You are on page 1of 12

Large-eddy simulation of turbulent natural-

bed flow
Cite as: Phys. Fluids 31, 085105 (2019); https://doi.org/10.1063/1.5116522
Submitted: 25 June 2019 . Accepted: 09 July 2019 . Published Online: 05 August 2019

G. Alfonsi , D. Ferraro , A. Lauria , and R. Gaudio

COLLECTIONS

This paper was selected as an Editor’s Pick

ARTICLES YOU MAY BE INTERESTED IN

Photon-counting laser interferometer for absolute distance measurement on rough surface


Review of Scientific Instruments 90, 083101 (2019); https://doi.org/10.1063/1.5109913

Fuel-shell interface instability growth effects on the performance of room temperature


direct-drive implosions
Physics of Plasmas 26, 082701 (2019); https://doi.org/10.1063/1.5104338

PAINeT: Similarity criteria and different approaches of kinetic theory


AIP Conference Proceedings 2132, 130005 (2019); https://doi.org/10.1063/1.5119625

Phys. Fluids 31, 085105 (2019); https://doi.org/10.1063/1.5116522 31, 085105

© 2019 Author(s).
Physics of Fluids ARTICLE scitation.org/journal/phf

Large-eddy simulation of turbulent


natural-bed flow
Cite as: Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522
Submitted: 25 June 2019 • Accepted: 9 July 2019 •
Published Online: 5 August 2019

G. Alfonsi, D. Ferraro, A. Lauria, and R. Gaudioa)

AFFILIATIONS
Dipartimento di Ingegneria Civile, Università della Calabria, 87036 Rende (CS), Italy

a)
Electronic mail: gaudio@unical.it

ABSTRACT
The turbulent flow in natural rough beds is a complex subject, still poorly understood despite the longstanding effort of several researchers. In
the present work, a turbulent open-channel flow experiment, with a pebble bed at Reynolds and Froude numbers, respectively, Re = 4.65 × 104
and Fr = 0.186, has been simulated using the Large-Eddy Simulation (LES) technique, in which the wall-adapting local eddy viscosity subgrid
scale closure model is used and in the presence of an air-water interface to take into account the effects of the interface deformation in the flow
turbulence statistics under a low relative submergence condition. The simulations have been compared with a companion experiment, where
the channel bottom is constituted by four pebble layers. For the simulations, the pebble-bed surface has been captured with a high-resolution
three-dimensional laser scanner and used to morphologically characterize the numerical channel bottom. Results are presented in terms of
turbulence statistics and turbulent laws, showing a good agreement with those obtained in the experiment. Since a good convergence between
simulation and experimental results was obtained, the LES dataset was used to compute the Turbulent Kinetic Energy (TKE) dissipation rate
across the water depth. The mesh resolution allows showing a detailed TKE dissipation rate distribution across the water depth. Moreover,
the equilibrium between TKE production and dissipation was checked to verify the overlap layer existence under low relative submergence
condition. Finally, a new procedure for vortex-visualization is implemented, based on the relationship between the vorticity and the TKE
dissipation rate.
Published under license by AIP Publishing. https://doi.org/10.1063/1.5116522., s

I. INTRODUCTION the reliability of this technique in representing both the mean and
instantaneous flow fields. Nevertheless, most of the numerical works
The fluid-dynamic phenomena occurring in turbulent open- on this subject are related to regular-roughness geometries; some-
channel flows in a low relative submergence condition play a key times, they use mathematical functions to represent the bed rough-
role in natural rivers and watercourses, where the strategic element ness. Among others, Stoesser and Nikora14 performed a LES of a tur-
that generates these complex phenomena is the presence of a macro- bulent flow over square ribs mounted on a wall. Bomminayuni and
roughness bed of natural origin. In recent years, this subject has been Stoesser15 performed a LES simulation of the flow over a channel bed
faced by researchers, both experimentally and numerically. With artificially roughened with hemispheres and investigated the effect
regard to the experimental field, a noticeable effort has been made of the roughness on higher-order turbulence statistical moments.
about the turbulent characteristics in open-channel flows.1–3 In the A further LES simulation in the case of an open-channel flow over
last two decades, the spatial-averaging method (double-averaging three kinds of macroroughness beds, obtained with spheres of dif-
method) became popular, and it was used in many works to capture ferent sizes and arrangements, was proposed by Fang et al.,16 who
the heterogeneous nature of the bed in terms of stress characteris- mainly investigated the effects of permeability on the hydrauli-
tics.4–8 Most recently, the low relative submergence raised interest cally macrorough flow characteristics. Recently, Fang et al.17 per-
in researchers, and some works about the turbulent characteristics formed a LES simulation in a square duct roughened with perpen-
in this condition have been published.7–13 dicular and V-shaped ribs. Taking into account a macroroughness
In the numerical field, the Large Eddies Simulation (LES) tech- obtained through three-dimensional (3D) dunes at the laboratory
nique was successfully used, and a large number of authors showed scale, Omidyeganeh and Piomelli18,19 executed a LES simulation

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-1


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

using the Lagrangian dynamic eddy-viscosity subgrid scale (SGS) Despite these techniques exploiting different physical and math-
closure model. The bedform three-dimensionality was imposed by ematical quantities, they can be made equivalent through the
shifting a standard two-dimensional dune shape in the streamwise introduction of an equivalent threshold.29 In this work, the Q-
direction according to a sine wave; the results were validated against criterion was chosen to validate the proposed procedure for vortex-
experiments. visualization used to extract coherent structures belonging to the
As for nonregular roughness, Hardy et al.20,21 performed a LES inertial subrange.
simulation of the flow over a gravel surface, investigating the role of
the near-bed turbulence and Reynolds number on the flow-structure
development process. Moreover, Stoesser22 proposed a physically II. EXPERIMENTATION
realistic method for the LES of turbulent open-channel flow over The experimental setup of the companion experiment is con-
a granular bed, in which the description of the bed roughness was cisely summarized. The experimental test was carried out at the
accomplished by means of a roughness-geometry function together Laboratorio “Grandi Modelli Idraulici” of the University of Calabria
with a forcing term in the momentum equation. They mainly inves- (Italy) in a 16 m long, 1 m wide, 0.8 m high tilting flume with a rect-
tigated the effects of permeability on the hydraulically macrorough angular cross section. The flume inlet incorporated a stilling tank,
flow. Overall, these studies provided a relatively good insight of the an uphill spillway, a fine grid, and a honeycomb for the damping
near-wall turbulence in the cases considered, but the case of the nat- of the residual pump vibrations and the reduction of the incom-
urally rough flow still needs more investigation. Furthermore, the ing turbulence. At the flume outlet, the water depth was regulated
free surface is often assumed as a rigid frictionless boundary so that by an adjustable tailgate. In the downstream restitution channel, a
its vertical movement is neglected, also in the cases of flows under a second honeycomb was placed upstream to a Bazin weir and was
low relative submergence condition. In summary, most of the exist- used to measure the discharge with an accuracy less than 2%. Natu-
ing research related to open-channel turbulent flows over a natural ral nonuniform pebbles were spread onto the channel bed in four
highly rough bed is based on experimental activities, whereas the layers, with median size d50 = 70 mm. The 2.5 m long test sec-
majority of numerical works are related to artificially and regular- tion was located 10.0 m downstream of the inlet for an appropriate
roughness beds. In this work, the rigid lid condition was abandoned, development of the boundary layer along the flume centerline, to
and the VoF (Volume of Fluid) technique was adopted in order to avoid side-wall effects.9,10,30 The measuring grid included 25 vertical-
follow the free-surface behavior. A LES of turbulent open-channel velocity profiles, 2 cm spaced along the streamwise direction, 0.5 cm
flow with a highly rough bed was then performed, in which the Wall- and 0.3 cm spaced in the vertical direction in the proximity of the
Adapting Local Eddy viscosity (WALE) SGS closure model23,24 was free surface and of the bottom, respectively.
used. Computational solvers embedded into the OpenFOAM C++ A down-looking four-beam acoustic Doppler velocimeter
digital library were utilized then, taking into consideration the effects (ADV) (Vectrino, by Nortek) was used for the local pointwise mea-
of the moving free surface in the flow turbulence statistics under a sure of the three instantaneous components of the fluid velocity. The
low relative submergence condition. The results of the simulations 3D instantaneous velocity components (streamwise u, spanwise v,
were compared with those obtained in a companion experiment. and vertical w1 and w2 , with fluctuations u′ , v ′ , w1′ , and w2′ , along
Shifting the focus to the flow visualization, we propose a the x, y and z axes, respectively) were measured in the x–z plane
new procedure for vortex-visualization based on the relationship along the centerline. For each point of the measuring grid, the data
between the vorticity and the Turbulent Kinetic Energy (TKE) dissi- sampling frequency and the sampling time were 100 Hz and 300 s,
pation rate, ε. Many techniques in the literature exploiting the eigen- respectively, which were found to be adequate in order to achieve
values of the velocity gradient tensor or the related invariants have the statistically time-independent turbulent quantities, as obtained
been developed in order to identify coherent turbulent structures. by Dey and Das.8
Jeong and Hussain25 proposed the λ2 method for the education of The velocity measurements were taken from the lowest accessi-
the coherent structures of turbulence by considering the problem ble position into the bed grain gaps up to an elevation of about 5 cm
of the pressure minimum. Zhou et al.26 proposed a method for below the water surface, owing to the limitation of the ADV down-
the detection of vortical structures. The authors adopted the frame looking probe (the measurement volume is located at about 5 cm
independent criterion of visualizing isosurfaces of the imaginary below the emitter). The high resolution ADV measurements can
part of the complex eigenvalue pair of the velocity gradient tensor suffer from parasitical noise.31–35 Considering that the four-receiver
that represents the local swirling strength of the vortex, quantified ADV gives redundant information on the third velocity component,
by λci . A further flow visualization technique is the Q-criterion,27 which is measured simultaneously twice (obtaining w1 and w2 ), the
which exploits the velocity-gradient tensor second invariant in variance of the noise was maintained less than 0.3 (this value is
order to visualize the coherent structures into the regions where commonly accepted as no noised, as it corresponds satisfactorily to
the vorticity magnitude prevails over the strain-rate magnitude.28 signal correlations greater than 70%). Moreover, the Signal-to-Noise

TABLE I. Characteristic parameters of the experimental and simulated flow cases.

S (%) B (m) Q (l/s) hw (m) U (m/s) ν (m2 /s) ks (m) Δ (-) Re (-) Fr (-)

0.4 1.00 46.5 0.185 0.251 1.11 × 10− 6 0.0573 3.13 2.04 × 105 0.186

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-2


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

For the numerical simulations, the physical pebble-bed surface


has been captured with a 3D scanner (Minolta Vivid 300) having
resolution of 0.1 × 0.1 mm2 and used for the characterization of the
bottom of the numerical channel (Fig. 1).

III. LES FORMULATION


The filtered unsteady Navier-Stokes equations for incompress-
ible fluids in three dimensions and conservative form were consid-
ered (Einstein summation convention applies to repeated indices, i,
j = 1, 2, 3),
∂ ūi ∂ 1 ∂ p̄ ∂τij ∂ 2 ūi
+ (ūi ūj ) = − +ν , (1)
∂t ∂xj ρ ∂xi ∂xj ∂xj ∂xj
FIG. 1. Bed topography.
∂ ūi
= 0, (2)
∂xi
where ρ is the fluid density, p is the fluid pressure, ui is the fluid
Ratio (SNR) was kept equal to about 15. Furthermore, the ellipsoid velocity along the ith axis, xi is the streamwise, vertical, and spanwise
method proposed by Goring and Nikora36 was applied for despik- axis for i = 1, 2, 3, respectively (overbars denote filtered quantities),
ing. Details about the reliability of the ADV measurements are given and t is time. The effect of the small turbulent scales is mirrored by
in Ferraro et al.,9 and Coscarella et al.10 the term (τij = ūi ūj − ui uj ), representing the SGS stress, that has to be
In Table I, the characteristic parameters of the experiment con- modeled. In the framework of the LES approach, the scales smaller
sidered in the present work for comparison with the numerical than the grid size are not resolved but accounted for through the
results are summarized, S being the longitudinal flume bottom slope, subgrid scale tensor τ ij , obtained from the following relation
B the flume breadth, Q the discharge, hw the water depth measured 1
above the maximum grain-crest level, ks the statistical roughness τij − τkk δij = −2νt Sij , (3)
3
scale defined as the biggest gap into the granular bed, Δ = hw /ks
the relative submergence, Re = 4Uhw /ν the Reynolds number, U the where:
bulk flow velocity, ν the kinematic water viscosity, Fr = U/(ghw )1/2 Sij = (
1 ∂ ūi ∂ ūj
+ ) (4)
the Froude number, and g the acceleration due to gravity. 2 ∂xj ∂xi

FIG. 2. Computational domain.

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-3


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

is the deformation-rate tensor of the resolved field. In Eq. (3), νt is TABLE II. Grid-refinement parameters.
the SGS eddy viscosity that is evaluated using the WALE SGS model,
as first proposed by Nicoud and Ducros.23 This closure, based on Grid Nx Ny Nz N tot Δx+ Δy+ +
Δzmin +
Δzmax
the square of the velocity gradient tensor, accounts for the effects
of both the strain and the rotation rate of the smallest resolved tur- 1 1024 160 384 6.3 × 107 23 22 1.50 29
bulent fluctuations. The LES governing equations were discretized 2 1024 160 420 6.9 × 107 23 22 1.25 24
by means of the Finite Volume Method (FVM). As for the solution 3 1024 160 512 8.4 × 107 23 22 1.19 22
domain, a structured mesh was built, in which the dependent vari-
ables were stored at the cell center of each cell space domain in a
co-located arrangement, where the structured grid was used in order velocity, and τ w is the mean shear stress at the wall], Ntot is the total
to avoid numerical errors due to nonorthogonality. Both convec- grid node number, Δx+ , Δy+ , Δzmin + +
, and Δzmax are the grid spacings
tive and diffusive terms were approximated with second-order Cen- in terms of wall units along the three directions, respectively, where
tral Differences (CD) schemes, while the Crank-Nicolson scheme the subscripts min and max indicate the minimum and the maxi-
was used in time. The Pressure Implicit with Split Operator (PISO) mum value, respectively. The grid was uniformly spaced in both the
technique suggested by Issa37 was employed to couple the pressure streamwise and spanwise directions, and the grid spacing in terms of
and velocity in transient computations via flux conservation. The wall units was Δx+ = Δy+ = 20. Along the vertical direction, which in
PISO procedure adopts a segregated approach, and the system of the dimensionless coordinates is ẑ = (z − zc )/hw , zc being the maximum
equations is solved sequentially. The stability of the solution proce- crest level, the grid is uniformly spaced among the pebbles (−0.54 ≤ ẑ
dure has been ensured utilizing an adaptive time step with an initial ≤ 0, Δz+ = 1), whereas in the main portion of the computing domain
value of 10−6 s in conjunction with a mean Courant-Friedrichs-Lewy (Fig. 3), a grid stretching law of hyperbolic tangent type was intro-
(CFL) number limit of 0.5. In order to obtain appropriate values duced, in order to have a better spatial resolution near both the wall
of the velocity and pressure fields at the free-surface, the governing and the free-surface52
equations were solved numerically by means of the InterFoam solver
tanh[M(1 − zj )]
®
that is embedded in the OpenFoam C++ libraries. The InterFoam
code is designed for incompressible, isothermal, immiscible fluids, zjstr = Pzj + (1 − P){1 −
tanh M
}, (5)
and uses the Volume of Fluid (VoF) phase-fraction-based interface
capturing approach38 that was satisfactorily utilized in several flow
cases.39–41 The VoF method locates and tracks the free surface in
such a way that each fluid phase (air and water) occupies an indi-
vidual fraction of volume, as widely explained in Calomino et al.42
The computational domain spanned in the range −2.7 ≤ x/hw ≥ 2.7
along the streamwise direction, in the range −0.4 ≤ y/hw ≥ 0.4
along the spanwise direction (so representing the pebble-bed sur-
face captured with the high-resolution 3D laser scanner), and in
the range −0.54 ≤ z/hw ≥ 1.54 in the vertical direction (Fig. 2).
The size of the computing domain had the same dimension as the
commonly accepted domain size for smooth-bed flows.15,16 More-
over, it was large enough to contain the large-scale turbulence struc-
tures observed in the literature.43–51 The setup and boundary con-
ditions were selected similar to those of the aforementioned flume
experiment. Boundary conditions were imposed as no-slip and zero
wall-normal velocity at the pebble-bed surface. Periodic boundary
conditions were used in the streamwise direction, according to the
uniform flow condition. Periodic boundary conditions were also
applied along the spanwise direction, while, in order to represent the
effects of the water surface variations, the VoF method was enforced.
The flow was driven by a pressure gradient that kept the flowrate
constant over time.

IV. COMPUTATIONAL SCHEME


As concerns the grid resolution, various numerical tests were
performed before reaching the final configuration of the grid (grid
3 in Table II), employing an increasing number of points mainly
along the vertical direction. The results of the numerical tests are
outlined in Table II, in which N x , N y , and N z are the grid node num-
ber in the streamwise, spanwise, and vertical directions, respectively
[recall that xi+ = xi u∗ /ν, u+i = ui /u∗ , u∗ = (τw /ρ)1/2 being the shear
FIG. 3. Computational grid in the generic x–z plane.

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-4


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

where zjstr are the stretched point elevations, zj indicates the uni- TABLE IV. Comparison of grid-point requirements between the present LES and a
form distribution, and P = 1.7 and M = 1.9 are two parameters hypothetical DNS (with uniform grid resolving η along the three coordinate directions).
characterizing the distribution. The partial derivatives along z were
calculated according to the distribution in Eq. (5). After the insertion u∗ (m/s) 0.0239 Experimental value
of appropriate initial conditions (an initial velocity profile evolving εbulk (m2 /s3 ) 7.75 × 10− 4 Experimental value
with time), the initial transient of the flow was first simulated by ν = Δx = Δy = Δz (m) 1.89 × 10− 4 Hypothetical DNS
means of the Reynolds Averaged Navier-Stokes (RANS) equations N tot (DNS) 8.5 × 109 Hypothetical DNS
coupled with the k − ω Shear Stress Transport (SST) closure model.42 N tot (LES) 8.4 × 107 Present LES
Then, the RANS steady state was mapped onto the LES domain and N tot (DNS)/N tot (LES) 101.2 ...
run for 15t (t = 5.4hw /U is the flow-through time) for an appropri-
ate turbulence development. Finally, the LES turbulent steady state
was run for 30t to build up the turbulent-flow database.
A CPU-based computational system was used for the simula- respectively. The results are reported in Table IV. It appears that
tions. The system included 1 worker node equipped with 2 eight- the present simulation executed via a hypothetical DNS on a uni-
core Intel E5-2640 CPUs (for a total of 16 cores/16 threads at form grid (able to resolve the Kolmogorov spatial microscale η
2.0 GHz), 128 GB RAM at 1899 MHz, and 4 TB of disk space. along all the three coordinate directions) would require a number
The code was parallelized through the public domain OpenMPI of grid points about 102 times greater than that of the present-work
implementation of the standard Message Passing Interface (MPI). LES.
The computational domain was decomposed into 16 subdomains (4
along the spanwise direction and 4 along the vertical direction) with V. RESULTS AND DISCUSSION
the Simple Geometric Domain Decomposition (SGDD) technique,
for an appropriate balancing of the computational weight among the A. Second-order statistics
different processors. Concerning the code performance, the simula- For a steady uniform flow over a rough bed, the spatially aver-
tion was preliminarily run on different machine configurations, and aged (SA) total shear stress ⟨τ⟩ resulting from the double-averaged
the results in terms of runtime in executing one full time step (Δt) (DA) Navier-Stokes (DANS) equations (see Refs. 2 and 4) is
of the calculation procedure on grid 3 were reported in Table III
(I/O operations were not included). It can be noted that, for the d⟨u⟩
⟨τ̄⟩ = −ρ⟨ũw̃⟩ − ρ⟨u′ w′ ⟩ + ρν , (8)
total number of grid points used, the execution time decreases, as dz
expected, with the number of processors involved in the calculations.
As mentioned before, the numerical simulation of the case at hand where −ρ⟨ũw̃⟩ is the form-induced shear stress, where the tilde
was run on the 16-core machine configuration on grid 3 (N tot = 8.4 accent represents the difference between the time averaged veloc-
× 107 , Re = 4.65 × 104 ). The elapsed computational time resulted ity and the corresponding DA value (e.g., ũ = u − ⟨u⟩), −ρ⟨u′ w′ ⟩
in a total of about 1500 h, which is a rather reasonable value for a the SA Reynolds shear stress and ρν d⟨u⟩/dẑ the SA viscous shear
relatively complex LES simulation using the worker node depicted stress. In Fig. 4, shear stresses from LES and experiment are com-
before. In this context, the grid requirements of the present LES pared. SA Reynolds stress, form-induced, and viscous shear stress
are evaluated in a comparison with those of a hypothetical Direct were made dimensionless dividing by ρu2∗ and expressed as a func-
Numerical Simulation (DNS). The Kolmogorov spatial microscale η tion of the nondimensional vertical axis. Both the experimental and
for open-channel flow numerical tests show a good agreement with the gravity line, so that
1/4 it is guaranteed that the experimental as well as the LES results mir-
ν3
η=( ) (6) ror a uniform-flow condition. Figure 4(a) does not show data in
ε the upper 5 cm of the water depth because the ADV down-looking
was first estimated using the method of the average dissipation rate configuration does not allow capturing velocity signals in the 5 cm
εbulk per unit mass53 below the emitter. In Fig. 4(b), the turbulent shear stress as well as
the form-induced stress are smooth if compared to the experimental
τw Lx Ly U u2∗ U
εbulk ≅ ≅ , (7) counterpart, suggesting that the fluctuations exhibited in panel (a)
ρhw Lx Ly hw are an ADV artefact. The total shear stress in panel (b) is in a good
where Lx and Ly are the actual (dimensional) lengths of the numer- agreement with the gravity line, so guaranteeing the uniform flow
ical channel along the streamwise and the spanwise directions, condition also in the LES test.
The form-induced stress can be seen as an indicator to define
TABLE III. Full-Δt runtime with different machine configurations (seconds per Δt).
the roughness layer, in which the turbulent stress should dominate
over −ρ⟨ũw̃⟩. As shown in Fig. 4, the roughness-layer upper edge is
ẑ < 0.1 in panel (a), while in panel (b), it can be placed at ẑ < 0.2,
CPU cores Simulation (N tot = 8.4 × 107 , Re = 4.65 × 104 )
where the form-induced stress is not negligible. Hence, above the
2 262.50 pebble crests, the main contribution to the SA total shear stress is
4 86.42 due to the SA turbulent shear stress −ρ⟨u′ w′ ⟩, as suggested by other
8 65.38 authors.54–56 The turbulent stress attains a peak close to the crest
16 55.00 level and shows a damping within the interfacial sublayer (ẑ < 0.1)
in both the experimental and LES tests. In the interfacial sublayer,

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-5


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 4. Reynolds shear stress distribu-


tion: (a) experimental data and (b) LES
data. Red squares represent the form-
induced shear stress, green triangles the
Reynolds shear stress, open bullets the
total shear stress, and blue dashed lines
the gravity line.

−ρ⟨u′ w′ ⟩ is compensated by −ρ⟨ũw̃⟩ that approaches its maximum 4/5-law in a fully turbulent flow, based on the statistical theory of
value among the roughness elements, contributing up to 20% to the turbulence.57 Finally, since the simulation showed a good agree-
total shear stress and upholding the works of Giménez-Curto and ment with measurements in all the proposed analyses of first and
Lera,2 Manes, Pokrajac, and McEwan.6 second order moments as well as the turbulent 4/5 law, we can
assert that the LES proposed here can be used to explore the iner-
B. Third-order laws of turbulence tial subrange and smaller scales, according to the grid resolution.
Figure 6 shows ⟨ε⟩ at several vertical elevations (from ẑ = 0 to
In real space, a stringent way to check the existence of the
ẑ = 0.9). This figure provides a more complete view, including posi-
Kolmogorov scaling is to consider the third-order structure func-
tions below which there were no available ADV data, owing to the
tions.57 One of the few exact theorems of turbulence is the Kol-
Taylor hypothesis violation (ẑ < 0.1),30 and positions in which it was
mogorov 4/5-law,58 given by the following equation in compensated
not possible to take measurements (ẑ > 0.68). The ⟨ε⟩ distribution
form:9,10,30
(see Fig. 7) decreases moving away from the source of turbulence,
5
⟨ε⟩ = − ⟨Δu3 ⟩, (9)
4r
where Δu is the double averaged space increment of the stream-
wise velocity and r is the increment vector of magnitude r. As
before, the overbar and the angle brackets identify time and spatial
average, respectively. Equation (9) gives a way to compute directly
the TKE dissipation rate in a turbulent flow. In Fig. 5, a compari-
son between the TKE dissipation rate from LES and experiment is
shown at several vertical dimensionless elevations ẑ. Once again,
the match between simulation and measurements is satisfactory.
Namely, all the compensated laws in Fig. 5 show a plateau in the
inertial subrange, which can be fitted to obtain the TKE dissipation
rate and a sharp decrease going toward the largest scales.9 More-
over, the resolved scales in the LES are smaller than those captured FIG. 6. TKE dissipation rate vs space increment. Five-pointed stars repre-
with the ADV using the LES data, showing the correct simulation sent ⟨ε⟩(ẑ = 0.9), downward-pointing triangles ⟨ε⟩(ẑ = 0.8), upward-pointing
of the inertial subrange that is the only need for the Kolmogorov triangles ⟨ε⟩(ẑ = 0.7), diamonds ⟨ε⟩(ẑ = 0.6), squares ⟨ε⟩(ẑ = 0.5), crosses
⟨ε⟩(ẑ = 0.4), points ⟨ε⟩(ẑ = 0.3), asterisks ⟨ε⟩(ẑ = 0.2), circles ⟨ε⟩(ẑ = 0.1),
and plus sign ⟨ε⟩(ẑ = 0).

FIG. 5. TKE dissipation rate vs space increment. Open bullets represent the exper-
imental data, whereas filled bullets are related to LES. Black markers represent
⟨ε⟩(ẑ = 0.2), blue markers ⟨ε⟩(ẑ = 0.3), green markers ⟨ε⟩(ẑ = 0.4), and red FIG. 7. TKE dissipation rate vs dimensionless elevation. Open bullets represent
markers ⟨ε⟩(ẑ = 0.5). the experimental data, whereas filled bullets are related to the LES data.

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-6


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

reflected from a bed of complex geometry, as in the case of pebbles


or cobbles.36 Namely, the cause of this kind of noise in a rough-bed
is the overlap of reflected echoes because of the nonregular bed sur-
face. Hence, the missing points observed in Fig. 9(a) are due to the
authors choice of discarding the measurements in which a high level
of noise was registered.9 The highest resolution of the numerical
velocity profiles (Fig. 9) shows a clear recirculation in the bed gaps
(1.4 < x/hw < 1.6) that was also captured by the ADV data. Uphold-
ing the work of Dey and Das,8 in the roughness gaps, a velocity pro-
files inflection is observed, whereas accelerated velocity profiles are
located on the roughness crests. Furthermore, Dey et al.60 imputed
that the velocity profile retard is due to the mixing process in the
wake region of roughness elements. According to literature observa-
tions,8,61,62 the S-shaped velocity profiles appear into the roughness
wake region (see Fig. 9, 1.2 < x/hw < 1.6). Moreover, an inversion of
the velocity direction is visible below the crest levels for ẑ < −0.1.
FIG. 8. Dimensionless mean velocity profiles at different locations in the com- This behavior is not observed in all the gaps, but it appears only
puting domain: open bullets represent the experimental results and solid line the for gap dimensions comparable to ks . Namely, the roughness wake
numerical ones. vortex, which is the cause of the velocity inversion, needs enough
space to develop. An extensively debated theme in the literature is
which in this case is the gravel bed. Simultaneously, ⟨ε⟩ shows an the absence of the universal log-law velocity distribution in low rel-
inertial subrange band shrinking close to the water surface. These ative submergence flows. According to the Townsend63 hypothesis,
facts suggest a damping operated by the water-air interfacial sur- the log-law arises if an overlap layer appears. This layer is character-
face, as well visible in Fig. 7. Once again, the comparison between ized by the wall distance that should be the only relevant length scale.
the experiment and the LES shows a good agreement. Both the The Townsend63 hypothesis is verified if the equilibrium between
tests present a quite linear trend in logarithmic scale from ẑ = 0 to the SA TKE production (⟨TKEp ⟩) and dissipation occurs (11),
ẑ = 0.68, that is, the maximum elevation measured in the flume test. 3/2
In this respect, it is important to recall the free surface treatment that ⟨(u′2 + v ′2 + w′2 ) ⟩
was handled with the VoF technique (instead of a rigid lid). This = ⟨ε⟩. (10)
z − zc
resulted to be extremely important in problems of fully turbulent
flow with small relative submergence. Here, an analysis based on the equilibrium between ⟨TKEp ⟩ and
⟨ε⟩ is proposed in order to validate the absence of the log-law.
C. Mean velocity profiles Figure 10 does not show an equilibrium layer in which the differ-
ence ⟨TKEp ⟩ − ⟨ε⟩ should be equal to 0. The last analysis upholds
Computed and measured velocity profiles are compared in
the idea that the low relative submergence does not allow the
Fig. 8, showing a good accordance. Some disagreements are visible
defect-law to hold, and, therefore, the overlap layer is also inex-
only in the bed proximity. These mismatches are related to the ADV
istent.64,65 Moreover, the influence of the relative submergence on
data because of the undoubted difficulties in measuring inside the
the occurrence of the log-law distribution was also observed by
layers influenced by the roughness. A further comparison is given
Nikora et al.4
in Fig. 9, where the streamwise time-averaged velocity-vector distri-
butions along the channel centerline are shown. This view presents
the resolution of the ADV data vs the LES data. The high resolu- D. Flow visualization
tion ADV measurements are known to suffer from parasitical noise An investigation on vortices in the turbulent open channel is
contributions.31–35,59 This condition made the despiking method conducted using the database generated by the large-eddy simula-
inapplicable. Such a situation may occur when the ADV pulses are tion and adopting the following relation:

FIG. 9. Snapshots of vertical mean-


velocity-vector distributions along the
channel centerline: (left panel) experi-
mental; (right) numerical.

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-7


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

This fact can be exploited in order to identify the smallest eddies


from the TKE dissipation rate ⟨ε⟩. As a consequence of Eq. (11), a
new vortex visualization method is proposed. The ω field is plot-
ted as a color map and a vorticity value ω = (⟨ε⟩/ν)1/2 is used to
compute a contour map. Figure 11 shows the contours of the vortic-
ity structure associated with the TKE dissipation rate resulting from
the Kolmogorov 4/5-law. The structures highlighted by the method
must have a dimension contained in the inertial subrange shown in
Fig. 6. As expected, a high population of vortices belonging to the
inertial subrange is visible close to the source of turbulence (pebble
bed). Gradually moving away from the bed, these vortices are less in
number. This is consistent with Figs. 6 and 7 that show a lower value
of ⟨ε⟩ far from the bed.
For further investigation, the simulated flow field was used to
extract the turbulent coherent structures with the Q-criterion.27 This
method is based on the velocity-gradient tensor second invariant
FIG. 10. LES results of ⟨TKEp ⟩ − ⟨ε⟩ distribution along the water depth. in order to visualize the coherent structures. The structures were
identified through a nonzero threshold value. A criticism of the Q-
criterion is the arbitrariness of the threshold value, which is a free
⟨ε⟩ = νω2 . (11) user choice.68 Namely, the user is not able to set an a priori thresh-
old value of Q to visualize the coherent structures populating the
Equation (11) shows that the energy dissipation is also associated inertial subrange. Contrariwise, in the proposed method the vor-
with the vorticity field ω,66 suggesting that the vorticity is connected ticity contours are directly connected to the energy dissipation rate
to the Richardson energy cascade process. According to Davidson,67 value. Namely, once obtained ⟨ε⟩ from the Kolmogorov 4/5-law, it is
in a turbulent flow, the smallest scale owns a high value of vortic- possible to compute a unique threshold value capable of visualizing
ity, which can be seen as a measure of the dissipation energy rate. the vortex contours contained into the inertial subrange. Figure 12 is

FIG. 11. Inertial subrange vortex visual-


ization for different water elevations at
a generic time instant. Colormaps rep-
resent the vorticity field magnitude and
black lines are the contours of the vor-
tices associated with ⟨ε⟩.

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-8


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 12. Flow visualization comparison at the ẑ = 0.7 plane, where the abscissa is about 0 < x < 0.2. Contour lines represent the vortices intersecting a plane by using
the proposed method. In subfigure (a), the coherent structures are highlighted through the contour lines rendering, whereas in subfigure (b) the coherent structures were
extracted using the Q-criterion.

a 3D representation of vortices and vortical structures that clarifies which is consistent with Fig. 11. Moreover, in Fig. 13, a reduc-
the way in which the latter structures cross the plane, represented tion in the number of the visualized structures is observed. This
through the contour lines defining the vortices. In Fig. 12(a), the fact is due to the limit dimension given by Eq. (9), which connects
contour lines are rendered for several contiguous planes in order to the TKE energy dissipation rate value to a characteristic length.
extract the coherent structures. Moreover, in Fig. 12(b), the contour Namely, the limit dimension of a vortex is the maximum length
lines representing the vortices belonging to the inertial subrange which gives a constant value of ε, and once a vortex overcomes the
are compared to the three-dimensional structures extracted with limit dimension it is no more belonging to the inertial subrange and,
the Q-criterion within a volume that encompasses several planes consequently, it is not visualized. This fact suggests that the vortex
slightly above the reference plane. Figure 12(a) shows worm-like reduction observed above ẑ > 0.4 is the result of a vortex stretching
shape coherent structures, which have an intense vortical motion process up to the overcoming of the limit dimension.
that is a characteristic of isotropic turbulence.68 This fact confirms
that the visualized structures are compatible with the inertial sub-
range. Figure 12(b) shows a good match between the coherent struc- VI. CONCLUSIONS
tures extracted through the proposed method and the Q-criterion. Overall, the LES code used for the simulations in conjunction
All the contours contain a coherent structure passing through the with the WALE SGS closure model has shown a rather good abil-
reference plane. In Fig. 13, the coherent structures (belonging to ity in correctly reproducing the turbulent phenomena at hand. It
the inertial subrange) are visualized across the water depth, reveal- appears that the present simulation executed via a hypothetical DNS
ing a great population of smaller structures near the bed, which is on a uniform grid (able to resolve the Kolmogorov spatial microscale
the source of the turbulence in the present work. The visualized η along all the three coordinate directions) would require a number
structure dimensions grow according to the distance from the bed, of grid points about 102 times greater than that of the present-work
LES. The adopted numerical approach guarantees a resolved-scale
resolution that is enough to glimpse into the inertial subrange and
perform the analysis of turbulence proposed here. A first portion
of the present work deals with the testing of the reliability of the
LES calculations with respect to the measured dataset. Once confi-
dent about the convergence of the second-order moment analysis,
as well as the more sensible third-order structures, we used the LES
dataset to compute the TKE dissipation rate (including the small-
est scales of the inertial subrange) that was not captured with the
ADV. Third-order structures clearly show the presence of an inertial
subrange as related to both experimental and numerical results, giv-
ing a way to precisely compute the value of ⟨ε⟩. Moreover, the ADV
down-looking configuration does not allow the measurements in the
upper 5 cm of the water depth. A deviation from the decreasing
trend was observed in the proximity of the water surface. Simulta-
neously, ⟨ε⟩ shows an inertial subrange band shrinking close to the
water surface. These facts suggest the existence of a damping caused
by the water-air interfacial surface. In this respect, it is important to
recall that the free surface treatment has been handled with the VoF
technique. The Townsend hypothesis was used to test the univer-
sal log-law existence. The TKE production and dissipation suggest
that an equilibrium layer does not exist, and, then, the absence of
FIG. 13. Flow visualization box in which the coherent structures, extracted through the universal log-law in flows under low relative submergence con-
the proposed method, are shown across the water depth (0 < ẑ < 1), whereas the dition. Here, a novel flow visualization method based on the TKE
streamwise abscissa is 0 < x < 0.2. dissipation rate definition is presented. This method is devoted to

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-9


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

21
visualizing coherent structures belonging to the inertial subrange. R. J. Hardy, J. L. Best, S. N. Lane, and P. E. Carbonneau, “Coherent flow struc-
Namely, once computed, the TKE dissipation rate from the Kol- tures in a depth-limited flow over a gravel surface: The role of near-bed turbu-
mogorov 4/5 law, the coherent structures having a vorticity value lence and influence of Reynolds number,” J. Geophys. Res.: Earth Surf. 114, 1,
linked to ⟨ε⟩ were visualized. As opposed to other criteria, the pro- https://doi.org/10.1029/2007jf000970 (2009).
22
T. Stoesser, “Physically realistic roughness closure scheme to simulate turbulent
posed method gives a precise way to extract the coherent structure channel flow over rough beds within the framework of LES,” J. Hydraul. Eng. 136,
belonging to the inertial subrange. Finally, the coherent structures 812–819 (2010).
extracted through the proposed method were compared with those 23
F. Nicoud and F. Ducros, “Subgrid-scale stress modelling based on the
coming from the Q-criterion, showing a good match between the square of the velocity gradient tensor,” Flow, Turbul. Combust. 62, 183–200
two methods. (1999).
24
S. Rezaeiravesh and M. Liefvendahl, “Effect of grid resolution on large eddy
simulation of wall-bounded turbulence,” Phys. Fluids 30, 055106 (2018).
25
REFERENCES J. Jeong and F. Hussain, “On the identification of a vortex,” J. Fluid Mech. 285,
69–94 (1995).
1
B. Kironoto, W. H. Graf et al., “Turbulence characteristics in rough uniform 26
J. Zhou, R. J. Adrian, S. Balachandar, and T. Kendall, “Mechanisms for gener-
open-channel flow,” Proc. Inst. Civ. Eng.: Water Marit. Energy 106, 333–344 ating coherent packets of hairpin vortices in channel flow,” J. Fluid Mech. 387,
(1994). 353–396 (1999).
2
L. A. Giménez-Curto and M. A. C. Lera, “Oscillating turbulent flow 27
J. Zhong, T. S. Huang, and R. J. Adrian, “Extracting 3D vortices in turbulent
over very rough surfaces,” J. Geophys. Res.: Oceans 101, 20745–20758, fluid flow,” IEEE Trans. Pattern Anal. Mach. Intell. 20, 193–199 (1998).
https://doi.org/10.1029/96jc01824 (1996). 28
V. Kolář, “Vortex identification: New requirements and limitations,” Int. J. Heat
3
A. Dittrich and K. Koll, “Velocity field and resistance of flow over rough sur- Fluid Flow 28, 638–652 (2007).
face with large and small relative submergence,” Int. J. Sediment Res. 12(3), 21–33 29
Q. Chen, Q. Zhong, M. Qi, and X. Wang, “Comparison of vortex identifica-
(1997 ), see https://publikationen.bibliothek.kit.edu/233297. tion criteria for planar velocity fields in wall turbulence,” Phys. Fluids 27, 085101
4
V. Nikora, D. Goring, I. McEwan, and G. Griffiths, “Spatially averaged open- (2015).
channel flow over rough bed,” J. Hydraul. Eng. 127, 123–133 (2001). 30
D. Ferraro, S. Servidio, and R. Gaudio, “Velocity scales in steady-nonuniform
5
V. Nikora, I. McEwan, S. McLean, S. Coleman, D. Pokrajac, and R. Wal- turbulent lows with low relative submergence,” Environ. Fluid Mech. 1–17 (2019).
ters, “Double-averaging concept for rough-bed open-channel and overland flows: 31
J. L. Garbini, F. K. Forster, and J. E. Jorgensen, “Measurement of fluid turbu-
Theoretical background,” J. Hydraul. Eng. 133(8), 873–883 (2007). lence based on pulsed ultrasound techniques. Part 2. Experimental investigation,”
6
C. Manes, D. Pokrajac, and I. McEwan, “Double-averaged open-channel flows J. Fluid Mech. 118, 471–505 (1982).
with small relative submergence,” J. Hydraul. Eng. 133(8), 896–904 (2007). 32
R. Lhermitte and U. Lemmin, “Open-channel flow and turbulence measure-
7
J. Aberle, K. Koll, and A. Dittrich, “Form induced stresses over rough gravel- ment by high-resolution Doppler sonar,” J. Atmos. Oceanic Technol. 11, 1295–
beds,” Acta Geophys. 56(3), 584–600 (2008). 1308 (1994).
8 33
S. Dey and R. Das, “Gravel-bed hydrodynamics: Double-averaging approach,” J. G. Voulgaris and J. H. Trowbridge, “Evaluation of the acoustic Doppler
Hydraul. Eng. 138, 707–725 (2012). velocimeter (ADV) for turbulence measurements,” J. Atmos. Oceanic Technol.
9
D. Ferraro, S. Servidio, V. Carbone, S. Dey, and R. Gaudio, “Turbulence laws in 15, 272–289 (1998).
natural bed flows,” J. Fluid Mech. 798, 540–571 (2016). 34
S. J. McLelland and A. P. Nicholas, “A new method for evaluating errors in high-
10
F. Coscarella, S. Servidio, D. Ferraro, V. Carbone, and R. Gaudio, “Turbulent frequency ADV measurements,” Hydrol. Processes 14, 351–366 (2000).
energy dissipation rate in a tilting flume with a highly rough bed,” Phys. Fluids 29, 35
D. Hurther and U. Lemmin, “A correction method for turbulence measure-
085101 (2017). ments with a 3D acoustic Doppler velocity profiler,” J. Atmos. Oceanic Technol.
11
S. Cameron, V. Nikora, and M. Stewart, “Very-large-scale motions in rough-bed 18, 446–458 (2001).
open-channel flow,” J. Fluid Mech. 814, 416–429 (2017). 36
D. G. Goring and V. I. Nikora, “Despiking acoustic Doppler velocimeter data,”
12
E. Padhi, N. Penna, S. Dey, and R. Gaudio, “Hydrodynamics of water-worked J. Hydraul. Eng. 128, 117–126 (2002).
37
and screeded gravel beds: A comparative study,” Phys. Fluids 30, 085105 (2018). R. I. Issa, “Solution of the implicitly discretised fluid flow equations by operator-
13
E. Padhi, N. Penna, S. Dey, and R. Gaudio, “Spatially averaged dissipation rate splitting,” J. Comput. Phys. 62, 40–65 (1986).
38
in flows over water-worked and screeded gravel beds,” Phys. Fluids 30, 125106 C. W. Hirt and B. D. Nichols, “Volume of fluid (VOF) method for the dynamics
(2018). of free boundaries,” J. Comput. Phys. 39, 201–225 (1981).
14 39
T. Stoesser and V. I. Nikora, “Flow structure over square bars at intermediate G. Alfonsi, A. Lauria, and L. Primavera, “Proper orthogonal flow modes in the
submergence: Large eddy simulation study of bar spacing effect,” Acta Geophys. viscous-fluid wave-diffraction case,” J. Flow Visualization Image Process. 20, 227
56, 876 (2008). (2013).
15 40
S. Bomminayuni and T. Stoesser, “Turbulence statistics in an open-channel flow G. Alfonsi, A. Lauria, and L. Primavera, “The field of flow structures gener-
over a rough bed,” J. Hydraul. Eng. 137, 1347–1358 (2011). ated by a wave of viscous fluid around vertical circular cylinder piercing the free
16
H. Fang, X. Han, G. He, and S. Dey, “Influence of permeable beds on hydrauli- surface,” Procedia Eng. 116, 103–110 (2015).
41
cally macro-rough flow,” J. Fluid Mech. 847, 552–590 (2018). G. Alfonsi, A. Lauria, and L. Primavera, “Recent results from analysis of flow
17
X. Fang, Z. Yang, B.-C. Wang, M. F. Tachie, and D. J. Bergstrom, “Large- structures and energy modes induced by viscous wave around a surface-piercing
eddy simulation of turbulent flow and structures in a square duct roughened with cylinder,” Math. Probl. Eng. 2017, 1.
42
perpendicular and V-shaped ribs,” Phys. Fluids 29, 065110 (2017). F. Calomino, G. Alfonsi, R. Gaudio, A. D’Ippolito, A. Lauria, A. Tafarojnoruz,
18
M. Omidyeganeh and U. Piomelli, “Large-eddy simulation of three-dimensional and S. Artese, “Experimental and numerical study of free-surface flows in a
dunes in a steady, unidirectional flow. Part 1. Turbulence statistics,” J. Fluid Mech. corrugated pipe,” Water 10, 638 (2018).
43
721, 454–483 (2013). K. Kim and R. Adrian, “Very large-scale motion in the outer layer,” Phys. Fluids
19
M. Omidyeganeh and U. Piomelli, “Large-eddy simulation of three-dimensional 11, 417–422 (1999).
44
dunes in a steady, unidirectional flow. Part 2. Flow structures,” J. Fluid Mech. 734, R. Adrian, C. Meinhart, and C. Tomkins, “Vortex organization in the outer
509–534 (2013). region of the turbulent boundary layer,” J. Fluid Mech. 422, 1–54 (2000).
20 45
R. J. Hardy, S. N. Lane, R. I. Ferguson, and D. R. Parsons, “Emergence of coher- A. D. Kirkbride and R. Ferguson, “Turbulent flow structure in a gravel-bed river:
ent flow structures over a gravel surface: A numerical experiment,” Water Resour. Markov chain analysis of the fluctuating velocity profile,” Earth Surf. Processes
Res. 43, W03422, https://doi.org/10.1029/2006wr004936 (2007). Landforms 20, 721–733 (1995).

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-10


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

46 57
T. Buffin-Bélanger, A. G. Roy, and A. D. Kirkbride, “On large-scale flow U. Frisch, Turbulence: The Legacy of an Kolmogorov (Cambridge University
structures in a gravel-bed river,” Geomorphology 32, 417–435 (2000). Press, 1995).
47 58
A. B. Shvidchenko and G. Pender, “Macroturbulent structure of open-channel A. N. Kolmogorov, “Dissipation of energy in locally isotropic turbulence,” Dokl.
flow over gravel beds,” Water Resour. Res. 37, 709–719, https://doi.org/10.1029/ Akad. Nauk SSSR 32, 16 (1941).
2000wr900280 (2001). 59
M. Franca and U. Lemmin, “Eliminating velocity aliasing in acoustic Doppler
48
B. Hofland and R. Booij, “Measuring the flow structures that initiate stone velocity profiler data,” Meas. Sci. Technol. 17, 313 (2006).
movement,” in River Flow (AA Balkema Rotterdam, The Netherlands, 2004), 60
S. Dey, S. Sarkar, S. K. Bose, S. Tait, and O. Castro-Orgaz, “Wall-wake flows
pp. 821–830. downstream of a sphere placed on a plane rough wall,” J. Hydraul. Eng. 137, 1173–
49
A. G. Roy, T. Buffin-Bélanger, H. Lamarre, and A. D. Kirkbride, “Size, shape and 1189 (2011).
dynamics of large-scale turbulent flow structures in a gravel-bed river,” J. Fluid 61
M. Franca and U. Lemmin, “The simultaneous occurrence of logarithmic and
Mech. 500, 1–27 (2004). S-shaped velocity profiles in gravel-bed river flows,” Arch. Hydro-Eng. Environ.
50
G. A. Marquis and A. G. Roy, “Effect of flow depth and velocity on the scales of Mech. 56(1-2), 29–41 (2009), ISSN 1231–3726.
62
macroturbulent structures in gravel-bed rivers,” Geophys. Res. Lett. 33, L24406, E. Mignot, E. Barthelemy, and D. Hurther, “Double-averaging analysis and local
https://doi.org/10.1029/2006gl028420 (2006). flow characterization of near-bed turbulence in gravel-bed channel flows,” J. Fluid
51
M. J. Franca and U. Lemmin, “Detection and reconstruction of large-scale Mech. 618, 279–303 (2009).
63
coherent flow structures in gravel-bed rivers,” Earth Surf. Processes Landforms A. Townsend, “Equilibrium layers and wall turbulence,” J. Fluid Mech. 11, 97–
40, 93–104 (2015). 120 (1961).
52 64
G. Alfonsi, A. Lauria, and L. Primavera, “On evaluation of wave forces and J. Jiménez and J. del Alamo, “Computing turbulent channels at experimen-
runups on cylindrical obstacles,” J. Flow Visualization Image Process. 20, 269 tal Reynolds numbers,” in 15th Australasian Fluid Mechanics Conference, 13–17
(2013). December 2004.
53 65
H. P. Bakewell, Jr. and J. L. Lumley, “Viscous sublayer and adjacent wall region W. K. George, “Is there a universal log law for turbulent wall-bounded flows?,”
in turbulent pipe flow,” Phys. Fluids 10, 1880–1889 (1967). Philos. Trans. R. Soc., A 365, 789–806 (2007).
54 66
C. Manes, D. Poggi, and L. Ridolfi, “Turbulent boundary layers over permeable H. Tennekes and J. L. Lumley, A First Course in Turbulence (MIT Press,
walls: Scaling and near-wall structure,” J. Fluid Mech. 687, 141–170 (2011). 1972).
55 67
S.-Q. Yang and A. T. Chow, “Turbulence structures in non-uniform flows,” Adv. P. A. Davidson, Turbulence: An Introduction for Scientists and Engineers
Water Resour. 31, 1344–1351 (2008). (Oxford University Press, 2015).
56 68
S.-Q. Yang and S.-Y. Lim, “Discussion of ‘Shear stress in smooth rectangular P. Chakraborty, S. Balachandar, and R. J. Adrian, “On the relationships
open-channel flows’ by Junke Guo and Pierre Y. Julien,” J. Hydraul. Eng. 132, between local vortex identification schemes,” J. Fluid Mech. 535, 189–214
629 (2006). (2005).

Phys. Fluids 31, 085105 (2019); doi: 10.1063/1.5116522 31, 085105-11


Published under license by AIP Publishing

You might also like