ2w αf )ws2 )] = fz − (αs + λp αf ) ∂t ∂x ∂z ∂z ∂us ∂ws ∂ ∂ws ∂ (νs αs + λu νf αf ) + (νs αs + λw νf αf ) +2 (νs αs + λw νf αf ) . (21) ppppP P + ∂x ∂z ∂x ∂z ∂z 168
With the definition of the bulk velocities, and the pressure of the mixture flow um = (αs + λu αf )us , wm = (αs + λw αf )ws , pm = (αs + λp αf )ps ,
169
the inertial coefficients Λuu =
170
(22)
αs + λ2u αf αs + αf λ2w αs + λu λw αf 1 1 , Λ = , Λuw = , Λu = , Λw = , (23) ww (αs + λu αf )2 (αs + λw αf )2 (αs + λu αf )(αs + λw αf ) αs + λu αf αs + λw αf
and the dynamical coefficients for mixture viscosities and pressure Ληu = νs αs + λu νf αf , Ληw = νs αs + λw νf αf , Λp = αs + λp αf ,
(24)
171
from (9), (20) and (21), the new dynamical model for bulk mixture flow consist of the non depth-averaged
172
two-dimensional quasi two-phase mass and momentum conservation equations: ∂um ∂wm + = 0, ∂x ∂z
(25)
∂um ∂ ∂ ∂pm + (Λuu u2m ) + (Λuw um wm ) = fx − ∂t ∂x ∂z ∂x ∂ ∂(Λu um ) ∂ ∂(Λu um ) ∂(Λw wm ) PPPPPPPPP + 2 Ληu + Ληu + Ληw , ∂x ∂x ∂z ∂z ∂x ∂wm ∂ ∂ ∂pm 2 + (Λuw um wm ) + (Λww wm ) = fz − ∂t ∂x ∂z ∂z ∂ ∂(Λu um ) ∂(Λw wm ) ∂ ∂(Λw wm ) PPPPPPPPP + Ληu + Λ ηw +2 Λ ηw , ∂x ∂z ∂x ∂z ∂z 9
(26)
(27)
173
where, for simplicity, the relations Λp ∂ps /∂x =: ∂pm /∂x, and Λp ∂ps /∂z =: ∂pm /∂z, are defined. The simple
174
situation being the one for which variation of Λp is negligible. pm is called the “kinematic pressure” for the
175
bulk. It appears that the dynamics of the bulk pressure, pm is complex. We will discuss on this in more detail
176
later. The inertial and dynamical coefficients in (26) and (27) are complex and coupled, and explicitly include
177
the physics and dynamics of mixture, such as the solid volume fraction, solid and fluid material densities
178
and viscosities; solid friction and yield strength, rate of deformation of solid, and velocity and pressure drift
179
parameters. Moreover, the mass and momentum equations in (25)-(27) appear in non-conventional form due
180
to the inertial coefficients (Λu , Λw ), and the non-equality of the mixture viscosities Ληu and Ληw . In the form,
181
the model equations (25)-(27) may look similar to the existing (effectively single-phase) mixture mass flow
182
models (Iverson, 1997; Iverson and Denlinger, 2001; Pudasaini et al., 2005). However, there are fundamentally
183
new aspects in the new model than those found in literature. This has been discussed in Section 5 and also
184
in the following sections.
185
From the application and computational point of view, the contraction (25)-(27) is useful, because it reduces
186
the full two-phase model to a quasi two-phase bulk model. This paves the way to apply the simple single-
187
phase computational tools of Domnik and Pudasaini (2012) and Domnik et al. (2013), with some amendments
188
to numerically solve the emerged model equations, that to a large extent, describe a mixture flow of solid
189
particles and viscous fluid. However, at the same time, the drift parameters (λu , λw , λp ) contain the basic
190
and important information relating the solid and fluid velocities, and the dynamic pressures. Further models
191
can be developed to describe these parameters. Alternatively, laboratory or field data can be utilized either
192
to obtain solid velocities (us , ws ) and pressure ps , or fluid velocities (uf , wf ) and pressure pf , and simulate
193
the dynamics for the other field variables. This makes the resulting model a generalized quasi two-phase bulk
194
mixture model. For λu = λw = λp = 1 the emerging model reduces to traditional effectively single-phase bulk
195
mixture model (Iverson and Denlinger, 2001; Pudasaini et al., 2005). The new model has a potential for a
196
wider application through constructing the functional relations for (λu , λw , λp ), or via parameterizations; the
197
advantage is that computationally it is faster than the full two-phase simulations, and, at the same time, more
198
accurate than the single-phase models.
10
199
4
200
Here, we present the final set of model equations that is written in a convenient form for applications and
201
numerical simulations (Domnik et al., 2013).
202
4.1
203
To calculate the pressure pm , we construct the pressure Poisson equation for the bulk mixture. The x- and
204
z- directional momentum equations (26) and (27) can be written as
205
The New Model Equations
Pressure Poisson Equations
∂pm ∂um =− + F, ∂t ∂x
(28)
∂wm ∂pm =− + G, ∂t ∂z
(29)
(30)
where ∂(Λu um ) ∂ ∂(Λu um ) ∂(Λw wm ) + Ληu +Ληw +fx , ∂x ∂z ∂z ∂x ∂ ∂ ∂(Λw wm ) ∂ ∂(Λu um ) ∂ ∂(Λw wm ) 2 )+ +Ληw G = − (Λuw um wm )− (Λww wm Λ ηu +2 Λ ηw +fz . ∂x ∂z ∂x ∂z ∂x ∂z ∂z F =−
206
∂ ∂ ∂ (Λuu u2m )− (Λuw um wm )+2 ∂x ∂z ∂x
(31)
Partially differentiating (28) with respect to x, and (29) with respect to z, and adding them, we get ∂ ∂x
207
Ληu
∂um ∂t
+
∂ ∂z
∂wm ∂t
=
∂ ∂t
∂um ∂wm + ∂x ∂z
=−
∂ 2 pm ∂ 2 pm ∂F ∂G − + + . ∂x2 ∂z 2 ∂x ∂z
This, with the continuity equation (25) results in the following pressure Poisson equation for the mixture flow ∂ 2 pm ∂ 2 pm ∂F ∂G + = + . ∂x2 ∂z 2 ∂x ∂z
(32)
208
4.2
The Bulk Mixture Model
209
With the definitions of F and G from (30) and (31), the following set of x- and z-directional momentum, and
210
pressure Poisson equation together constitute a new generalized quasi two-phase mixture mass flow model: ∂pm ∂um =− + F, ∂t ∂x
(33)
∂wm ∂pm =− + G, ∂t ∂z
(34)
∂ 2 pm ∂ 2 pm ∂F ∂G + = + . ∂x2 ∂z 2 ∂x ∂z
(35)
11
211
The set (33)-(35) is a system of three highly non-linear partial differential equations with three unknowns,
212
namely, the generalized mixture velocities um , wm , and pressure pm . So, the system is closed and the numerical
213
integration is possible provided that the coefficients Λuu =
αs + λ2u αf αs + αf λ2w αs + λu λw αf 1 1 , Λww = , Λuw = , Λu = , Λw = , 2 (αs + λu αf ) (αs + λw αf )2 (αs + λu αf )(αs + λw αf ) αs + λu αf αs + λ w αf
Ληu = νs αs + λu νf αf , Ληw = νs αs + λw νf αf , Λp = αs + λp αf ,
(36)
214
are known. Equations (33)-(35) require appropriate boundary conditions for the velocities and pressure at
215
the free and the basal surface, and also the initial conditions for these variables. This has been discussed in
216
Section 8.
217
4.3
218
There are several ways to model the drift coefficients (Zuber and Findlay, 1965; Wallis, 1969; Manninen et
219
al., 1996; Krasnopolsky et al., 2016; Pudasaini and Fischer, 2016). Simplified flow situations can be utilized
220
to reduce the drift parameters λu and λw into a single one. Assuming locally small variations of αs and λu ,
221
the mixture mass balance (25), and the steady state solid mass balance (1) leads to
Modelling the Drift Coefficients
wf ≈ λ0 + λu ws ,
(37)
222
where λ0 is an integration constant whose value can be fixed from the boundary and physical conditions.
223
Since wf = λw ws , by replacing λw = λ0 /ws + λu , the model equations contain effectively a single velocity drift
224
parameter, λu . Note that this does not produce singularity, because while substituting λw in the respective
225
terms in the model equations (33)-(36), ws in the denominator cancels out. In particular, this shows that
226
when λ0 ≈ 0, λw ≈ λu . This is advantageous, because explicit descriptions are available for λu . GhoshHajra
227
et al. (2015) analytically derived an expression for λu , as a ratio of the fluid and solid pressure parameters
228
βf and βs , λu =
229
averaged equations (Kafle et al., 2016; Kattel et al., 2016). Then, we may dynamically obtain λu = uf /us .
230
Furthermore, we may also parameterize λw as a function of λu , e.g., a linear or a quadratic relation.
p βf /βs . Alternatively, us and uf can be obtained from the fast simulations of the depth
12
231
5
Discussions on Physical Aspects of the New Model
232
The model equations (33)-(35) are written as well structured partial differential equations in conservative form.
233
There are several important features of the new model. Some physical aspects, implications, and applicabilities
234
are discussed here. We have presented two new structures (concepts) of bulk mixture viscosities and pressures
235
as functions of the solid volume fraction, solid and fluid material densities and viscosities, solid friction and
236
yield strength, rate of deformation of solid, and velocity and pressure drift parameters. One of the advantages
237
of the coupled and generalized model (33) - (35) is that structurally it is the same as the single phase
238
granular or debris flow model (Domnik and Pudasaini, 2012; Domnik et al., 2013). So, in principle, the same
239
computational strategy can be applied to solve the new system. However, the complexity inherited by the
240
two-phase bulk mixture flow is contained in the parameters Λu , Λw , Λuu , Λuw , Λww , Ληu , Ληw , Λp . Importantly,
241
we can reconstruct some fundamental properties and dynamics of two-phase flows.
242
5.1
243
The two-phase dynamics is largely lost in the classically derived effectively single-phase mixture mass flow
244
model (Iverson and Denlinger, 2001; Pudasaini et al., 2005). But, in our derivation, once um , wm and pm are
245
obtained from (33)-(35), the two-phase dynamics of solid (us , ws , ps ) and fluid (uf , wf , pf ) can be calculated
246
from (22) and (8). That is, (us , ws , ps ) are obtained from (22), and then (uf , wf , pf ) are systematically re-
247
constructed from (8). However, how much reality these solutions represent depends on the physics contained
248
in (λu , λw , λp ), the viscosities (Ληu , Ληw ) and the description of αs .
249
5.2
250
The viscous terms appear in the momentum equations (20) and (21). From the third term on the right hand
251
side of (20), the bulk viscosity can be obtained
Reconstruction of the Two-phase Dynamics
Effective Bulk Viscosity
∂us ∂(λu us ) ∂ 1 ∂ λu ∂um + νf αf = νs αs um +νf αf um ≈ (νs αs Λu + νf αf λu Λu ) . (38) νs αs ∂x ∂x ∂x αs + λu αf ∂x αs + λu αf ∂x 252
253
Similarly, from the fourth term on the right hand side of (21), another expression for the bulk viscosity can be obtained νs αs
∂ws ∂(λw ws ) ∂ 1 ∂ λw ∂wm + νf αf = νs αs wm +νf αf wm ≈ (νs αs Λw + νf αf λw Λw ) . (39) ∂z ∂z ∂z αs + λw αf ∂z αs + λw αf ∂z
13
254
These expressions are valid if Λu and λu vary linearly with x, and Λw and λw vary linearly with z. These
255
approximations separately concern the x- and z-directional bulk viscosities. The effective (representative)
256
‘bulk viscosity’ of the mixture is obtained by averaging over (38) and (39): 1 1 ΛB ηm := νs αs (Λu + Λw ) + νf αf (λu Λu + λw Λw ). 2 2
(40)
257
If λu and λw are increasing, then (Λu + Λw ) and (λu Λu + λw Λw ) are decreasing. Thus, the viscosity of the
258
mixture is also decreasing. This means the higher are the fluid drifts, the faster is the fluid motion, and thus,
259
the higher is the deformation. This results in the lower bulk viscosity, as in a shear-thinning fluid (Pudasaini,
260
2011). This is physically meaningful.
261
5.3
262
The shear viscosities are contained in the fourth term on the right hand side of (20), and similarly, the third
263
term on the right hand side of (21):
264
265
Effective Shear Viscosity
(νs αs + λu νf αf )
∂us ∂ws ∂um ∂wm + (νs αs + λw νf αf ) ≈ (νs αs + λu νf αf )Λu + (νs αs + λw νf αf )Λw , ∂z ∂x ∂z ∂x
(41)
(νs αs + λu νf αf )
∂us ∂ws ∂um ∂wm + (νs αs + λw νf αf ) ≈ (νs αs + λu νf αf )Λu + (νs αs + λw νf αf )Λw . ∂z ∂x ∂z ∂x
(42)
This shows that there are two contributions also for the shear viscosity. Averaging these, we obtain a representative ‘shear viscosity’ of the mixture: 1 1 ΛSηm := νs αs (Λu + Λw ) + νf αf (λu Λu + λw Λw ). 2 2
(43)
266
S So, since ΛB ηm and Ληm are identical, (43) can be considered as the “representative mixture viscosity”, Ληm .
267
However, these are only simplifications, and that the better representations of the viscosities are as they
268
appear in (20), (21). In simple situations, (43) can be further reduced. If the x- and z-directional velocity
269
drifts are identical, i.e., λu ≈ λw (isotropic drifts), then Ληm ≈ (νs αs + νf αf λu ) Λu =
νs αs + νf αf λu . αs + λ u αf
(44)
270
This is a very important structure, because the mixture viscosity Ληm can be measured in a rheometer.
271
Figure 2 reveals that the mixture viscosity, as represented by the simplified form (44), varies strongly with
272
the velocity drift. The chosen parameters are νs = 4.99 m2 /s, νf = 10−3 m2 /s (Domnik et al., 2013; von 14
273
Boetticher et al., 2016, 2017), and for different values of αs . For higher solid fraction (dense flow), the drop in
274
the mixture viscosity with the drift is slower, however, for the lower solid fraction (dilute flow), the mixture
275
viscosity drops exponentially with the velocity drift. This indicates that the mixture flow behaves as a type
276
of shear-thinning material: weakly for higher solid fraction but strongly with the lower solid fraction. Such
277
a rheological behaviour of the mixture is often observed in debris flows (von Boetticher et al., 2016; 2017).
278
Figure 2 shows that Ληm is an inverse exponential function of λu . So, it can be approximated by a hyperbolic
279
tangent function:
Ληm = Λ0ηm (1 − ξ tanh(λχu )) .
(45)
αs = 0.75 αs = 0.50
4
αs = 0.25
m
Mixture viscosity: Λη (m2s−1)
5
3
αs = 0.15
2 1 0 0
0.5
1
1.5
2 2.5 3 Velocity drift: λu
3.5
4
4.5
5
Figure 2: Variation of the mixture viscosity with velocity drift: mixture viscosity depends strongly on the velocity drift and the solid volume fraction in the debris mixture and shows a type of shear-thinning behaviour.
Theoretical derivation Approx. with tanh function
4
m
Mixture viscosity: Λη ( m2 s−1)
5
3 2 1 0 0
0.5
1
1.5
2 2.5 3 Velocity drift: λu
3.5
4
4.5
5
Figure 3: Approximation of the mixture viscosity by a hyperbolic tangent function of drift.
15
12 αs = 0.25 Velocity drift: λu
10
αs = 0.20 αs = 0.15
8
αs = 0.10
6 4 2 0.2
0.3
0.4
0.5 0.6 0.7 Viscosity: Λη (m2 s−1)
0.8
0.9
1
m
Figure 4: Variation of the velocity drift with mixture viscosity.
280
As Λ0ηm is the reference value of Ληm at λu = 0, this can be determined relatively easily. However ξ and χ may
281
be determined from an experiment which is not within the scope here, but can be numerically determined for
282
results presented in Fig. 2. Figure 3 shows that the Ληm in (44) can very well be represented by a hyperbolic
283
tangent function of λu with the parameters Λ0ηm = 4.99 m2 /s, χ = 0.5 and ξ = 0.97. This is important from
284
an application point of view as a relatively simple function (45) can replace (44).
285
Furthermore, (44) shows that λu can be obtained analytically, and explicitly: λu =
(νs − Ληm ) αs . (Ληm − νf ) αf
(46)
286
Since usually νs − Ληm > 0 and Ληm − νf > 0, αs , αf > 0, λu > 0. Such an explicit form of the drift is novel,
287
and technically very important. Because, as mentioned earlier, Ληm may be measured in the lab, then with the
288
knowledge of νs and νf , (46) provides values for the drift, λu , as a function of Ληm . This is a great advantage
289
because this can be used to reconstruct the two-phase dynamics of the mixture. See later, Fig. 5 and Fig. 6,
290
and the corresponding descriptions on the phase reconstructions. Figure 4 plots (46) with νs = 4.99 m2 /s,
291
νf = 10−3 m2 /s, αs = 0.25, 0.20, 0.15, 0.10, and shows that higher is the solid volume fraction, higher are the
292
velocity drifts which decrease with higher viscosity. Figure 2 and Fig. 4 show that the viscosity and velocity
293
drift of the mixture flow vary inversely to each other. The same analysis can be performed for more complex
294
mixture viscosity represented by (43).
295
In connection to (22)1 and (22)2 , (46) can be utilized to reconstruct the solid and fluid velocities in terms of
296
the mixture velocity um and the drift λu . Figure 5 shows an incipient motion of a mixture material. The 16
297
chosen parameters are νs = 4.99m2 /s, νf = 10−3 m2 /s, αs = 0.25. This figure explains how the solid and fluid
298
velocities can be analytically reconstructed as a function of the mixture velocity. As the mass is released,
299
the fluid accelerates faster than the solid phase, but after a while the fluid velocity tends to saturate, which
300
is realistic (Pudasaini, 2011), as the fluid-phase is a viscous material that can reach to a terminal velocity.
301
However, the solid-phase continues to accelerate, this is also realistic because the solid phase is described by
302
a frictional granular material. So, such an incipient behaviour can be described by our modelling approach.
303
Figure 6 shows the depositional behaviour. Assume a situation, as described by Fig. 5 that after traveling
304
a certain distance, the mixture enters the depositional area. At the entry into the depositional area, the
305
solid-phase has higher velocity than the fluid-phase (as seen in Fig. 5). However, as the flow enters into this
306
area, the velocity of both the phases decrease steadily. Nevertheless, the depositional behaviour of the solid
307
and the fluid phase are completely different. The solid decelerates much faster than the fluid. This is so,
308
because the solid material is frictional which dissipates the energy more quickly than the fluid phase which
309
is viscous, that is relatively weaker (because of the viscous dissipation) than the solid. These are realistic
310
scenarios of mixture flow in depositional region.
311
Figure 5 and Fig. 6 show several important features of solid and fluid velocities in the mixture flows. First,
312
these are strongly non-linear. Second, both the solid and fluid velocities behave fundamentally differently.
313
The rates at which the solid and fluid velocities accelerate after the mass release, and the rates of their
Phase velocities: us, uf (ms−1)
2.5 2
Solid−phase velocity Fluid−phase velocity
1.5 1 0.5 0 0
0.2
0.4 0.6 0.8 1 The mixture velocity: um (ms−1)
1.2
1.4
Figure 5: Evolution of the incipient motion of mixture material that explains how the solid and fluid velocities can be reconstructed as a function of the mixture velocity.
17
Phase velocities: us, uf (ms−1)
2.5 Solid−phase velocity Fluid−phase velocity
2 1.5 1 0.5 0
1.2
1
0.8 0.6 0.4 The mixture velocity: um (ms−1)
0.2
0
Figure 6: Depositional behaviour of the solid and the fluid phases velocities as a function of the mixture velocity.
314
decelerations as they enter the depositional region are completely different. These are natural phenomena,
315
but simulated here for the first time in terms of the mixture velocity. This clearly shows the application
316
potential of the new model presented here.
317
If, further, the drifts are close to unity, then Ληm ≈ νs αs + νf αf ,
(47)
318
which is the simplest description of the viscosity of the bulk mixture, or the concentration-weighted average
319
viscosity (von Boetticher et al., 2016). The complex mixture viscosity as it appears in (40) or (43) is newly
320
constructed here, which is rich in its physics, and evolves mechanically as a complex and coupled function of
321
several physical and mechanical parameters and dynamical variables including the solid volume fraction αs ;
322
solid and fluid material densities and viscosities (because the viscosities are kinematic), solid friction, yield
323
strength and rate of deformation of solid (that appears in νs ), and velocity and pressure (because νs containing
324
yield strength that presents on solid pressure) drift parameters, etc.
325
5.4
326
The pressure of the bulk mixture is defined in (22). The solid and fluid pressures can be expressed as
Mixture Pressure
αs ρs ps =
pm αs ρs , αs + λp αf
18
αf ρf pf =
(pm αf ρf )λp . αs + λp αf
(48)
327
With this, the mixture pressure can also be written as pm =
Λp (αs ρs ps + αf ρf pf ) . αs ρs + λp αs ρf
(49)
328
In terms of generalized mixture density (ρpm = αs ρs + λp αf ρf ) with respect to pressure drift λp , the mixture
329
pressure is pm =
Λp (αs ρs ps + αf ρf pf ) . ρpm
330
This constitutes a general concept of dynamically evolving (bulk) mixture pressure. From (22) and (49), we
331
have two equivalent representations of the mixture pressure. For dilute flows (if the extra pressure generated
332
by the grain contacts is very small, i.e., λp ≈ 1), then pm reduces to pm =
1 (αs ρs ps + αf ρf pf ), ρm
(50)
333
where, ρm = αs ρs + αf ρf . pm in (50) is the classical definition of the mixture (barycentric) pressure. So,
334
(49) is called the drift induced generalized barycentric pressure for the mixture flow. This shows that λp
335
may play a crucial role in the pressure dynamics. If the fluid pressure increases (say strongly, i.e., more than
336
hydrostatic), then pm becomes much greater than ps . In this situation, pm ≈ λp αf ps . If λp is very small, from
337
(48), we obtain pm = αs ps . The pressures ps and pf are dynamic. So, both can be positive, or negative, or
338
one of them positive and the other negative. Assume that ps > 0, and the fluid pressure is negative (λp < 0).
339
But, whether pm > 0 or < 0 depends on αs + λp αf . If αs + λp αf > 0, pm > 0, otherwise pm < 0. So,
340
one of the phases may have locally negative pressure, however the bulk pressure can still be positive. It may
341
depend on, how the negative fluid pressure develops and the amount of the locally available fluid. Many
342
other interesting possibilities can be described. For example, large shearing of the solid matrix can result in
343
dilatation leading to the decreased fluid pressure. This implies that bulk pressure can be smaller than the solid
344
pressure, because αs + λp αf < 1. On the contrary, during compaction of the solid phase, the fluid pressure
345
increases substantially, leading to the higher bulk pressure than that of the solid. Therefore, the dynamics of
346
the bulk pressure pm is much more complex than just the dynamics of the solid and fluid pressures separately.
19
347
5.5
Mixture Velocity
348
As for the bulk pressure, the bulk mixture velocities um = (αs + λu αf )us , wm = (αs + λw αf )ws can be
349
expressed as um =
350
αs ρs us + αf ρf uf αs ρs ws + αf ρf wf , wm = . Λu (αs ρs + λu αf ρf ) Λw (αs ρs + λw αf ρf )
With the generalized mixture densities with respect to velocity drifts λu and λw , ρum = αs ρs + λu αf ρf ; ρw m = αs ρs + λw αf ρf ,
351
the mixture velocities can be written in alternative forms as um =
1 1 (αs ρs us + αf ρf uf ) , wm = (αs ρs ws + αf ρf wf ) . Λu ρum Λw ρw m
(51)
352
These are called the drift induced generalized barycentric velocities, or simply the generalized mixture veloc-
353
ities. In special situation when the velocity drifts are close to unity, these expressions reduce to the classical
354
barycentric velocities: um =
1 1 (αs ρs us + αf ρf uf ) , wm = (αs ρs ws + αf ρf wf ) . ρm ρm
(52)
355
As for the pressure, the properties of the generalized velocities can be discussed.
356
6
357
The models obtained in (25)-(27) can be further simplified. Here, we derive reduced models as characterized
358
either by the volume fraction, or by the drift coefficients.
359
6.1
360
Up to now, no assumption has been made on the solid (or, fluid) volume fraction αs . As it stands now, αs
361
appears to be an internal variable. There are two possibilities. (i) αs can be considered as a state variable
362
which is the general situation. This requires an extra closure or, a transport equation namely, the advection-
363
diffusion (Iverson and Denlinger, 2001; Pudasaini et al., 2005):
Reduced Models
Characterization with the Volume Fraction
∂αs ∂αs ∂αs ∂ 2 αs ∂ 2 αs + us + ws − Dx 2 − Dz 2 = 0, ∂t ∂x ∂z ∂ x ∂ z 20
(53)
364
which says that αs advects with the velocity (us , ws ) and diffuses with (Dx , Dz ). Exact solutions can be
365
constructed in computation assuming that (us , ws ) and (Dx , Dz ) are known for some particular time step.
366
This can be updated as required. This way, the model equations (25)-(27) and (53) are dynamically coupled.
367
However, much advanced sub-advection, sub-diffusion models (Pudasaini, 2016) can also be utilized. (ii)
368
In simple situations, either αs can be parameterized, or be considered as constant, which is the scenario in
369
homogeneous flow.
370
6.2
371
6.2.1
372
In situations when the solid and fluid pressure are close to each other, the pressure drift coefficient is close
373
to unity, i.e., λp ≈ 1. When instantaneous pressures are in equilibrium, this can be a valid assumption, for
374
example, if the curvature and/or the surface tension is negligible. This holds only for substantially dilute
375
flows with very low particle contacts. When the contact area is a small fraction of the total interfacial area
376
(e.g., loose/intermittent contacts), interfacial solid and fluid pressures can be approximated by fluid pressure
377
(Kuo et al., 1976; Gough, 1980; Drew, 1983). Then, the model equations (9), (20) and (21) reduce to
Characterization with Drifts Pressure Drift Factor Close to Unity
i i ∂ h ∂ h (αs + λu αf )us + (αs + λw αf )ws = 0, ∂x ∂z
378
∂ ∂ ∂ ∂ps [(αs + λu αf )us ] + (αs + λ2u αf )u2s + [(αs + λu λw αf )us ws ] = fx − ∂t ∂x ∂z ∂x ∂ ∂us ∂ ∂us ∂ws (νs αs + λu νf αf ) + (νs αs + λu νf αf ) , (54) ppppP P + 2 + (νs αs + λw νf αf ) ∂x ∂x ∂z ∂z ∂x ∂ ∂ ∂ ∂ps [(αs + λw αf )ws )] + [(αs + λu λw αf )us ws ] + [(αs + λ2w αf )ws2 )] = fz − ∂t ∂x ∂z ∂z ∂ ∂us ∂ws ∂ ∂ws ppppP P + (νs αs + λu νf αf ) + (νs αs + λw νf αf ) +2 (νs αs + λw νf αf ) . (55) ∂x ∂z ∂x ∂z ∂z
379
There are only two velocity drifts remaining in these equations.
21
380
6.2.2
Identical Velocity Drift Factors
381
The velocity drift factors may have the identical values in the x- and z-directions: λu = λw 6= 1. Then, (9),
382
(54) and (55) reduce to i i ∂ h ∂ h (αs + λu αf )us + (αs + λu αf )ws = 0, ∂x ∂z
∂ ∂ ∂ps ∂ [(αs + λu αf )us ] + (αs + λ2u αf )u2s + (αs + λ2u αf )us ws = fx − (αs + λp αf ) ∂t ∂x ∂z ∂x ∂ ∂us ∂ ∂us ∂ws ppppP P + 2 (νs αs + λu νf αf ) + (νs αs + λu νf αf ) , + ∂x ∂x ∂z ∂z ∂x
383
∂ ∂ ∂ ∂ps [(αs + λu αf )ws )] + [(αs + λ2u αf )us ws ] + [(αs + λ2u αf )ws2 )] = fz − (αs + λp αf ) ∂t ∂x ∂z ∂z ∂us ∂ws ∂ ∂ws ∂ (νs αs + λu νf αf ) +2 (νs αs + λu νf αf ) . ppppP P + + ∂x ∂z ∂x ∂z ∂z
(56)
(57)
(58)
384
As compared to the general situation, the advantage now is that the bulk mixture viscosity is simply given by
385
the single expression νs αs + λu νf αf , which, however, still evolves as a function of the velocity drift λu , and
386
other physical and dynamical variables included in νs and αs .
387
6.2.3
388
In the most simple situation, when the velocity drifts are close to unity (λu , λw ≈ 1), (56)-(58) further reduce
389
to
Velocity Drift Factors Close to Unity
∂us ∂ws + = 0, ∂x ∂z
(59)
∂us ∂ ∂us ∂ws Λm + Λm + , ∂x ∂z ∂z ∂x ∂ws ∂(us ws ) ∂(ws2 ) ∂ps ∂ ∂us ∂ws ∂ ∂ws + + = fz − + Λm + +2 Λm , ∂t ∂x ∂z ∂z ∂x ∂z ∂x ∂z ∂z ∂us ∂(u2s ) ∂(us ws ) ∂ps ∂ + + = fx − +2 ∂t ∂x ∂z ∂x ∂x
(60) (61)
390
where Λm = νs αs + νf αf is now the traditional simple bulk mixture viscosity which is independent of the
391
velocity drift. And, since the velocity drifts are close to unity, the pressure drift may also be close to unity,
392
which is legitimate. However, for similar solid and fluid velocities, the solid and fluid pressures can be different.
393
Even the simplest set of equations (59)-(61) is important because this includes the first order basic dynamics
394
of evolving viscosity that governs the dynamics of the mixture flow.
22
395
7
The Flow Rheology
396
The mixture motion and the settlement are described by constitutive law. Geophysical mass flows can be mod-
397
elled by viscoplastic rheology (Jop et al., 2006; Ancey, 2007; Balmforth et al., 2007; Forterre and Pouliquen,
398
2008). Domnik et al. (2013) proposed the pressure and rate-dependent Coulomb viscoplastic rheological
399
model to describe the full dynamics of the rapid flows of granular materials down the channels impinging on
400
rigid walls. Unlike Herschel-Bulkley, or Bingham plasticity, Domnik et al. (2013) model does not include
401
any fit parameter, and is fully described by the phenomenological parameters of the materials. With several
402
flow simulations, from the material collapse and silo outlet to the final depositions, they have demonstrated
403
the applicability of their new rheological model with very high computational performances. Their model
404
is capable of describing several important observation in wide range of geophysical mass flows as recently
405
demonstrated by von Boetticher et al. (2016, 2017).
406
Iverson (1997), Iverson and Denlinger (2001), Massoudi (2003, 2010) and Pudasaini et al. (2005) used the
407
total stress in a mixture as the sum of solid and fluid-phases stress tensors. On contrary, our approach is
408
completely different. We employ the mixture pressure, velocity and mixture shear viscosity (as constructed
409
in Section 5, and mixture strain-rate tensor, see below), to construct the mixture stress tensor. So, flow
410
rheology intrinsic to the model (25)-(27) is rather complex and non-conventional. It simultaneously includes
411
the physical and dynamical properties of both the solid and fluid components. From (25)-(27), the Cauchy
412
stress tensor for the mixture takes the form: σ m = −pm I + 2Ληm Dm ,
413
(62)
where, pm is the normalized pressure (49), Ληm is the mixture viscosity (43) 1 1 Ληm = νse αs (Λu + Λw ) + νf αf (λu Λu + λw Λw ). 2 2
414
Λ T Λ Dm = 12 [∇uΛ m + (∇um ) ] is the strain-rate tensor for mixture, where um = (Λu um , Λw wm ). So, the velocity
415
gradient terms in Dm are in extended and non-conventional form, and reduce to the standard form if Λu =
416
Λw = 1.
23
417
The effective kinematic viscosity for solid νse = νs +
τys ||Dm ||
(63)
418
tends to infinity as ||Dm || → 0. To overcome this problem in computation, Domnik et al. (2013) and von
419
Boetticher et al. (2016, 2017) introduced the exponential factor my as νse = νs +
τys 1 − e−my ||Dm || . ||Dm ||
(64)
420
To capture typical flow behavior of a geophysical mass flows, Domnik et al. (2013) proposed a pressure
421
dependent yield stress τys = τc + τp pm /Λp , where τc is cohesion. Without cohesion, this forms a Drucker-
422
Prager yield criterion (Prager and Drucker, 1952) q ΠσDm > τp pm ,
(65)
423
where, ΠσDm is the second invariant of the deviatoric stress tensor (Prager and Drucker, 1952). The relation
424
(65) simply tells us that the mixture material undergoes plastic yielding when deviatoric stress is greater than
425
the yield stress. With several complex flow simulations, from silo outlet to the flow obstacle interactions, and
426
depositions with solid-fluid and fluid-solid transitions, Domnik et al. (2013) and von Boetticher et al. (2016,
427
2017) demonstrated that the pressure-dependent yield stress discussed above is a very efficient rheological
428
model.
429
8
430
To solve the equations (25), (28)-(31) numerically, we need boundary conditions. Usually, the free surface of
431
the debris bulk is assumed to be traction free. The friction is induced by the movement of debris mixture
432
on the sliding plane. Consider the normal vector n and a tangential vector t on sliding plane. In rapid
433
flows of debris material down the slopes, even the lowest particle layer in contact with the bottom boundary
434
moves with a non-zero and non-trivial velocity. Therefore, the generally used no-slip boundary condition does
435
not represent the flow physics (Domnik and Pudasaini, 2012). A non-zero slip velocity is determined by the
436
frictional strength, which depends on the load the material exerts on the rigid boundary. So, as in Domnik
437
b = σ b n · t to the normal pressure N b = σ b n · n at the sliding et al. (2013), we relate the shear-stress Tm m m m
Boundary Conditions
24
438
surface (b). This can be achieved by using the Coulomb sliding law for bulk: b Tm =
ubm b tan δm Nm , |ubm |
(66)
439
where the tangent of the effective bed friction angle δm for the mixture defines the “effective proportionality
440
constant”, and ubm is the velocity of the bulk at the base. Consequently, δm ≤ δs implies the emergence of a
441
new effective mixture friction coefficient µm = tan δm . This is uniquely defined via (66), and reveals that µm
442
or, δm , can evolve dynamically. Higher values of tan δm go along with a higher shearing and, therefore, with
443
a higher frictional force at the rigid boundary. The relation (66) asserts that the material yields plastically if
444
the shear stress attains the critical value. The flowing material is assumed to be non penetrating through the
445
boundary. So, n · ubm = 0 and also [(t · ∇)(n · um )]b = 0. Then, from (62), the normal and shear stresses at
446
the bottom are given by b Nm = pbm + 2Λbηm [(t · ∇)(t · um )]b ,
b and Tm = Λbηm [(n · ∇)(t · um )]b ,
(67)
447
which has been expressed in terms of the tangential velocity and the pressure at the boundary, and leads to
448
a pressure and rate-dependent Coulomb-viscoplastic sliding law. Using (67) in (66), one obtains [(n · ∇)(t · um )]b − 2cFm [(t · ∇)(t · um )]b =
449
450
cFm b p = Ferm pbm , Λbηm m
where the ratio between the debris friction and the viscous friction is given by Ferm =
(68) cFm (the effective Λbηm
friction ratio), and the friction factor cFm is defined by b um b |ub | tan δm , um 6= 0, m cF = 0, ubm = 0.
451
The pressure and rate-dependent Coulomb-viscoplastic sliding law for generalized bulk mixture (68), dynami-
452
cally defines the bottom boundary velocity [t · um ]b (Domnik and Pudasaini, 2012). In case, if the bed friction
453
angle δm = 0, then there is no solid material friction, and cFm = 0. This results in the free slip condition
454
[(n · ∇)(t · um )]b = 0, so the basal shear-stress vanishes. For very high shear viscosity, Ferm ≈ 0, and there is
455
no more pressure dependency. So, (68) becomes 2cFm =
[(n · ∇)(t · um )]b , [(t · ∇)(t · um )]b 25
(69)
456
i.e., the friction factor is the ratio of the normal to the tangential velocity at the bottom.
457
We ponder the more general case, Ferm 6= 0. If we consider the flow in a narrow rectangular inclined channel
458
with the main flow direction parallel to the x-axis, the shearing mainly occurs at the bottom boundary with
459
the basal normal vector parallel to the z-direction. The shearing at the sidewalls may not have a significant
460
influence on the flow (Pudasaini et al., 2007; Domnik et al., 2013), and so will be neglected. Then, the
461
Coulomb shear-stress at the bottom (66) can be written as b Tm
=
cFm
"
pbm
+
2Λbηm
∂um ∂x
b #
.
(70)
462
So, we have extended the pressure- and rate-dependent Coulomb-viscoplastic granular flow rheology (Domnik
463
and Pudasaini, 2012; Domnik et al., 2013) to debris mixture flow. Moreover, the extended sliding law (68),
464
for the rectangular inclined channel is given by
∂um ∂z
b
−
2cFm
∂um ∂x
b
=
cFm b p · Λbηm m
(71)
465
The equation (71) is for the pressure in terms of velocity gradient or vice-versa (Domnik and Pudasaini, 2012).
466
So, the pressure dependent velocity is employed as the boundary condition at the sliding surface and the top
467
surface is traction free. Furthermore the bottom shear stress in (71) depends only on the normal stress, i.e.,
468
on the velocity near the sliding surface.
469
We further employ the von Neumann boundary condition for the pressure by applying the momentum con-
470
servation in the direction of the normal at the rigid boundaries, i.e., n · 4pm = n · F·
(72)
471
This closes the derivation of the model equations and the boundary conditions for a generalized quasi two-
472
phase mixture mass flow.
473
9
474
The full dimensional (non-depth-averaged) simulation of the debris mixture flow is a new topic. However,
475
here, we mention some similarities and major differences between our new model and some other seemingly
476
similar or, alternative models existing in the literature. This will make it clear the usefulness of the proposed
Essence of the New Model
26
477
model and its application potential.
478
Since, in general, all the coefficients in (36) are away from unity, even in reduced form, the convective and
479
viscous terms in (26) and (27) are different than in Domnik and Pudasaini (2012) and Domnik et al. (2013).
480
Differences emerge from the fact that all Λ’s may vary as functions of αs ’s and λ’s. Furthermore, (um , wm )
481
and pm are state variables for the mixture, from which, with the knowledge of αs and λ, the state variables
482
for the solid (us , ws ; ps ), and for the fluid (uf , wf ; pf ) can be reconstructed. From the technical point of view,
483
this is a major advantage of the proposed model that could not have been achieved from classical formulations
484
of the mixture models (Iverson and Denlinger, 2001; Pudasaini et al., 2005).
485
As an alternative model for the debris mixture flows, we consider von Boetticher et al. (2016) in which the
486
bulk density is computed as a linear combination of the phase densities from the knowledge of the volume
487
fractions as transport quantities. In our formulation, we do not require to dynamically compute the bulk
488
density. So, we do not have to deal with the difficulties with boundary conditions in simulations with the
489
mixture density gradients and its evolutions as in von Boetticher et al. (2016). This stems from the fact
490
that in von Boetticher et al. (2016), the bulk transport equations are considered whereas we begin with two
491
explicit mass and momentum balances for the solid and fluid phases. Then, we constructed the bulk mixture
492
model consisting of the mass and momentum balances for the mixture. This resulted in the emergence of the
493
velocity and pressure drift factors, inertial and dynamical coefficients for mixture velocities, viscosities and
494
pressures. Our simple formulation provides possibilities for the construction of the full dynamics of the solid
495
and fluid phases. This also allowed to construct several technically important reduced models that can be
496
applied from simple to complex flow situations. The inertial, pressure and viscous terms in (33)-(35) appear
497
in non-classical forms. Effective viscosity takes a very general form. The bulk and shear viscosities, that are
498
in extended and general form, are different. However, the reduced bulk and shear viscosities for the mixture
499
appear to be equivalent. This representative mixture viscosity has further been reduced to simple mixture
500
viscosity as in von Boetticher et al. (2016) that also includes the power law rheology together with the yield
501
strength for the slurry fluid, whereas we model that as a non-Newtonian viscous fluid. In von Boetticher et
502
al. (2016, 2017), the effective mixture viscosity is a complex linear combination between the phase viscosities,
503
whereas in our formulation the mixture viscosities are complex and non-linear combinations between the solid
27
504
and fluid-phase viscosities. The same applies for the velocities and pressures in our model while in the von
505
Boetticher et al. (2016) model the velocities and pressure are simply the bulk velocities and pressure.
506
10
507
The depth-averaged version of the two-phase mass flow model (Pudasaini, 2012) has been successfully applied
508
to different flows of mixture materials. Here, we focused on some particular aspects of the full-dimensional
509
mixture flows down a channel. First, we considered the general three-dimensional and two-phase particle-fluid
510
mixture mass flow model. Then, in order to develop a reduced but, generalized quasi two-phase bulk mixture
511
model, we introduced the solid and fluid velocity drift coefficients, and also the pressure drift coefficient. From
512
the application and computational point of view, this is useful, as it reduces the full two-phase model to quasi
513
two-phase bulk model. This paves the way to apply the simple single-phase computational tools to numerically
514
solve the new model equations. At the same time, the drifts contain important information relating the solid
515
and fluid velocities, and the dynamic pressures. This makes the resulting model a generalized quasi two-phase
516
bulk mixture model.
517
We developed two different mixture viscosities and pressures. The drift induced generalized mixture velocities
518
are obtained. The effective bulk and shear viscosities are obtained by respectively averaging two alternative
519
bulk and shear viscosities appearing in the model derivation. We noticed that the bulk and shear viscosities
520
are identical. This can be considered as the mixture viscosity. The constructed mixture viscosity is rich in
521
its physics and evolves mechanically as a coupled function of several physical and mechanical parameters and
522
dynamical variables including the solid volume fraction, solid and fluid material densities and viscosities, solid
523
friction and yield strength, rate of deformation of solid, and velocity and pressure drift parameters. This
524
shows that the higher are the fluid drifts, faster is the fluid motion, and thus, higher is the deformation. This
525
results in the lower bulk viscosity. In simple situations, the viscosity can be further reduced to the simplest
526
description of the viscosity of the bulk mixture. We have also introduced a general concept of dynamically
527
evolving drift induced generalized pressure for the bulk mixture flow. This shows that the pressure drift
528
coefficient plays a crucial role in the pressure dynamics. We reveal that, during large shearing, dilatation and
529
compaction, the mixture pressure may respond completely differently as compared to the solid or fluid pressure
Discussions and Summary
28
530
separately. Therefore, the dynamics of the bulk pressure is much more complex than just the dynamics of
531
the solid and fluid pressures separately. For dilute flows, it reduces to the classical definition of the mixture
532
pressure. The emergence of an effective mixture friction coefficient, that evolve dynamically, reveals one of the
533
most important mechanical aspects of our mixture model. We have extended the pressure- and rate-dependent
534
Coulomb-viscoplastic granular flow rheology to debris mixture flow. Such a sliding law for generalized bulk
535
mixture dynamically defines the bottom boundary velocity.
536
With the constructed bulk mixture velocities and the pressure, the inertial coefficients, and the coefficients
537
of dynamical mixture viscosities and pressure, our model for bulk mixture flow consists of the full two-
538
dimensional quasi two-phase mass and momentum conservation equations. It appears that the dynamics of
539
the bulk pressure, the inertial and dynamical coefficients are complex and coupled, and explicitly include the
540
physics and dynamics of mixture. The final model equations are written as a well structured system of three
541
non-linear partial differential equations in conservative form with three unknowns, namely, the generalized
542
mixture velocities and pressure. This system is written in a convenient form for numerical simulations.
543
Moreover, the newly derived mass and momentum equations appear in non-conventional form due to the
544
inertial coefficients, and the non-equality of the mixture viscosities.
545
We outlined some important physical aspects, implications, and applicabilities of the new model. The model
546
has a potential for a wider applications, it will be computationally faster than the full two-phase simulations,
547
and, at the same time, more accurate than the single-phase models. The great advantages of the coupled
548
and generalized model is that structurally it is the same as the existing single phase granular or debris flow
549
model. So, in principle, the same computational strategy can be applied to solve this new system. However,
550
the complexity and generality inherited by the two-phase bulk mixture flow are contained in the inertial,
551
dynamical and pressure coefficients. Once mixture velocity and pressure are obtained, the two-phase dynamics
552
of solid and fluid can be re-constructed from the definition of the mixture quantities and drift coefficients.
553
We derived reduced models as characterized either by the volume fraction, or by the drift coefficients. Solid
554
volume fraction appears to be an internal variable for which an evolution equation can be constructed, or
555
simply parameterized. In the most simplified situation, the model equations consist of a single velocity drift
556
parameter. This is advantageous, because there are analytical expressions available for this. Alternatively, 29
557
this factor can be obtained from the fast simulations of the depth averaged equations. This allows us to
558
dynamically obtain the drift coefficient. For simple flow configuration with unit drift coefficients, the emerging
559
model reduces to traditional effectively single-phase bulk mixture model. In special situation when the velocity
560
and drifts are close to unity, the generalized mixture velocities, and pressure reduce to the classical barycentric
561
velocities and pressure. This further reduces the complexity of the model system. For the identical velocity
562
drift coefficients, the generalized viscosities reduce to a single classical mixture viscosity. In the most simple
563
flow configuration, if further the velocity drifts are close to unity, then the mixture viscosity becomes the
564
traditional simple bulk mixture viscosity. Even the simplest set of equations is important because this includes
565
the first order basic dynamics of evolving mixture viscosity that governs the dynamics of the mixture flow.
566
Acknowledgments
567
We thank the reviewers, and the editor Giuseppe Saccomandi for their constructive reviews and suggestions
568
that resulted in the substantially improved paper. The idea of the model development and the model reduction
569
was generated by Shiva P. Pudasaini. We gratefully acknowledge the financial support provided by the German
570
Research Foundation (DFG) through the research project, PU 386/3-1: “Development of a GIS-based Open
571
Source Simulation Tool for Modelling General Avalanche and Debris Flows over Natural Topography”. Puskar
572
R. Pokhrel acknowledges University Grants Commission (UGC), Nepal for the financial support provided as
573
a PhD fellowship.
574
References
575
Ancey C (2007) Plasticity and geophysical flows, A review. J. Non-Newtonian Fluid Mech. 142 (1-3): 4-35
576
Atkin RJ and Craine RE (1976a) Continuum Theories of Mixtures: Applications. J. Inst. Maths Appl., 17,
577
578
579
580
581
153-207 Atkin RJ and Craine RE (1976b) Continuum theories of mixtures: Basic theory and historical development. The Quarterly Journal of Mechanics and Applied Mathematics, 29, 209-244 Bagnold RA (1954) Experiments on a gravity-free dispersion of large solid spheres in a Newtonian fluid under shear. Proc. R. Soc. Lond., Ser. A, 225, 49-63
30
582
583
584
585
Balmforth NJ and Frigaard I (2007) Viscoplastic fluids: From theory to application. J. Non-Newtonian Fluid Mech., 142(1-3), 1-3 Bedford A and Drumheller DS (1983) Recent advances: theories of immiscible and structured mixtures. Int. J. Eng. Sci., 21, 863-960
586
Bingham E (1922) Fluidity and Plasticity. New York, McGrawHill
587
Bohorquez R. de M. P (2008) Study and Numerical Simulation of Sediment Transport in Free-Surface Flow.
588
PhD thesis, M´ alaga University, M´ alaga
589
Bowen RM (1976) Theory of Mixture. A.C. Eringen (Ed.), Continuum Physics, 3, p1
590
Chen CL (1988) Generalized viscoplastic modelling of debris flow. J. Hydraul. Res., 114(3), 237-258
591
Domnik B and Pudasaini SP (2012) Full two-dimensional rapid chute flows of simple viscoplastic granular
592
materials with a pressure-dependent dynamic slip-velocity and their numerical simulations. J. Non-
593
Newtonian Fluid Mech., 173-174, 72-86
594
595
Domnik B, Pudasaini SP, Katzenbach R and Miller SA (2013) Coupling of full two-dimensional and depthaveraged models for granular flows. J. Non-Newtonian Fluid Mechanics, 201, 56-68
596
Drew DA (1983) Mathematical modelling of two-phase flow. Ann. Rev. Fluid Mech., 15, 261-291
597
Drew DA and Lahey RT (1987) The virtual mass and lift force on a sphere in rotating and straining inviscid
598
599
600
flow. Int. J. Multiphase Flow, 13, 113-121 Fernandez-Nieto ED, Bouchut F, Bresch D, Daz MJC and Mangeney A (2008) A new Savage-Hutter type model for submarine avalanches and generated tsunami. J. Comput. Phys., 227(16), 7720-7754
601
Forterre Y and Pouliquen O (2008) Flows of Dense Granular Media. Annu. Rev. Fluid Mech., 40, 1-24
602
GhoshHajra S, Kandel S and Pudasaini SP (2015) Lie symmetry solutions for two-phase mass flows. Interna-
603
tional J. Non-linear Mechanics, 77, 325-341
604
Goddard JD (1996) Dissipative materials as constitutive models for granular materials. Acta Mech., 63, 3-13
605
Hungr O, McDougall S and Bovis M (2005) A real two-phase submarine debris flow and tsunami. Springer,
606
Berlin
607
Ishii M (1975) Thermo-Fluid Dynamic Theory of Two-Phase Flow. Eyrolles, Paris
608
Ishii M and Hibiki T (2006) Thermo-Fluid Dynamics of Two-Phase Flow. Springer, New York
609
Ishii M and Zuber N (1979) Drag coefficient and relative velocity in bubbly, droplet or particulate flows. 31
610
AIChE J., 25, 843-855
611
Iverson RM (1997) The physics of debris flows. Rev. Geo-phys., 35(3), 245-296
612
Iverson RM and Denlinger RP (2001) Flow of variably fluidized granular masses across three-dimensional
613
614
615
terrain: 1. Coulomb mixture theory. J. Geophys. Res., 106(B1), 537-552 Iverson RM (2012) Elementary theory of bed-sediment entrainment by debris flows and avalanches. J. Geophys. Res., 117 (F03006), 197-200
616
Johnson AM (1970) Physical Processes in Geology. Freeman, San Francisco
617
Jop P, Forterre Y and Pouliquen O (2006) A constitutive law for dense granular flows, Coulomb mixture
618
theory. Nature, 441(7094), 727-730
619
Kafle J, Pokhrel PR, Khattri KB, Kattel P, Tuladhar BM and Pudasaini SP (2016) Submarine landslide and
620
particle transport in mountain lakes, reservoirs and hydraulic plants. Annals of Glaciology 57(71) doi:
621
10.3189/2016AoG71A034
622
Kattel P, Khattri KB, Pokhrel PR, Kafle J, Tuladhar BM and Pudasaini SP (2016) Simulating
623
Glacial lake outburst floods with a two-phase mass flow model. Annals of Glaciology, 57(71) doi:
624
10.3189/2016AoG71A039
625
626
627
628
629
630
631
632
633
634
635
636
637
Krasnopolsky B, Starostin A, Osiptsov AA (2016), Unified graph-based multi-fluid model for gas-liquid pipeline flows. Computers and Mathematics with Applications. 72, 1244-1262 Kuo KK, Koo JH, Davis TR and Coates GR (1976) Transient combustion in mobile, gas-permeable propellants. Acta Astron. 3, 575-91 Manninen M, Taivassalo V and Kallio S (1996) On the mixture model for multiphase flow. Espoo, Technical Research Center of Finland, VTT Publications 228 (ISBN 951-38-4946-5; ISSN 1235-0621) Massoudi M (2003) Constitutive relations for the interaction force in multicomponent particulate flows. International J. Non-linear Mechanics, 38, 313-336 Massoudi M (2010) A Mixture Theory formulation for hydraulic or pneumatic transport of solid particles. Int. Jour. Engin. Sci., 11, 1440-1461 Maxey RM and Riley JJ (1993) Equation of motion for a small sphere in a non-uniform flow. Phys. Fluids, 26, 883-889 Mergili M, Fischer JT, Krenn J and Pudasaini SP (2017a) r.avaflow v1, an advanced open source computational 32
638
framework for the propagation and interaction of two-phase mass flows. Geosci. Model Dev.,10, 553-569
639
Mergili M, Fischer JT, Pudasaini SP (2017 b) Process chain modelling with r.avaflow: lessons learned from
640
multi-hazard analysis (WLF4)
641
Mergili M, Emmer A, Juricova A, Cochachin A, Fischer JT, Huggel C and Pudasaini SP (2017 c) Can we
642
simulate complex hydro-geomorphic process chains? The 2012 multi-lake outburst flood in the Santa
643
Cruz Valley (Cordillera Blanca, Peru). Earth Surface Processes and Landforms (ESPL, in Review)
644
Mills M (1966) Incompressible mixtures of newtonian fluids. Int. J. Engng Sci., 4, 97-112
645
O’Brien JS, Julien PY and Fullerton WT (1993) Two-dimensional water flood and mud flow simulation. J.
646
647
648
649
650
651
652
653
654
Hyd. Eng., 119(2), 244-261 Pailha M and Pouliquen O (2009) A two-phase flow description of the initiation of underwater granular avalanches. J. Fluid Mech., 633, 115-135 Pitman EB and Le L (2005) A two fluid model for avalanche and debris flows. Phil. Trans. R. Soc. A, 363(3), 1573-1601 Pitman EB, Patra AK, Kumar D, Nishimura K and Komori J (2013) Two phase simulations of glacial lake outburst flows. J. Computational Science, 4, 71-79 Prager W and Drucker DC (1952) Soil mechanics and plastic analysis or limit design. Q. Appl. Math., 10(2), 157-165
655
Pudasaini SP (2011) Some exact solutions for debris and avalanche flows. Phys. Fluids, 23(4), 043301
656
Pudasaini SP (2012) A general two-phase debris flow model. J. Geophysics. Res., 117. F03010 doi:
657
658
659
660
661
662
663
664
665
10.1029/2011JF002186 Pudasaini SP (2016) A novel description of fluid flow in porous and debris materials. Engineering Geology, 202, 62-73. Pudassaini SP and Fischer J-T (2016) A mechanical model for phase-separation in debris flow. arXiv:1610.03649v1 Pudasaini SP and Krautblatter M (2014) A two-phase mechanical model for rock-ice avalanches. J. Geophys. Res. Earth Surf., 119, doi:10.1002/2014JF003183 Pudasaini SP and Miller SA (2013) The hypermobility of huge landslides and avalanches. Engineering Geology, 157, 124-132 33
666
Pudasaini SP, Hutter K, Hsiau S-S, Tai S-C, Wang Y and Katzenbach R (2007) Rapid flow of dry granular ma-
667
terials down inclined chutes impinging on rigid walls. Phys. Fluids, 19(5), 053302, doi:10.1063/1.2726885
668
Pudasaini SP, Wang Y and Hutter K (2005) Modelling debris flows down general channels. Nat. Hazards
669
Earth Syst. Sci., 5, 799-819
670
Rajagopal KR and Tao L (1995) Mechanics of mixtures. 754, Singapore, World scientific
671
Richardson JF and Zaki WN (1954), Sedimentation and fluidization: Part 1. Trans. Inst. Chem. Eng., 32,
672
673
674
675
676
677
678
35-53 Rivero M, Magnaudet J and Fabre J (1991) Quelques resultants nouveaux concernat les forces exercees sur une inclusion spherique par unecoulement accelere. C. R. Acad. Sci., Ser. II, 312, 1499-1506 Savage SB and Hutter K (1989) The motion of a finite mass of granular material down a rough incline. J. Fluid Mech., 199, 177-215 Schaefer T (2014) Viscosity spectral function of a scale invariant non-relativistic fluid from holography. Phys. Rev., 90, 106008
679
Takahashi T (2007) Debris Flow: Mechanics, Prediction and Countermeasures. Taylor and Francis, New York
680
Tardos GI (1997) A fluid mechanics approach to slow,frictional flow of powders. Powder Technol., 92(1), 61-74
681
Truesdell C (1957) Sulle basi della thermomeccanica. Rand Lincei, Series. 8(22), 33-38, 158-166
682
Truesdell C (1984) Rational thermodynamics Springer-Verlag
683
von Boetticher A, Turowski JM, McArdell BW, Rickenmann D and Kirchner JW (2016) DebrisInterMixing-
684
2.3: a finite volume solver for three-dimensional debris-flow simulations with two calibration parameters
685
- Part 1: Model description. Geosci. Model Dev., 9, 2909-2923, doi: 10.5194/gmd-9-2909-2016
686
von Boetticher A, Turowski JM, McArdell BW, Rickenmann D, H¨ urlimann M, Scheidl C and Kirchner JW
687
(2017) DebrisInterMixing-2.3: a finite volume solver for three-dimensional debris - flow simulations with
688
two calibration parameters - Part 2: Model validation. Geosci. Model Dev. Discuss., doi:10.5194/gmd-
689
10-2017-25 (in Review)
690
Wallis GB (1969) One-dimensional two-phase flow. McGraw-Hill, New York, 408
691
Zuber N and Findlay J (1965) Average volumetric concentration in two-phase flow systems. J. Heat Transfer,
692
87, 453-468
34