International Journal of Hydraulic Engineering
p-ISSN: 2169-9771 e-ISSN: 2169-9801
2018; 7(2): 33-42
doi:10.5923/j.ijhe.20180702.03

Wisam Alawdi1, 2, T. D. Prasad1
1School of Computing, Science & Engineering; University of Salford, Salford, M5 4WT, UK
2Department of Civil Engineering, University of Basra, Basra, Iraq
Correspondence to: Wisam Alawdi, School of Computing, Science & Engineering; University of Salford, Salford, M5 4WT, UK.
| Email: | ![]() |
Copyright © 2018 The Author(s). Published by Scientific & Academic Publishing.
This work is licensed under the Creative Commons Attribution International License (CC BY).
http://creativecommons.org/licenses/by/4.0/

Several analytical models of velocity distribution for turbulent uniform open channel flows were lately developed by analysis and simplification of the Reynolds-averaged Navier-Stokes equations (RANS). These simplified RANS-based models, which are called dip-modified laws, are frequently employed to predict the velocity profile in flow cases where the maximum velocity may occur below the water surface. In this paper, the performance of two simplified RANS models, namely the dip-modified log law (DML-law) and the dip-modified log wake law (DMLW-law) are compared against the full 3D RANS model used in the computational fluid dynamics (CFD) modelling. The results show that although the simplified RANS models can predict the velocity dip phenomenon, the accuracy of such models is less than the full RANS (CFD) model. This is likely to be due to the assumption imposed for approximating the secondary current term in the governing equations. It is also found that the DMLW-law can give results closer to that obtained by the full RANS model. This may because of including the wake effect in eddy viscosity calculation.
Keywords: Open channel flow, Turbulence, Velocity dip phenomenon, Velocity distribution, RANS
Cite this paper: Wisam Alawdi, T. D. Prasad, Comparison of Performance of Simplified RANS Formulations for Velocity Distributions against Full 3D RANS Model, International Journal of Hydraulic Engineering, Vol. 7 No. 2, 2018, pp. 33-42. doi: 10.5923/j.ijhe.20180702.03.
![]() | Figure 1. Definition sketch for the steady uniform flow in a rectangular channel |
![]() | (1) |
and
are mean velocity in the streamwise (x), lateral (y), and vertical (z) directions, respectively;
is gravitational acceleration;
is channel slope;
and
are Reynolds stress tensor components; and ν is fluid kinematic viscosity of fluid. In the central zone (Figure 1), it is assumed that the horizontal gradients (d/dy) are negligible comparing to the vertical gradients (d/dz) which are dominating in this zone, [7]. Therefore, Eq. (1) can be simplified to:![]() | (2) |
![]() | (3) |
is the friction velocity, h is the flow depth and
In most simplified RANS approaches, two additional assumptions were often imposed on Eq. (3). One for modelling the Reynolds shear stress
and the other for approximating the secondary flow term
The third term on the right-hand side of Eq. (3), which reflects the influence of secondary currents, are often approximated using a linear relationship for simplicity, [7]:![]() | (4) |
is a positive coefficient. On the other hand, the Boussinesq assumption are frequently used to model the Reynolds shear stress as following: ![]() | (5) |
is the eddy viscosity. Substituting Eq. (4) and Eq. (5) into Eq. (3), the following partial differential equation (PDE) is obtained:![]() | (6) |
is expressed, different formulations for calculating the velocity distribution can be obtained from the simplified RANS equation, Eq. (6).![]() | (7) |
![]() | (8) |
is the distance from the bed at which the velocity is hypothetically equal to zero and
is he dip-correction parameter. Equation (8) predicts the velocity-dip phenomenon by the term
which includes the dip-correction parameter
[7]. This law returns into the classical log law if 
![]() | (9) |
is Coles parameter representing the wake strength in the boundary turbulent flow. If the eddy viscosity model given by Nezu and Rodi [4], Eq. (9), is used instead of the parabolic profile, the integration and simplification of Eq. (6) would yield the following dip-modified log wake law (DMLW-law): ![]() | (10) |
![]() | (11) |
![]() | (12) |
into the governing equations. These extra unknown quantities comprise the so-called Reynolds stresses. Hence, a turbulence model is then required to account for the Reynolds stresses in order to close the system of equations.![]() | (13) |
is the production of the turbulence,
is the turbulent diffusion due to the fluctuations,
is the diffusion of the Reynolds stresses due to molecular mixing,
is the pressure-strain redistribution term and
is the viscous dissipation of Reynolds stresses. Terms (I), (II) and (IV) contain only mean velocity components and the Reynolds stresses, thus, they do not require modelling when Eq. (13) is used to close the mean flow Eq. (12). On the other hand, terms (III), (V) and (VI) introduce 22 new unknowns into the governing equations, thus they need to be modelled to close the equations. The modelled form of exact transport equation for the Reynolds stresses are written in the CFX as follows, [23]:![]() | (14) |
is modeled within CFX by the following constitutive relation:![]() | (15) |
is the specific dissipation rate,
is the unit tensor (Kronecker’s delta) and
is the mean rate of strain tensor. The production tensor of Reynolds stresses is given by:![]() | (16) |
is given by:![]() | (17) |
![]() | (18) |
![]() | (19) |
|
Therefore,
should be first estimated before application of the models.An empirical formula, which is proposed by Yang et al. [7], could be used for finding
This formula relates the dip-correction parameter to (z/h) and is given as:![]() | (20) |
is calculated by using Eq. (20), [9, 13, 27]. Absi [11] proposed another relation for calculating the dip-correction parameter:![]() | (21) |
is normalized distance of maximum velocity measured from the channel bed. In the central zone, Yan et al. found that
depends closely on the aspect ratio (Ar), and
decreases with the increase of Ar, [7]. Equation (21) produces results consistent well with this fact. Hence, in this study, Eq. (21) was employed to estimate the value of
in both simplified RANS models (i.e. DML-law and DMLW-law).Added to the estimation of the dip-correction parameter, the wake strength parameter (Π) also need to be estimated when the DMLW-law is applied. Π seems to be not universal and its value depends on turbulence structure and the effect of secondary currents. Cebeci and Smith found experimentally that Π increases with Reynolds number in zero-pressure-gradient boundary layers, and at high Reynolds numbers, Π rises to a value of 0.55, [28]. Nezu and Rodi indicated through Laser Doppler anemometry velocity measurements that Π increases significantly with the Reynolds number but becomes nearly constant (Π ≈ 0.2) for
, where
is the mean bulk velocity [4]. Cardoso et al. observed in their experiments on smooth open channel that a wake of small strength (Π ≈ 0.08) occurred in the core of the outer region (0.2 < z/h < 0.7), followed by the retarding effect near the free surface due to the downflow of the secondary currents [5]. This suggests that the secondary currents may influence the wake strength and cause the value of Π to be lower. The flow cases considered in this study are almost narrow channels (except S3) and have a relatively high Reynolds number, therefore the effect of the secondary currents at the center of the flow is considerable. Therefore, a value in a range from 0.0 to 0.2 could be selected for the wake strength parameter. However, in all velocity calculations conducted by the dip-modified laws herein, it was found that the agreement between computed and experimental results was rather better when Π takes a value of 0.2. Hence, the value of Π = 0.2 was used for all test cases considered in the present work.![]() | Figure 2. Schematic of domain geometry and boundary conditions |
of zero). Thus, DML-law may not be able to predict the maximum velocity location and, at the same time, lead to accurate computations of the velocity profile by only adjusting the factor
.![]() | Figure 3. Mean streamwise velocity profile, comparing the simplified RANS model (DML-law) and full RANS model (3D CFD model) with the experimental data from Tominaga et al. [26] |
as shown in Figure 5. Although a modified treatment for the free surface boundary conditions are used, the magnitude of the computed turbulence anisotropy at free surface zone (z/h > 0.7) is comparatively low with respect to the measured data. This may affect predicting the secondary velocity components (V, W) which are responsible for generating the velocity dip. To improve the predicted results from the full RANS (CFD) model, more sophisticated boundary condition is required to impose at the free surface. This makes the application of the full RANS model more impractical for engineering problems compared to using the simplified RANS-based models.![]() | Figure 4. Mean streamwise velocity profile, comparing the simplified RANS model (DMLW-law) and full RANS model (3D CFD model) with the Experimental data from Tominaga et al. [26] |
![]() | Figure 5. Turbulence anisotropy of the normal stresses for case S3 at the channel centre: computed by CFD model and measured by Tominaga & Ezaki [25] |
![]() | (22) |
,
are predicted and measured values respectively, and n is the total number of data in each of the individual profiles.The RMSD of
for all models used and for all test cases are summarized and shown by the bar-graph in Figure 6. It can be seen that the 3D CFD models based on full RANS equations gives nearly the lowest values of RMSD (0.3, 0.4, and 0.6), compared to those obtained for both dip modified formulations based on the simplified RANS equations. However, the DMLW-law, which include the wake effect, may give velocity results with RMSD values (0.5, 0.6 and 0.9) close to those for full RANS model. DML-law gives the greatest RMSD of 1.8 for the case S3 (i.e. for the narrowest channel) while DMLW-law improves the prediction with RMSD of 0.9. Additionally, Figure 6 shows that the velocity profiles computed by conventional log wake law deviate from those measured with RMSD being less or equal to that obtained for DML-law.![]() | Figure 6. Root mean square deviation (RMSD) for predicted velocity from measured data |