Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

Contents lists available at ScienceDirect

Journal of Non-Newtonian Fluid Mechanics


journal homepage: www.elsevier.com/locate/jnnfm

Emerging, ripening, and attenuating stages of granular roll waves


Jianbo Fei a, b, d, Yuxin Jie c, Zezhou Wu b, *, Min Zhu a, b, d
a
Underground Polis Academy, Shenzhen University, Shenzhen 518060, China
b
College of Civil and Transportation Engineering, Shenzhen University, Shenzhen 518060, China
c
State Key Laboratory of Hydroscience and Engineering, Tsinghua University, Beijing 100084, China
d
Key Laboratory of Coastal Urban Resilient Infrastructures (MOE), Shenzhen University, Shenzhen 518060, China

A R T I C L E I N F O A B S T R A C T

Keywords: Adopting the form of the dynamic μ(Fr,h) basal friction law and introducing a second-order viscous term into the
Granular flow Saint Venant-type equations, we simulated two-dimensional granular roll waves generated in a rectangular chute
Roll waves applying the second-order-accurate total-variation-diminishing MacCormack scheme. Consistent with previous
Depth-averaged
findings, we found that the amplitude and wavelength of the wave increased with transit distance. Although the
Viscous effect
μ(Fr,h) dynamic basal friction law is the more influential factor in shaping the waves, introducing a viscous term
slightly reduces the wave peaks. A sensitive analysis indicates a quadratic relationship between wave peak height
and the coefficient in the depth-averaged viscosity with ν≤2.4 × 10− 2 m1.5/s. In the numerical simulation of roll
waves subject to initial perturbations of different magnitudes, three stages in their evolution (namely, emerging,
ripening, and attenuating) are evident from their difference in transit behavior. In the emerging stage, the
normalized maximum height of the roll waves increases exponentially during propagation. In the ripening stage,
this height grows linearly with transit distance whereas the growth rate decreases with the strength of the initial
perturbation. During the attenuating stage, the maximum roll-wave height forms a hump in the upstream region
where the waves attenuate. Roll-wave peaks and troughs for all three stages tend to approach equilibrium heights
further down the chute. Subcritical flow roll-wave simulations show that as flow speed decreases, a reverse
transformation pattern is found; specifically, the amplitude and wavelength of the wave decrease further down
the chute.

1. Introduction flows, i.e., the μ(I) rheology. The rheology initially correlates the shear
stress and confined pressure [7] and further links the strain rate and
In classical fluids, unstable free-surface flows and long waves ulti­ deviatoric stress [8].
mately form if the flows accelerate to a certain velocity. This instability The μ(I) rheology furnishes the continuum governing equations (e.g.,
is usually referred to as the “Kapitza instability” for viscous fluid and Navier–Stokes (N–S) or Saint Venant equations) with terms to describe
“roll waves” for turbulent flow. The instability is not only experimen­ the rapid motion of granules more accurately. The introduction of μ(I)
tally observed in hydraulic open channel flows, but also in complex rheology into the depth-averaged granular flow model (Saint Venant-
flows such as mud flows, grain suspension flows, and dry granular flows. type) generated two different longitudinal viscous stresses as reported
Savage [1] and Vallance [2] noticed successive surges in chute granular by Forterre [9] and Gray and Edwards [10], which enabled the calcu­
flow experiments. Daerr [3] reported the instability of granular flow at lation of the cutoff frequency for the instability. Within the
the rear of the front of laboratory avalanches. Forterre and Pouliquen depth-averaged framework and the inclusion of μ(I) rheology, roll waves
[4] found that the uniform granular flow on an inclined plane becomes were studied analytically in Gray and Edwards [10]. Later on,
unstable in the experiment if a threshold of the Froude number Fr=0.57 one-dimensional μ(I)-rheology-based depth-averaged equations were
is exceeded. They also calculated Fr=2/3 for the theoretical threshold of numerically solved by Razis and co-workers to simulate roll waves
the Kapitza instability by adopting the depth-averaged avalanche dy­ adopting a shock capturing central scheme [11], Edwards and Gray
namic model [5] and the μ(Fr,h) basal friction law [6]. μ(Fr,h) friction [12], and Viroulet and co-workers [13]. Using two-dimensional N–S
law further evolves to a phenomenological constitutive law for granular equations coupled with μ(I) rheology as a framework without

* Corresponding author.
E-mail address: wuzezhou@szu.edu.cn (Z. Wu).

https://doi.org/10.1016/j.jnnfm.2020.104411
Received 1 April 2020; Received in revised form 5 August 2020; Accepted 6 October 2020
Available online 9 October 2020
0377-0257/© 2020 Elsevier B.V. All rights reserved.
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

depth-integration, the governing equation was solved numerically by Froude number is expressed as
Baker and co-workers [14] when simplified to two-dimensional steady
|u|
fully-developed flows. Nevertheless, roll waves were not produced Fr = √̅̅̅̅̅̅̅̅̅̅̅̅̅̅, (3)
ghcosθ
because the μ(I) rheology was ill-posed. In latter research using the
governing framework without depth-integration, granular roll waves In Eq. (1b, c), Vx and Vy are two viscous-like terms. Gray and
were simulated numerically by setting an initial perturbation with Edwards [10] and Baker and Gray [15] derived mathematical expres­
partially regularized μ(I) rheology by Barker and Gray [15]. sions for these terms by including the gradient of the in-plane shear
In general, the above-mentioned studies on granular roll waves were stresses with the μ(I)-rheology, specifically,
mostly based on the analysis or simulation using a one-dimensional ( ) ( ( ))
depth-averaged governing equation or a two-dimensional N–S equa­ ∂ νh3/2 ∂∂ux ∂ 12 νh3/2 ∂∂uy + ∂∂vx
tion. The present paper conducted a two-dimensional depth-averaged Vx = + , (4a)
simulation on granular roll waves. Thus, the three-dimensional config­ ∂x ∂y
uration of roll waves can be illustrated. In the previous studies on the roll ( ) ( ( ))
waves modelling with depth-integration, periodic-box simulations that ∂ νh3/2 ∂∂vy ∂ 1
2
νh3/2 ∂u
∂y + ∂∂xv
connect the upstream and downstream sections were conducted by Razis Vy = + . (4b)
∂y ∂x
and co-workers [11] and Edwards and Gray [12]. For our study, linear
uniform granular flow in a long flume was investigated. The coordinates Recall that the basal friction law [i.e., Eq. (2)] was proved to be
in our simulation were fixed, in contrast to the wavefront-centered important in determining the configuration of the granular roll waves
traveling coordinates used by Gray and Edwards [10]. Thus, contin­ [10,11,12]. In comparison, the magnitude of each viscous term [i.e., Eq.
uous evolutionary behaviors of roll waves can be studied, which closely (4a, b)] is relatively small [13], but all terms contribute an important
resembles the natural granular flow cases. In an attempt to model roll part in predicting the cutoff frequency and wavenumber [10,15] and
waves more accurately, μ(I)-rheology-based viscous depth-averaged coarsening dynamics [13].
equations reported by Gray and Edwards [10,12] was adopted as the
basic theoretical framework to describe granular roll waves. With the 3. Numerical scheme and experiment set-up
adoption of a second-order accurate finite-element method (i.e.,
TVD-MacCormack scheme), two-dimensional uniform granular flows in 3.1. Numerical scheme
an inclined chute were able to be simulated more precisely, with which
the mechanism of steady granular roll waves can be investigated more The traditional first-order governing equations of continuum dy­
readily. namics models for avalanches, i.e., partial differential equations of mass
and momentum conservation, can be efficiently solved using various
2. Governing equations numerical simulation methods, e.g., finite-element methods. Several
traditional numerical schemes have been adopted to solve hyperbolic-
Consider a stream of granules flowing down an inclined rectangular type partial differential equations, e.g., the first-order upwind scheme
chute. When the depth of the granular flow h is much smaller than its and second-order upwind beam-warming and Lax–Wendroff schemes. A
lateral spread in an inclined plane, a depth-averaged continuum model Eulerian scheme similar to the MacCormack scheme [17] was adopted
for granular flow can be implemented by integrating the three- by Savage and Hutter [18] and Lagrangian finite-difference schemes by
dimensional N–S equations along the depth of the flow. In rectangular Anderson and co-workers [19], Greve [20], and Savage and Hutter [21].
Cartesian coordinates Oxy, we have These methods may cause dispersity that leads to large oscillations
during calculations, and fail to capture shock wave phenomena accu­
∂h ∂(hu) ∂(hv) rately. A high-resolution shock-capturing numerical scheme is needed to
+ + = 0, (1a)
∂t ∂x ∂y simulate the moving front and rolling waves accurately. Since Nessyahu
( ) and Tadmor [22] proposed the second-order-accurate non-oscillatory
( 2
) ∂ 1
gh2 cosθ ( ) central total-variation-diminishing (TVD) scheme, research on central
∂(hu) ∂ hu ∂(huv) 2 u
+ + + = ghcosθ tanθ − μb + V x , TVD schemes has become intense with modifications made in successive
∂t ∂x ∂y ∂x |u| studies [23–25], and the scheme has been recognized as an effective and
(1b) commonly-used approach. The central TVD schemes have already been
( ) applied to simulate the movement of dense and shallow granular
( 2
) ∂ 1
gh2 cosθ “liquid”. Gray and co-workers [26] adopted the TVD Lax–Friedrichs
∂(hv) ∂(huv) ∂ hv 2 v
+ + + =− ghcosθμb + Vy , (1c) scheme to simulate some typical phenomena when a granular flow
∂t ∂x ∂y ∂y |u|
striking an obstacle, such as shock waves, vacuum and dead areas.
Following Gray and co-workers [26], a non-oscillatory central difference
where θ denotes the angle of inclination, u and v the depth-averaged
scheme integrated with the TVD limited or weighted essentially
flow velocities along the x- and y-coordinates, u = (u2 + v2 )1/2 the non-oscillatory cell formation approach was introduced by Tai and
magnitude of the velocity field tangential to the flow base, and g the co-workers [27] to model shock waves and get a smooth solution. To
acceleration due to surface gravity. Rather than a traditional Coulomb- model the movement of landslides on natural terrain, Pitman and
type basal friction law, the coefficient of dynamic basal friction μb(h,Fr) co-workers [28] developed an adaptive mesh and parallel Godunov
is adopted in this study to simulate roll waves according to Pouliquen solver. Denlinger and Iverson [29] applied the Riemann solver algo­
and Forterre [4], Baker and Gray [15] rithm to model granule-laden liquid flows. The shock-capturing central
μ2 − μs scheme [24] with a Runge–Kutta–Chebyshev adaptive time stepper [30]
μb (h, Fr) = μs + , (2)
βh
λFr
+1 and weighted essentially non-oscillatory flux limiter [22,31] has been
applied to simulate roll waves [11], erosion-deposition waves [12], and
where β denotes a dimensionless empirical constant, with β=0.65 ac­ segregation-induced fingering [32].
cording to Pouliquen and Forterre [16] and Forterre and Pouliquen [4], We shall next use the mathematical framework described above, i.e.,
λ the length scale, and μ2 and μs denoting the threshold and start-up depth-averaged governing equations coupled with the dynamic friction
coefficients of friction, which correspond to the maximum and mini­ law and the second-order viscous terms, to simulate a uniform depth of
mum angle of inclination at which steady uniform flows occur. The granules flow in a long inclined rectangular chute. Note that the

2
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

inclusion of the depth-averaged viscous terms is a singular perturbation Note that granules slide downstream a chute without transversal
of the model and turns it from a hyperbolic system to a hyperbolic- inclination is chosen as a scenario, an initial inflow rate in the x- di­
parabolic system of equations. Thus, a high-resolution shocking rection is set while that in the y- direction is zero. In addition, the
capturing scheme is essential for solving the convection-diffusion transversal dimension (along the y- direction) of the chute flow is much
equations numerically. The total variation diminishing (TVD)–Mac­ smaller than the longitudinal (along the x- direction) one. Thus, the flow
Cormack scheme was adopted [33–34], and a fast and precise program in the downstream direction prevails during the whole flow process, the
developed to perform time-dependent numerical simulations for roll mass conservation holds along the x- direction in the simulation.
waves. The scheme uses the two-step (predictor and corrector) Mac­ Knowing that erosion–deposition waves were generated by initial­
Cormack scheme as a basic framework, and a TVD term is introduced to izing a sinusoidal or random perturbation to a steady uniform flow by
eliminate steep-gradient oscillations, ensuring the scheme is advanta­ Edwards and Gray [12], these two types of zero-mean inflow pertur­
geous over other single-step schemes, e.g., the Lax–Wendroff and clas­ bations are also introduced in the present study to trigger roll waves in
sical explicit Euler schemes. The second-order viscous terms are the simulation through the depth-averaged governing equations, i.e., Eq.
discretized by the central difference scheme during the numerical (1a–c). The frequency and amplitude of the oscillatory perturbation
implementation. The scheme provides second-order spatial and tempo­ were set to 0.47 Hz and 0.1 mm. To set the random perturbations, values
ral accuracy of the calculation program during the calculation. Details of were taken from within the interval [− 0.05 mm, 0.005 mm]. In the
the numerical scheme are provided in Appendix A. simulations, the perturbation was applied via the inflow volumetric flux.
Interestingly, the simulated results with sinusoidal and random pertur­
bations were found to be almost identical. Therefore, we only present
3.2. Initial conditions and parameter and boundary perturbation settings
results pertaining to sinusoidal perturbations.
The set-up of the numerical simulation example begins with the
setting of the initial conditions, simulation parameters, and boundary 4. Analysis of the simulation results
perturbation. To simulate a stream of steady granular flow in an inclined
chute at the angle of θ=35.1◦ (see Fig. 1), the flow thickness is set at h0= In the periodic-box simulations conducted by Gray and Edwards
4.15 mm. We selected a 5-m-long section of the 0.45-m-wide chute for [10], it was reported that waves with large amplitude traveled faster
analysis. than those with small amplitude. This is consistent with our simulation
Razis and co-workers [11] and Edwards and Gray [12] defined the results: tracked flow depths at x=2.5 m and x=4.5 m (Fig. 2) and the
steady state of the granular layer flowing with constant speed and sequences of profiles for various wave configurations (Fig. 3) both
thickness. The inflow velocity and the initial velocity in the x-direction is indicate that the steady amplitude of the wave height increases with
determined by the flow depth, angle of inclination, and empirical pa­ distance transited. In addition, consistent with experimental observa­
rameters, that is, tions by Viroulet and co-workers [13], computed roll waves of large
√̅̅̅̅̅̅̅̅̅̅̅̅ amplitude had longer wavelengths; see comparison (Fig. 3b) of the
u0 =
β gcosθ 3/2
h0 , (5) distance between adjacent wave peaks of different colors. Apart from
λγ these already recognized findings, we further analyzed the simulation
results by changing the parameter settings of simulations, e.g., the co­
where λ=1 mm is used, and γ can be determined from efficient in the depth-averaged viscosity, the initial perturbations, and
μ2 − μs the initial flow condition.
γ= − 1, (6)
tanθ − μs

with μs=0.74 and μ2=0.827 being used. The value of coefficient in the 4.1. Effect of the μ(I)-rheology-based viscous terms
depth-averaged viscosity in the viscous-like terms in Eqs. (4a, b) were
obtained from In studying the influence of the μ(I)-rheology-based viscous term on
√̅̅̅ the modelling of roll waves, we solved Eq. (1a–c) with ν=2.4 × 10− 3
2λγ gsinθ
ν= √̅̅̅̅̅̅̅̅̅ . (7)
9β cosθ
In addition, natural boundary conditions for the flow velocity were
specified at the outflow boundaries. The free-slip wall boundary con­
dition was chosen for the two side boundaries. In the calculation along
the y-direction, natural boundary conditions for flow depth were applied
to the two side boundaries, i.e., zero-gradient flow depth was specified
along the confining side walls of the chute.

Fig. 1. Schematic of granular flow down an inclined rectangular chute used for
roll-wave simulations along with the chute dimensions and the Cartesian co­ Fig. 2. Comparison of two evolutions of the simulated avalanche depth with
ordinate frame. viscous term at x=2.5 m and x=4.5 m.

3
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

term, and indicates that in a single period the perturbation grew until
stable waves were attained.
To investigate this viscous effect further, we conducted quantitative
sensitive analysis on the coefficient in the depth-averaged viscosity.
Fig. 5 illustrates the tracked wave peak height at x=2.5 m and x=4.5 m
with ν=2.4 × 10− 4 m1.5/s, 5.0 × 10− 4 m1.5/s, 2.4 × 10− 3 m1.5/s, 5.0 ×
10− 3 m1.5/s, 1.0 × 10− 2 m1.5/s, and 2.4 × 10− 2 m1.5/s, and indicates
that the steady maximum height decreased with increasing value of the
coefficient in the depth-averaged viscosity at the two positions. Com­
bined with the calculated results from ν=1.5 × 10− 2 m1.5/s and 2.0 ×
10− 2 m1.5/s, the correlation between the maximum wave depth and the
coefficient in the depth-averaged viscosity (Fig. 6) indicates a quadratic
relationship when ν≤2.4 × 10− 2 m1.5/s. The physical origin of the
simulation results can be interpreted by analyzing the energy trans­
formation of the dynamic system: as the dissipation of energy raises with
the growth of viscosity, the wave peak height that is positively corre­
lated to the energy diminishes with increasing viscosity. To interpret the
mathematical quadratic law between the wave peak height and the

Fig. 3. Sequences of profiles of simulated rolling waves at intervals of 0.2 s


represented by different colors: (a) without and (b) with the viscous term.

m1.5/s and ν=0, respectively. Note that Eq. (1a–c) with ν=0 reduces to
that proposed by Forterre and Pouliquen [4]. In Fig. 3, the wave peaks
were suppressed with the introduction of the viscous term, which adds a
diffusive effect to the governing equation. We also conclude from Fig. 3
that the effective basal friction law plays a more significant role in
shaping the configuration of waves when the viscous term is present.
Nevertheless, the inclusion of a viscous term is essential considering it
generates roll waves in the y-direction, which can be seen by the
thickness of the colored outlines [Fig. 3(a, b)] and a comparison of the
cross-section configurations (Fig. 7). Fig. 4 illustrates the evolution of
the calculated avalanche depth at x=4.5 m with and without the viscous

Fig. 4. Comparison of the evolution of the calculated avalanche depth at x=4.5 Fig. 5. Evolution of tracked wave peak height at x=2.5 m (a) and x=4.5 m (b)
m with and without the viscous term. for different coefficients in the depth-averaged viscosity.

4
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

1
h0 u0 2 − 2h0 u0 uw + h0 uw 2 − gcosθ(hw + h0 )hw = 0, (9)
2

in which hw and uw are the wave peak height and wave speed, respec­
tively. h0 and u0 are the flow thickness and velocity on the forward side
of the shock, respectively; they are deemed to obey Eq. (9), i.e.,

νh0 1/2
u0 ∼ . (10)
L
The introduction of Eq. (10) into Eq. (9) derives another quadratic
equation for the coefficient in the depth-averaged viscosity. This is likely
to be the reason why the computed wave peak scales quadratically with
the viscosity.
The calculated cross sections and profiles of the roll-wave configu­
Fig. 6. Correlation between the wave peak at x=4.5 m and the coefficient in ration for different values of the coefficient in the depth-averaged vis­
the depth-averaged viscosity indicating a quadratic relationship when ν≤0.024 cosity (Figs. 7 and 8, respectively) show that the roll wave along the y-
m1.5/s. direction was more visible because the ratio of the wave amplitudes in
the x- and y-directions grows with increasing value of the coefficient in
( )
the depth-averaged viscosity. This is likely because the flowing mass
∂ νh3/2 ∂∂ux
with larger viscosity can sustain greater cross-stream gradient, consid­
viscosity, we firstly obtainhu⋅u ∼ νh3/2uL deduced from ∂(hu)
∂t ∼ ∂x ering the depth gradient term is influenced by the viscous term with
by referencing the momentum conservative equation Eq. 1(b) (where L reference to Eq. (1c). The results indicate that the application of natural
is the characteristic length scale in the x- direction). This expression boundary conditions combined with the effect of viscous term produce
further evolves to side-wall viscous drag-like effect on the flow in our simulation: the effect
νh1/2 becomes more obvious with the increase of the coefficient in the depth-
u∼ . (8) averaged viscosity. In general, the introduction of viscous-like term into
L
the μ(Fr,h)-law-based depth-averaged avalanche dynamic model en­
With the introduction of the Rankine–Hugoniot condition [35] for ables the simulation of cross-stream gradients that can be observed in
the mass and momentum conservation across a wave, Razis and natural granular roll waves cases.
co-workers [11] derive a quadratic equation for the wave speed

Fig. 7. Cross sections of five consecutive sequences of a calculated roll wave configuration at intervals of 0.2 s. Different coefficients in the depth-averaged viscosity ν
are used for comparison: (a) 5.0 × 10− 4 m1.5/s, (b) 2.4 × 10− 3 m1.5/s, (c)5.0 × 10− 3 m1.5/s, and (d) 2.4 × 10− 2 m1.5/s.

5
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

3
Fig. 8. Resulting profiles of the calculated roll wave configurations at intervals of 0.2 s represented by different colors with (a) 5.0 × 10− m1.5/s, and (b) 2.4 × 10− 2

m1.5/s.

4.2. Emerging, ripening, and attenuating stages exponentially; to be discussed further in Section 3.3). (2) The ripening
stage was calculated with an intermediate strength for the initial per­
We set different initial perturbations and check the resulting effects. turbations (specifically, 0.415 mm < pi ≤1.0 mm), the peak height grew
From another viewpoint, waves with different initial perturbations as the wave proceeded downstream until reaching an equilibrium
correspond to different positions and timepoints; that is, in the down­ height. The growth rate increased initially and subsequently decreased
stream region, larger initial perturbations represent mature waves. The to a constant value. (3) The attenuating stage was produced with initial
tracked flow depth at x=4.5 m with different magnitudes of initial perturbations of large strengths (i.e., 2.0 mm≤pi≤5.0 mm). A hump in
perturbations (Fig. 9) indicate that the steady wave peak height in­ the bound of the wave peak height appears in the upstream region. This
creases with the strength of the initial perturbation. indicates that the maximum height of the waves increased rapidly
In addition, 14 independent numerical trials with different pertur­ initially reaching a maximum value, and then attenuated, the peak
bation strengths were conducted. Note that the magnitude of initial height decreasing to an equilibrium value. Note that although higher
perturbation pi is restrained (i.e., pi ranges from 0.05mm to 5.0mm); an humps were observed with initial perturbations of increased strength,
initial perturbation pi larger than 5.0 mm causes a calculation failure in the attenuating waves converged to an identical equilibrium value.
our simulation. Boundaries for the peak and trough heights of the roll The prediction from Fig. 10 is that the wave in all three stages tended
waves at different time points during the whole process of wave evo­ to adjust its peak height to an identical equilibrium peak height (h’~15
lution were outlined (transformed from the upper and lower boundary mm); if the amplitude of the wave is larger than this value, it autono­
lines) and separately plotted (Figs. 10 and 11). These two diagrams mously reduces its height to an equilibrium value while propagating. In
assist in our understanding of the physics of the wave transition. In the contrast, if the amplitude is smaller than this value, the wave grows
diagrams, h’ denotes the wave peak height above the initial flow depth instantaneously until the equilibrium value is reached. Note that waves
h0=4.15 mm in Fig. 10 and below the depth in Fig. 11. in the emerging stage cannot reach the equilibrium peak height within
Regarding Fig. 10, we separated the outlined boundary lines of the the calculation domain, but outside the calculation domain. The
peak wave height into three stages: emerging, ripening, and attenuating magnitude of initial perturbation is restrained; an initial perturbation pi
stages. (1) The emerging stage was obtained with small initial pertur­ larger than 5.0 mm causes a calculation failure in our simulation.
bations (i.e., pi≤0.3 mm for this study). The wave peak height increased In addition, an equilibrium height of the wave trough was also
as the wave proceeded downstream at a growing rate, i.e., observed (Fig. 11). If the initial perturbation is smaller than the

6
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

4.3. Normalized wave peak height versus transit distance in the emerging
and ripening stages

To account for the configuration of waves peak in the emerging and


ripening stages following distinctive shapes but similar trends, we
introduced the difference between wave peak height and the initial flow
depth, denoted by h’, normalized by the magnitude of the initial
perturbation, denoted by h*, and fitted the calculated data h* against the
transit distance x for different initial perturbation strengths (Figs. 12 and
13). In the emerging stage (Fig. 12; pi =0.05 mm, 0.1 mm, 0.15 mm, 0.2
mm, 0.25 mm, and 0.3 mm), the maximum peak height grows expo­
nentially with transit distance and obeys the same fitting correlation,
( )
h∗ = 0.4e0.5197x , R2 = 0.9877 . (11)

Fig. 13 indicates that, in the ripening stage (pi =0.415 mm, 0.5 mm,
0.7 mm, and 1.0 mm), the wave peaks roughly grow linearly with transit
distance from the same original point (x=0, h*=0.4), and the growth
rate decreases with the initial perturbation strength.
These findings are consistent with the calculated transient evolution
pattern of the free surface of roll waves reported by Barker and Gray
[15], i.e., an initial exponential trend and later linear.

4.4. Comparison of numerical data with the theoretical prediction

Fig. 9. Evolution of the tracked height of three stable roll waves with different With the introduction of the Rankine–Hugoniot conditions [35] for
magnitudes of the initial perturbation at position x=4.5 m in the chute. mass and momentum conservation across the shock, as well as the
assumption of the flow thickness on the rearward side of the shock h- as
equilibrium value, the wave trough gradually deepens when travelling its peak height hw (i.e., h-≈hw), the wave speed uw of a shock can be given
downstream. However, if the initial perturbation is larger than this as [11]
value, the wave attenuates, and its wave trough becomes shallower (the √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
1
depth of wave trough from the initial flow height decreases) when uw (hw ) = u+ + gcosθ[1 + (hw /h+ )]hw , (12)
travelling downstream until the equilibrium trough height is reached. 2
These findings are consistent with the existence of bounds for the in which, h+ and u+ are the flow thickness and the depth-averaged ve­
wave peak and trough reported by Razis and co-workers [11]. If the locity on the forward side of the shock, respectively (Fig. 14). Eq. (12)
wave peak or trough exceeds respective equilibrium heights, the gran­ can be further simplified by linearizing it to the first order in (hw-h0),
ular mass has to be distributed over the next several roll waves. The neglecting higher order terms that are relatively much smaller in
existence of the equilibrium value is significant because it indicates that magnitude and approximating depth on the forward side of the shock
both peaks and troughs of the waves are unable to grow indefinitely with h+ as the uniform flow depth h0 [11]
large.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ 3 √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
uw (hw ) ≈ u0 + gh0 cosθ + gcosθ/h0 (hw − h0 ). (13)
4
To check the validity of our numerical simulation results, the

Fig. 10. Profiles of the boundaries for wave peaks generated by different strengths of the initial perturbation, emphasizing the three distinct stages: emerging
(marked with dashed-dotted lines), ripening (solid lines), and attenuating (dashed lines).

7
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

Fig. 11. Boundaries for the trough of roll waves with different strengths of the initial perturbation. Trough of waves significantly turn lower with the increase the
initial perturbation at the original point. They converge to an equilibrium value with increasing position in the chute (x-coordinate).

Fig. 12. In the emerging stage, ratio of wave peak height to the magnitude of the initial perturbation against position in the chute (x-coordinate) with different initial
perturbations pi=0.05 mm, 0.1 mm, 0.15 mm, 0.2 mm, 0.25 mm, and 0.3 mm. Black dots are the captured data points for the upper boundary position, blue line is a
fitted curve.

calculated data of uw and hw for each shock are measured at four time­ the variations of wave speed (i.e., Δuw=uw(i)-uw(j)) and wave peak height
points (i.e., t=20 s, 22 s, 24 s, and 26 s). The magnitude of uw for each (i.e., Δhw=hw(i)-hw(j)) at two adjacent locations (i.e., i and j) are more
wave is obtained by measuring the travel distance of its peak during a conspicuous than these between uw and hw. We plotted the computed
short period of computation time. As the scale of wave amplitude is data of Δuw and Δhw at t=20 s, 22 s, 24 s, and 26 s in the Δuw -Δhw plane
relatively smaller than that of the flow depth, the correlation between in Fig. 15, and compared these data with the following theoretical

8
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

Fig. 13. In the ripening stage, boundaries of the trough of roll waves generated with different magnitudes of the initial perturbation.

indicates that the wave amplitude decreases to an equilibrium value


soon after a single cycle of the wave. Considering these findings, we
concluded that slower flows with smaller Froude number tend to form
decreasing roll waves with smaller amplitudes when propagating.

5. Conclusion

Within the depth-averaged continuum framework, we adopted the


μ(Fr,h) basal friction law [6] and the μ(I)-rheology-based second-order
viscous term of Gray and Edwards [10] in conducting simulations of
granular roll waves generated in a rectangular chute. One main inno­
vation of this study is that the simulations were performed in two di­
mensions using a second-order accurate TVD–MacCormack scheme. As a
consequence, roll waves were simulated precisely, enabling their phys­
Fig. 14. Schematic of the flow depth on the forward and rearward side of an ical properties to be revealed more readily. A scenario whereby linear
inviscid shock (h+ and h-, respectively), and the velocity on the forward side u+.
uniform granular flow in a long and inclined flume was selected for
simulation, in which a fixed coordinate frame was used rather than a
prediction that was derived from Eq. (13) wavefront-centered traveling coordinate frame.
3 √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ From simulation results, the amplitude and wavelength of the waves
Δuw (hw ) = gcosθ/h0 ⋅Δhw . (14) increased with transit distance. The diffusion effect introduced by the
4
viscous term reduced the maximum height of the waves slightly, but the
It can be concluded from the numerical calculation results in Fig. 15
dynamic μ(Fr,h) basal friction law was still the main factor that
that the tracked data of Δuw are linearly related to those of Δhw, and the
contributed to shaping the configuration of the waves. Sensitive analysis
numerical data fit well with the theoretical solution (i.e., Eq. (14)). This
quantitatively revealed a quadratic relationship between the wave peak
proves the validity of the simulation in this paper.
height and the coefficient in the depth-averaged viscosity when ν≤2.4 ×
10− 2 m1.5/s. The ratio of the wave amplitude in the x-direction to that in
4.5. Simulation of subcritical flow the y-direction was found to increase with the value of the coefficient in
the depth-averaged viscosity.
The above-mentioned results on granular roll waves are based on the From results of numerical simulations with initial perturbations with
simulations of critical granular flows (with Froude number larger than different magnitudes, three stages in the propagation of roll waves were
1). To simulate granular roll wave behavior in the subcritical regime, we identified—emerging, ripening, and attenuating stages—and they were
decreased the initial flow velocity and adjusted the corresponding flow found to obey different transit trends. The peak height of the roll waves
depth h0 to 3.46 mm [Eq. (7)]. Therefore, the Froude number satisfies in the emerging stage increased exponentially during this transit period.
Fr=0.63<β<1 (β the Kapitza instability threshold); settings of the other Roll waves during the ripening stage grew at an increasing rate initially
parameters remained the same as for the simulation of critical flow (see and later at a decreasing rate until an equilibrium height was reached.
Sections 3.1 and 3.2). For roll waves during the attenuation stage, a hump in the bound of the
The stable amplitude and the wavelength of the waves corresponding maximum wave height was observed in the upstream region, and when
to the subcritical flow regime decreased during wave propagation transiting downstream the waves gradually attenuated reaching an
[Fig. 16(a)]. The tracked granular flow depth at x=4.5 m [Fig. 16(b)]

9
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

Fig. 15. Comparison between the measured data of Δuw and Δhw and the theoretical solution for Eq. 14.

for the wave peak and trough, indicating that the height of the wave
peak and trough tended to converge onto respective values when trav­
elling downstream.
In the emerging stage, the normalized wave peak height was found to
grow exponentially with transit distance following the same correlation
law. In contrast, in the ripening stage, the normalized wave peak height
grew linearly and the growth rate decreased with the strength of the
initial perturbation. We also performed simulations on roll waves of
subcritical flows by slowing down the flow speed and found that the
amplitude of the stable waves and their wavelengths decreased in
transiting downstream.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgements

The research is funded by the National Natural Science Foundation of


China (NSFC) under grant No. 52008261 and 41790434, and the Key
Research and Development Program of China Railway (Grant No.
K2019G033).

Fig. 16. Simulation results of subcritical flow when Froude number is less than
1:(a) amplitude of a roll wave decrease with position in the chute (x-coordi­
nate); (b) evolution of the tracked avalanche depth at x=4.5 m.

equilibrium height. In addition, equilibrium heights were also observed

10
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

Supplementary materials

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.jnnfm.2020.104411.
Appendix A. Implementation of the TVD-MacCormack numerical scheme

To explain the numerical calculation process, the governing equations (i.e., Eq. 1) can be expressed as
→ →
∂Q ∂ f ∂→g → → → →
+ + = S + H + V x + V y, (A.1)
∂t ∂x ∂y

→ → → →
in which Q is conservative variables vector, f and → g are the flux vectors along the x- and y-directions, respectively, S and H the source terms, and
→ →
V x and V y the diffusive flux. More specifically, the vectors can be expressed as
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
h hu hv
→ ⎝ ⎠ → ⎝ 2/ /
2
Q = hu , f = (hu) h + gh cosθ 2 , g = ⎠ → ⎝
/ huv /
⎠ (A.2a,b,c)
hv huv (hv)2 h + gh2 cosθ 2
⎛ ⎞ ⎛ ⎞
0 0
→ ⎜ u ⎟ → ⎜ ⎜


⎜ μb ghcosθ ⎟ 0
S = ⎜ ghsinθ − ⎟, H = ⎜


⎟ (A.2d,e)
⎝ |u| ⎠ ⎝ v ⎠
− μ ghcosθ
0 |u| b

⎛ ⎞ ⎛ ⎞
0 0
⎜ ( ) ( ( )) ⎟ ⎜ ⎟
→ ⎜ ∂ νh1/2 ∂hu 1
∂ νh1/2
∂hu ∂hv
+ ⎟ → ⎜
⎜ ( 0 ⎟

⎜ ⎟
Vx =⎜ ∂x 2 ∂y ∂x ⎟, V y = ⎜ ) ( ( )) ⎟ (A.2f,g)
⎜ + ⎟ ⎜ 1/2 ∂hv 1 1/2 ∂hu ∂hv ⎟
⎝ ∂x ∂y ⎠ ⎜ ∂ νh ∂ νh + ⎟
⎝ ∂y 2 ∂y ∂x ⎠
+
0 ∂y ∂x
(A. 1) further splits into two sets of one-dimensional equations by the operator-splitting method as
→ → →
∂Q ∂ f → → ∂ Q ∂→ g → →
+ = S + V x, + = H + V y. (A.3a,b)
∂t ∂x ∂t ∂y
Next, we apply a numerical discretization procedure on (A. 3a) within the MacCormack framework [(A. 3b) follows similarly], which includes two
successive steps during each marching time step,
( ) ( ) /
→n+1 →n →n →n →n
Qi = Qi − f i − f i− 1 ⋅Δt Δx + S i ⋅Δt, (A.4a)

[( )′ (( )′ ( )′ ) / ( )′ ]
→n+1 1 →n+1 →n →n →n →n+1 → →
Qi = Qi + Qi − f i+1 − f i ⋅Δt Δx + S i ⋅Δt + V x + T i , (A.4b)
2

where the subscript and superscript are the spatial and temporal indices, respectively. Central difference scheme is adopted to discretize the diffusive

flux V x . Note that different from the traditional MacCormack scheme, a symmetric five point TVD term Ti is inserted into the corrector step to suppress
large oscillations caused by steep gradients in the calculation [36–38]; the term is expressed as
( ) ( )
→ [ ( ) ( − )] →n →n [ ( + ) ( )] →n →n
T i = G ri+ + G ri+1 ⋅ Q i+1 − Q i − G ri+1 + G ri− ⋅ Q i − Q i− 1 , (A.5a)

ΔHi−n n
1/2 ⋅ΔHi+1/2 + Δqnxi− 1/2 ⋅Δqnxi− 1/2 + Δqnyi− 1/2 ⋅Δqnyi+1/2
ri+ = n n
, (A.5b)
ΔHi+1/2 ⋅ΔHi+1/2 + Δqnxi+1/2 ⋅Δqnxi+1/2 + Δqnyi+1/2 ⋅Δqnyi+1/2

ΔHi−n n
1/2 ⋅ΔHi+1/2 + Δqnxi− 1/2 ⋅Δqnxi+1/2 + Δqnyi− 1/2 ⋅Δqnyi+1/2
ri− = , (A.5c)
ΔHi−n 1/2 ⋅ΔHi−n 1/2 + ΔHxni− 1/2 ⋅Δqnxi− 1/2 + Δqnyi− 1/2 ⋅Δqnyi− 1/2

in which
ΔHi−n 1/2 = hni − hni− 1 , n
ΔHi+1/2 = hni+1/2 − hni , (A.6a,b)

Δqnxi− 1/2 = hn−i n ui − hni− 1 , Δqnxi+1/2 = hn−i+1n ui+1 hn−i n ui, (A.6c,d)

Δqnyi+1/2 = hn−i n vi − hn−i− 1n vi− 1 , Δqnyi+1/2 = hn−i+1n vi+1 − hn−i n vi . (A.6e,f)

The function G(x) in (A. 5a) can be expressed as


1
G(x) = C{1 − max[0, min(2x, 1)]}, (A.7)
2

11
J. Fei et al. Journal of Non-Newtonian Fluid Mechanics 287 (2021) 104411

where the expression of variable C is


{
Cr(1 − Cr), Cr ≤ 0.5
C= (A.8)
0.25, Cr > 0.5

with Cr is the Courant number, which can be expressed as


√̅̅̅̅̅
(u + gh )Δt
Cr = . (A.9)
Δx
We remark that the numerical scheme is computationally more efficient than most shock-capturing numerical methods, as the solution of the
eigenvectors and eigenvalues of (A. 5c) is not necessary. Moreover, second-order spatial and temporal accuracy can be obtained using the two-step
predictor-corrector scheme mentioned above, ensuring the scheme is advantageous over other single-step schemes, e.g., the Lax–Wendroff and
classical explicit Euler schemes.

References [20] R. Greve, Zur Ausbreitung einer Granulatlawine entlang gekrümmter Flächen -
Laborexperimente und Modellrechnungen. Diplomarbeit, Fachbereich Physik,
Technische Hochschule Darmstadt, Deutschland, 1991. (in German).
[1] S.B. Savage, Gravity flow of cohesionless granular materials in chutes and
[21] S.B. Savage, K. Hutter, The dynamics of avalanches of granular materials from
channels, Journal of Fluid Mechanics 92 (1979) 53.
initiation to run out, Acta Mechanica 86 (1991) 201–223.
[2] J.W. Vallance, Experimental and field studies related to the behavior of granular
[22] H. Nessyahu, E. Tadmor, Non-oscillatory central differencing for hyperbolic
mass flows and the characteristics of their deposits, PhD Thesis, Michigan
conservation laws, Journal of Computational Physics 87 (1990) 408–463.
Technological University, 1994.
[23] J. Balbás, E. Tadmor, C.C. Wu, Non-oscillatory central schemes for one-and two-
[3] A. Daerr, Dynamical equilibrium of avalanches on a rough plane, Physics of Fluids
dimensional MHD equations, I,, Journal of Computational Physics 201 (2004)
13 (2001) 2115–2124.
261–285.
[4] Y. Forterre, O. Pouliquen, Long-surface-wave instability in dense granular flows,
[24] A. Kurganov, E. Tadmor, New high-resolution semi-discrete central schemes for
Journal of Fluid Mechanics 486 (2003) 21–50.
Hamilton–Jacobi equations, Journal of Computational Physics 160 (2000)
[5] S.P. Pudasaini, K. Hutter, Avalanche Dynamics, Springer, Berlin, Heidelberg, 2007.
720–742.
[6] O. Pouliquen, Scaling laws in granular flows down rough inclined planes, Physics
[25] Y. Liu, Central schemes on overlapping cells, Journal of Computational Physics 209
of Fluids 11 (1999) 542.
(2005) 82–104.
[7] GDR MiDi, On dense granular flows, The European Physical Journal E 14 (2004)
[26] J.M.N.T. Gray, Y.C. Tai, S. Noelle, Shock waves, dead zones and particle-free
341–365.
regions in rapid granular free-surface flows, Journal of Fluid Mechanics 491 (2003)
[8] P. Jop, Y. Forterre, O. Pouliquen, A constitutive law for dense granular flows,
161–181.
Nature 441 (2006) 727–730.
[27] Y.C. Tai, S. Noelle, J.M.N.T. Gray, K. Hutter, Shock-capturing and front-tracking
[9] Y. Forterre, Kapitza waves as a test for three-dimensional granular flow rheology,
methods for granular avalanches, Journal of Computational Physics 175 (2002)
Journal of Fluid Mechanics 563 (2006) 123–132.
269–301.
[10] J.M.N.T Gray, A.N. Edwards, A depth-averaged μ(I)-rheology for shallow granular
[28] E.B. Pitman, C.C. Nichita, A. Patra, A. Bauer, M. Sheridan, M. Bursik, Computing
free-surface flows, Journal of Fluid Mechanics 755 (2014) 503–534.
granular avalanches and landslides, Physics of Fluids 15 (2003) 3638–3646.
[11] D. Razis, A.N. Edwards, J.M.N.T. Gray, V.D.W. Ko, Arrested coarsening of granular
[29] R.P. Denlinger, R.M. Iverson, Flow of variably fluidized granular masses across
roll waves, Physics of Fluids (2014), 26123305, https://doi.org/10.1063/
three-dimensional terrain 2. Numerical predictions and experimental tests, Journal
14904520.
of Geophysical Research: Solid Earth. 106 (2001) 553–566.
[12] A.N. Edwards, J.M.N.T. Gray, Erosion–deposition waves in shallow granular free-
[30] A.A. Medovikov, High order explicit methods for parabolic equations, BIT 38
surface flows, Journal of Fluid Mechanics 762 (2015) 33.
(1998) 372–390.
[13] S. Viroulet, J.L. Baker, F.M. Rocha, C.G. Johnson, B.P. Kokelaar, J.M.N.T. Gay, The
[31] S. Noelle, The MoT-ICE a new high-resolution wave-propagation algorithm for
kinematics of bidisperse granular roll waves, Journal of Fluid Mechanics 848
multidimensional systems of conservation laws based on Fey’s method of transport,
(2018) 836–875.
Journal of Computational Physics 164 (2000) 283–334.
[14] T. Barker, D.G. Schaeffer, P. Bohorquez, J.M.N.T Gray, Well-posed and ill-posed
[32] J.L. Baker, C.G. Johnson, J.M.N.T. Gray, Segregation-induced finger formation in
behaviour of the μ(I)-rheology for granular flow, Journal of Fluid Mechanics 779
granular free-surface flows, Journal of Fluid Mechanics 809 (2016) 168–212.
(2015) 794–818.
[33] D. Liang, R.A. Falconer, B. Lin, Comparison between TVD-MacCormack and ADI-
[15] T. Barker, J.M.N.T. Gray, Partial regularisation of the incompressible (I)-rheology
type solvers of the shallow water equations, Advances in Water Resources 29
for granular flow, Journal of Fluid Mechanics 828 (2017) 5–32, https://doi.org/
(2006) 1833–1845.
10.1017/jfm2017428.
[34] D. Liang, B. Lin, R.A. Falconer, Simulation of rapidly varying flow using an efficient
[16] O. Pouliquen, Y. Forterre, Friction law for dense granular flows: application to the
TVD–MacCormack scheme, International Journal for Numerical Methods in Fluids
motion of a mass down a rough inclined plane, Journal of Fluid Mechanics 453
53 (2007) 811–826.
(2002) 133–151.
[35] J.M.N.T. Gray, X. Cui, Weak, strong and detached oblique shocks in gravity driven
[17] R.W. Maccormack, An efficient explicit-implicit-characteristic method for solving
granular free-surface flows, Journal of Fluid Mechanics 579 (2007) 113–136.
the compressible Navier-Stokes equations, SIAM-AMS Proceedings 11 (1978)
[36] C.G. Mingham, D.M. Causon, D.M. Ingram, A TVD MacCormack scheme for
130–155.
transcritical flow, ICE Proceedings Water and Maritime Engineering 148 (2001)
[18] S.B. Savage, K. Hutter, The motion of a finite mass of granular material down a
167–175.
rough incline, Journal of Fluid Mechanics 199 (1989) 177–215.
[37] S.F. Davis, TVD finite difference schemes and artificial viscosity, ICASE Report 84
[19] D.A. Anderson, J.C. Tannehill, R.H. Pletcher, Computational Fluid Mechanics and
(20) (1984).
Heat Transfer, McGraw-Hill Book Co., New York, 1984.
[38] M. Louaked, L. Hanich, TVD scheme for the shallow water equations, Journal of
Hydraulic Research 36 (1998) 16.

12

You might also like