JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, F04028, doi: /2012jf002452, 2012

Size: px
Start display at page:

Download "JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, F04028, doi: /2012jf002452, 2012"

Transcription

1 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi: /2012jf002452, 2012 Numerical analysis of the effect of momentum ratio on the dynamics and sediment-entrainment capacity of coherent flow structures at a stream confluence George Constantinescu, 1 Shinjiro Miyawaki, 1 Bruce Rhoads, 2 and Alexander Sukhodolov 3 Received 6 April 2012; revised 11 October 2012; accepted 15 October 2012; published 13 December [1] The flow and turbulence structure at stream confluences are characterized by the formation of a mixing interface (MI) and, in some cases, of streamwise-oriented vortical (SOV) cells flanking the MI. Depending on the junction angle and planform symmetry, as well as the velocity ratio across the MI, the MI can be in the Kelvin-Helmholtz (KH) mode or in the wake mode. In the former case, the MI contains predominantly co-rotating large-scale quasi two-dimensional (2-D) eddies whose growth is driven by the KH instability and vortex pairing. In the latter case, the MI is populated by quasi 2-D eddies with opposing senses of rotation. This study uses eddy resolving simulations to predict details of flow structure for KH- and wake-mode conditions at a confluence for which field measurements are available. Results indicate that SOV cells at this confluence, which occur in both modes, redistribute momentum and mass, enhancing the potential for entrainment of bed material beneath the cells and for extraction of fluid and suspended sediment from the MI. The simulations predict that the cores of some of the primary SOV cells are subject to large-scale bimodal oscillations toward and away from the MI that contribute to amplification of the turbulence close to the MI and enhance the capacity of the SOV cells to entrain sediment. At this confluence, which has a concordant bed and a large angle between the incoming streams - conditions that generate strong adverse lateral pressure gradients adjacent to the MI - the oscillating SOV cells interact with MI eddies to generate large bed friction velocities in the zone of scour immediately downstream of the confluence. Citation: Constantinescu, G., S. Miyawaki, B. Rhoads, and A. Sukhodolov (2012), Numerical analysis of the effect of momentum ratio on the dynamics and sediment-entrainment capacity of coherent flow structures at a stream confluence, J. Geophys. Res., 117,, doi: /2012jf Introduction [2] Confluences are fundamental components of river networks [Paola, 1997] that play an important role in regulating the movement of sediment through these networks [Rhoads and Kenworthy, 1995; Rice, 1998]. The convergence of flow induced by the configuration, or planform, of the two conjoining channels results in highly three-dimensional patterns of fluid motion and the production of turbulence. Following Kenworthy and Rhoads [1995], the region of complex flow in the immediate 1 Departments of Civil and Environmental Engineering and IIHR- Hydroscience and Engineering, University of Iowa, Iowa City, Iowa, USA. 2 Department of Geography, University of Illinois at Urbana-Champaign, Urbana, Illinois, USA. 3 Department of Ecohydrology, Institute of Freshwater Ecology and Inland Fisheries, Berlin, Germany. Corresponding author: G. Constantinescu, Department of Civil and Environmental Engineering, University of Iowa, Iowa City, IA USA. (sconstan@engineering.uiowa.edu) American Geophysical Union. All Rights Reserved /12/2012JF vicinity of the junction is referred to as the confluence hydrodynamic zone. This zone generally extends several channel widths downstream of the upstream junction corner, or confluence apex, where the two channels initially meet. Mixing effects can extend far beyond the confluence hydrodynamic zone [Bouchez et al., 2010]. [3] The structure of flow in the confluence hydrodynamic zone influences patterns of sediment transport, which, in turn, shape the channel bed. This zone typically is characterized by the formation of a large scour hole. Confluence scour constitutes an important environmental and river-management concern (e.g., for the design of bridges and structures anchored on the channel bottom), especially for large rivers. Scour holes can be deep, with depths up to five times greater than the mean depth of the upstream channels, and the orientation of these features can change considerably over time [Best and Ashworth, 1997; Paola, 1997; Rhoads et al., 2009]. The diverse morphologic and flow patterns at confluences also provide a variety of favorable habitat conditions for fish and other aquatic organisms [Rice et al., 2008]. [4] A prominent feature of flow within the confluence hydrodynamic zone (Figure 1) is the formation of a shear 1of21

2 Figure 1. Sketch showing the main types of coherent structures present in the confluence hydrodynamic zone for the case the mixing interface (MI) is in (a) wake mode (momentum ratio is close to unity); (b) Kelvin-Helmholtz (KH) mode (high momentum ratio, Mr 1). The coherence and transverse large-scale oscillations (red arrow) of the SOV cells are generally the strongest in the region where the axes of the two incoming channels reach the MI. For the KH mode, the higher momentum stream pushes the MI toward the bank situated on the other side of the main channel, the MI eddies are co-rotating, the recirculation region is relatively small and the coherence of the main SOV cell is much larger of the high-momentum side. The gray region visualizes the stagnation zone. layer and associated mixing interface (MI). Field and laboratory experiments have examined in considerable detail the large-scale turbulent eddies within the MI and the relation of these eddies to momentum exchange and mixing between the two incoming streams [Best, 1988; Biron et al., 1993b, 1996; Paola, 1997; Rhoads and Kenworthy, 1999; Bradbrook et al., 2000a, 2000b; Sukhodolov and Rhoads, 2001; Rhoads and Sukhodolov, 2001, 2004, 2008]. The main parameters that determine the position and alignment of the MI are the momentum and velocity ratios between the two incoming streams, the magnitude of the angles between the incoming streams and the downstream channel, and whether or not changes in bathymetry at the entrance to the confluence are gradual (concordant bed) or abrupt (non-concordant bed). Flow structure at natural stream confluences is further complicated by topographic steering effects induced by bar forms on the channel bed, by large scale bed roughness [Hardy et al., 2009, 2010] and by largescale irregularities of the channel banks [Rhoads and Sukhodolov, 2001]. [5] Based on theoretical consideration on how vortical structures develop in mixing layers between two parallel streams of different mean velocities and how wake vortices develop behind a bluff body, Constantinescu et al. [2011a] proposed classifying confluence MIs into two types: Kelvin-Helmholtz (KH) and wake modes. Although past studies have generally characterized confluence hydrodynamics using the momentum ratio (Mr =[r 1 Q 1 U 1 ]/[r 2 Q 2 U 2 ] where r i,q i, and U i are the density, discharge, and bulk velocity in the lateral [i = 1] and main [i = 2] tributaries), the occurrence of the KH and wake modes of MI dynamics is governed by transverse fluid shear, indicating that the velocity ratio (Vr =U 1 /U 2 ), rather than Mr, is the most appropriate metric for characterizing these modes. It is possible to have different flows with different momentum flux ratios, but similar velocity ratios; however, for the cases analyzed in this paper Vr corresponds directly to the value of Mr. [6] When Vr differs greatly from unity, the transverse shear across the MI and the Kelvin-Helmholtz (KH) instability play dominant roles in the formation and growth of the large-scale eddies [see also Sukhodolov et al., 2010]. In the KH mode, the MI is populated by co-rotating eddies whose axes are close to vertical relative to the bed (Figure 1b). Similar to the classical case of a mixing layer developing between two streams of unequal velocities, the main mechanism responsible for the increase in the thickness of the MI with the distance from the upstream junction corner is vortex pairing [Winant and Browand, 1974]. [7] When the mean transverse shear across the MI is relatively small (Vr 1) and a zone of flow stagnation exists at the upstream junction corner, the shear layers along the two sides of the confluence apex can interact in a way similar to the separated shear layers (SSLs) forming on the two sides of a bluff body. These interacting shear layers generate large eddies (rollers) containing vorticity of opposite signs, i.e., the MI is populated by eddies with alternating senses of rotation (Figure 1a). In this wake mode, the eddy structure of the MI is similar to the von Karman vortex street that forms behind cylinders and other bluff bodies [Chen and Jirka, 1995]. Flow conditions and confluence planform geometries that favor the development of a large stagnation zone at the upstream junction corner should strengthen the wake mode and enhance the generation of high circulation rollers. As the MI eddies move away from the upstream junction corner, their size increases but no mechanism exists to increase their circulation. This condition contrasts with MIs 2of21

3 in the Kelvin-Helmholtz mode where vortex pairing results in an increase in circulation, enhancing the capacity for MI eddies to erode the channel bed at large distances from the confluence apex. At natural stream confluences both modes can simultaneously influence the development of eddies within the MI, but often one mode will dominate over the other. Flow conditions, confluence geometry, and bathymetry will determine which mode is dominant. [8] Besides the MI, streamwise-oriented vortical (SOV) cells flanking the MI (Figure 1) are another prominent type of coherent flow structure at concordant confluences, i.e., those with channel beds at nearly the same elevation upstream of the confluence [e.g., Rhoads and Kenworthy, 1995; Paola, 1997]. Such cells, or regions of strong helical motion, are expected to form and considerably affect momentum and mass exchange processes at concordant confluences where the angle between the two incoming streams is large. Under such conditions, the two converging streams have significant transverse momentum with respect to the orientation of the MI [Rhoads and Sukhodolov, 2008]. The convergence of the flows results in an increase in pressure, or the elevation of the water surface, within the center of the confluence. The rapid loss of transverse momentum in the region where the two streams collide leads to the development of strong adverse pressure gradients in the center of the confluence adjacent to the MI that steer or turn the flow into the downstream channel. The raised free surface induces strong downwelling of the fluid along the MI. The descending fluid moves laterally along the bed in a divergent pattern on the two sides of the MI and then rises toward the free surface. The result is the formation of a pair of counter-rotating vortices (primary SOV cells) parallel to the orientation of the MI. Depending on the spatial pattern of the pressure gradients in the transverse direction and bathymetry on the two sides of the MI, numerical simulations suggest that the formation of the primary SOV cells may be accompanied by the formation of secondary SOV cells that are co-rotating with the primary SOV cell on each side of the MI [Constantinescu et al., 2011a]. Evidence for the formation of primary SOV cells is quite strong for confluences with a concordant bed [e.g., Ashmore et al., 1992; Rhoads and Kenworthy, 1995, 1998; Rhoads, 1996; Paola, 1997; Rhoads and Sukhodolov, 2001; Sukhodolov et al., 2010]. Their presence and role at confluences with discordant beds are less clear because bed discordance may disrupt [De Serres et al., 1999, Ribeiro et al., 2012] the coherence of the SOV cells. In such cases, MI vortices distorted by the rapidly changing bed morphology are key drivers of flow dynamics and mixing. Even at confluences with strongly coherent SOVs, MI vortices can play an important role in transport and mixing, especially at confluences where the KH mode is dominant. The coherence of the SOV cells will diminish with the angle between the incoming streams and no SOV cells are expected to form at confluences when the two tributaries are nearly parallel an outcome confirmed by Kirkil and Constantinescu [2008, 2009a], who applied the same numerical model used in the present work for the limiting case of a MI developing between two parallel streams of unequal velocities. [9] Laboratory and field studies have led to an improved understanding of confluence dynamics, mixing, and transport. These investigations have provided insight into bed morphology and distributions of mean velocity at cross sections within and downstream of confluences. Some of these studies also provide detailed information on velocity spectra, turbulent stresses, and sediment transport at discrete locations [e.g., Rhoads, 1996; Sukhodolov and Rhoads, 2001; Boyer et al., 2006]. Such studies are, however, limited if the dynamics of the large-scale turbulence and the details of the secondary flow within the entire confluence hydrodynamic zone are of critical interest. The most important limitation is the spatial resolution of the measurements, which typically is not sufficient to completely resolve flow within the SOV cells or turbulent structures generated within the MI. Determining the exact position and extent of these important flow features based on sparse measurements is challenging. [10] Recent advances in the numerical simulation of complex turbulent flows at large (field) Reynolds numbers [e.g., Spalart, 2009] allow an alternative approach to investigate confluence hydrodynamics [Bradbrook et al., 2000b; Constantinescu et al., 2011a]. In the case of stream confluences, results of eddy-resolving numerical simulations can be used to visualize and examine the unsteady dynamics of the large-scale coherent structures (e.g., SOV cells, MI eddies) and to quantitatively assess the effects of these structures on instantaneous and mean bed shear stresses. Such an approach is an important step toward evaluating the capacity of these structures to entrain sediment and, ultimately, to understanding the main mechanisms driving morphological adjustments at confluences. [11] This study uses 3-D eddy-resolving simulations for an asymmetrical (y-shaped) confluence with a concordant bed and a large angle between the two incoming streams to examine in detail the flow physics for two different cases in which the MI is primarily in the wake mode and, alternatively, mainly in the Kelvin-Helmholtz mode. Abundant data from past field investigations at this confluence confirm that flow is characterized by the development of SOV cells. The research builds on the modeling work of Constantinescu et al. [2011a], which examined the effects of the SOV cells on mean flow and turbulence structure for a low momentum ratio test case at this confluence in which the wake mode was dominant. The model predicted that some SOV cells have a large capacity to generate high bed shear stresses, and thus entrain sediment, but did not offer an explanation for this prediction - an issue explored in this paper. In addition, the study presented here investigates numerically the flow physics for a high momentum ratio case in which the Kelvin-Helmholtz mode is dominant. Another novel contribution is an evaluation of the capacity of the MI eddies and SOV cells to transport mass in both cases based on simulations of scalar transport. The simulations address several important sets of questions related to the role of largescale turbulence in momentum and mass exchange processes in the confluence hydrodynamic zone. [12] 1) Do strongly coherent SOV cells form along the MI when the Kelvin-Helmholtz mode is dominant? If so, does the structure of the SOV cells differ from the structure of such cells when the wake mode is dominant? If differences in SOV cells do occur, how do these differences influence the capacity of the cells to entrain bed material compared to the transport capacity of the quasi 2-D eddies convected inside the MI? These questions are important because the 3of21

4 coherence of MI eddies should be substantially enhanced in cases where the Kelvin-Helmholtz mode is dominant. [13] 2) What are the spatial and temporal dynamics of SOV cells under conditions of steady mean flow and how are these dynamics related to turbulence amplification, to MI modes, to the flow inside the MI, to exchange of fluid and fine suspended sediment between the MI and adjacent freestream, and to spatial and temporal patterns of bed shear stress? [14] The numerical method is discussed in section 2. The same section reviews validation of the numerical solver for several types of flows that are relevant to the present investigation of confluence dynamics. Section 3 provides details on the site at which the field experiments were conducted together with the flow conditions in the two simulated test cases. The effect of the momentum ratio on the instantaneous and mean flow structure is outlined in section 4. Section 5 describes the predicted dynamics of the SOV cells. The simulated effect of large-scale coherent turbulence structures on the distribution of the depth-averaged turbulent kinetic energy is discussed in section 6. Section 7 examines the influence of the velocity (momentum) ratio on the potential capacity of the large-scale coherent structures to entrain sediment from the bed and section 8 explores the predicted effect of the SOV cells on mass transport based on the advection of a passive scalar injected close to the upstream junction corner. Section 9 summarizes the main findings and section 10 discusses the relevance of the present results in a more general context. 2. Numerical Model and Validation [15] The simulations in the present study are performed using Detached Eddy Simulation (DES) with the Spalart- Allmaras (SA) Reynolds-Averaged Navier-Stokes (RANS) model as the base model [Spalart, 2000, 2009]. Based on a comparison with well-resolved Large Eddy Simulation (LES), the SA version of DES predicts more accurately mass transport compared to RANS and other versions of DES [Chang et al., 2007]. No special treatment is required to match the solutions at the boundary between the regions where DES is in LES mode and in RANS mode. In particular, the calculation of the bed friction velocity does not rely on the assumption of the presence of a logarithmic layer in the velocity profile, as is the case when wall functions are used. The one-equation SA model solves a transport equation for the modified eddy viscosity, ~n. The boundary condition implementation for rough walls is described in Spalart [2000]. The SA version of DES is obtained by replacing the turbulence length scale d (distance to the nearest wall) in the destruction term of the transport equation for ~n with a new length scale d DES = min(d, C DES D), where the model parameter C DES is equal to 0.65 and D is a measure of the local grid size. When the production and destruction terms in the transport equation for ~n are balanced, the length scale in the LES regions, d DES =C DES D, becomes proportional to the local grid size and yields an eddy viscosity proportional to the mean rate of strain and D 2, as in LES. This approach allows for an energy cascade down to grid size. [16] Detailed descriptions of the viscous solver, governing equations, and DES-SA model used to perform the present simulations are given in Constantinescu and Squires [2004], Chang et al. [2007] and Constantinescu et al. [2011a] and are not repeated here. The governing 3D incompressible Navier-Stokes equations formulated in generalized curvilinear coordinates are discretized on a non-staggered grid and then integrated using a fully implicit fractional-step method. The convective terms in the momentum and the advection-diffusion equations used to simulate scalar transport are discretized using the fifth-order accurate upwind biased scheme. All the other terms in the momentum, scalar transport and pressure-poisson equations are discretized using second-order central differences. Source terms in the turbulence model equations are treated implicitly. The discrete momentum (predictor step), scalar transport and turbulence model equations are integrated in pseudo-time using the alternate direction implicit approximate factorization scheme. Time integration is done using a double time stepping algorithm with local time stepping. The time discretization is second order accurate. The simulations were performed using a parallel version of the code based on Multiple Parallel Interface. [17] The capacity of present DES code to accurately simulate shallow mixing layers, flow in curved channels with naturally deformed bathymetry, and the unsteady dynamics of large-scale vortices induced by strong adverse pressure gradients is directly relevant for the confluence flow studied in the present work. Constantinescu et al. [2011b] reported a detailed validation study of flow in a sharply curved channel (ratio between mean radius of channel curvature and channel width close to 1.3) with equilibrium scour bathymetry at a channel Reynolds number of 68,400. DES successfully captured the redistribution of the streamwise velocity within the bend and the details of secondary flow revealed by the comprehensive experiments of Blanckaert [2010]. DES predictions were a better match to experimental data compared to both RANS and LES with wall functions. [18] The same DES model [Kirkil and Constantinescu, 2008, 2009a] successfully captured the experimentally measured streamwise variation of the centerline displacement and width of a shallow mixing layer developing downstream of a splitter plate separating two parallel turbulent streams of unequal velocities in a flat-bed channel. Analysis of the velocity spectra confirmed the quasi-2d character of the large-scale KH billows within the mixing layer observed in experiment. The spanwise variation of the streamwise velocity profiles across the mixing layer was accurately captured by DES. The configuration considered by Kirkil and Constantinescu [2008] is a limiting case for a river confluence with concordant bed in which the angles between the two tributaries and the downstream channel are close to zero and the MI is in the Kelvin-Helmholtz mode (Case 2). [19] Finally, Constantinescu et al. [2011a] reported a detailed validation study for Case 1 based on the measurements collected as part of a field study by Rhoads and Sukhodolov [2001] and Sukhodolov and Rhoads [2001]. Very good agreement with the field data was obtained for the two-dimensional distributions of the streamwise velocity at the three cross sections (A1, A and C in Figure 2a) where measurements were available. Additionally, Figure 3 provides a direct comparison of the predicted vertical profiles of the streamwise velocity with field measurements at three stations for each of the three sections. The model predicted a 4of21

5 Figure 2. Bathymetry around the confluence between the Kaskaskia River (KR) and Copper Slough (CS). (a) Case 1; (b) Case 2. The bathymetry is visualized using nondimensional bed elevation values, z /D. The datum corresponds to the position of the free surface in Case 1. The free surface in Case 2 is situated at z /D = (c) Also shown is the full computational domain used in simulation of Case 1. narrow zone of high turbulent kinetic energy near the center of the confluence similar to that observed in the pattern of field data (see Figure 9 in Constantinescu et al. [2011a]). The exact locations of the predicted and measured zones of high TKE show some differences, but overall the DES predictions are greatly superior to RANS simulations, which fail to predict important features of the measured distributions of TKE and secondary flow (Figure 10 in Constantinescu et al. [2011a]). In particular, RANS cannot reproduce accurately the coherence and circulation of the SOV cells which results in underprediction of TKE within the MI and its vicinity by 1 2 orders of magnitude. Possible reasons for the discrepancy in the position of the zone of high TKE for field measurements and DES simulation include failure of the model to capture completely the influence of bathymetric and bank conditions on the flow field due to the rather coarse resolution of the topographic information on which the computational grid is based and/or failure of the field data to capture fully the effects of lowfrequency oscillations of coherent structures in the vicinity of the MI during the measurement interval ( 90 s). 3. Test Cases and Simulations Setup [20] The flow at the confluence of Kaskaskia River (KR, west [W] side of the confluence) and Copper Slough (CS, east [E] side of the confluence) in east central Illinois, U.S. A. is chosen as a representative test case of an asymmetrical confluence with a concordant bed. The cross sections of both tributaries upstream of the confluence are trapezoidal. The upstream channel for the Kaskaskia River is fairly well aligned with the downstream channel. The Copper Slough joins the Kaskaskia River at an angle of about 60. The inner and the outer banks correspond to the east (E) and the west (W) banks, respectively. The radius of curvature of the flow around the east bank divided by the mean width of the downstream channel is close to 3.5. Results of field research at this confluence are reported by Rhoads [1996], Rhoads and Kenworthy [1995, 1998], Rhoads and Sukhodolov [2001, 2004, 2008], Rhoads et al. [2009], and Sukhodolov and Rhoads [2001]. The present paper describes the results of simulations for two test cases corresponding to confluence bathymetry (Figure 2) and flow conditions in the two incoming streams (Table 1) as documented on May 27, 1998 [Rhoads and Sukhodolov, 2001; Sukhodolov and Rhoads, 2001] and on May 26, 1999 [Rhoads and Sukhodolov, 2008]. Little or no bed material was transported by the flow during these events and no bed forms were present on the bed. Typical of flows at this stage [Kenworthy and Rhoads, 1995], suspended sediment consisted entirely of wash load (silt, clay, very fine sand). Detailed information on the textural characteristics of the bed material is given in Rhoads et al. [2009]. The main difference between the two test cases is the magnitude of Mr, which is about one (1.03) for Case 1 and about 5.4 for Case 2. In the latter test case, Copper Slough has much more momentum than does 5of21

6 Figure 3. Comparison of predicted mean streamwise velocity (solid line) with the measurements (symbols) for Case 1. The profiles are compared at several stations within cross sections (top) A1, (middle) A and (bottom) C. For a given cross section, the nondimensional transverse distance from the west bank line is indicated at the top of each frame. Each vertical profile obtained from the numerical simulation (solid line) starts at the bed and ends at the free surface (z/d = 0). Kaskaskia River. Thus, the wake mode is dominant in Case 1, while the Kelvin-Helmholtz mode is dominant in Case 2. The physical Reynolds number defined using average values of flow velocity and depth in the downstream channel is 166,000 for Case 1 and 77,000 for Case 2. For both test cases, all quantities are non-dimensionalized using the mean velocity (U) and the mean flow depth in the downstream channel (D) for Case 1, i.e., U = 0.45 m/s and D = 0.36 m. [21] Bathymetry data were collected within the confluence hydrodynamic zone with an average resolution of 2 m in the streamwise direction and 1 m in the transverse direction. This resolution includes all major bathymetric features (bars, scour hole, failed blocks of bank material) within the confluence. The bathymetry transects available within the region of interest from the field experiment were used to define the computational domain. No smoothing was applied on the measured bathymetry profiles except near the banks. As can be seen from Figure 2, the minimum bed elevations of the scour holes in the two test cases measured with respect to the position of the free surface in Case 1 (z = 0.0D) were Table 1. Main Flow Parameters of the Confluence Flow in Case 1 and Case 2 Case Stream Q i (m3/s) U i (m/s) rq i U i (kg m/s) D i (m) Re Fr Mr 1 Ave. a , KR CR Ave. a , KR CR a Ave. row contains variables calculated using the average value of the bulk velocity and mean depth in the channel downstream of the junction between the two incoming streams Kaskaskia River (KR) and Copper Slough (CS). 6of21

7 about the same ( 3.0D), but the volume of the scour hole in Case 2 was smaller than the volume in Case 1. The mean flow depth in Case 2 was about two thirds of that in Case 1. The mean width of the downstream channel was close to 25D in both cases. A submerged block of failed bank material was present in Case 1 along the west bank of the confluence, close to where the Kaskaskia River enters the junction (see bathymetry in Figure 2a around Section A3). In Case 2, the bank line at this same location was much different. Furthermore, between 1998 and 1999 a submerged bar developed close to the east bank between Sections A and C; thus, the flow in Case 2 was much shallower in this region compared to Case 1. [22] The computational domain (e.g., see Figure 2c for Case 1) is meshed with close to 5 million cells. The number of grid points in the vertical direction is close to 30. This domain adequately represented individual coherent structures in the flow. A typical computational cell has dimensions of 0.05D 0.1D in the spanwise direction within and close to the MI, 0.1D 0.25 D in the streamwise direction and less than 0.04D in the vertical direction. The wall-normal grid spacing of the first row of cells off the bed and the banks is less than two wall units based on the average value of the bed friction velocity in the incoming channels. Using information from the field studies, the channel bed and banks are treated as rough surfaces with a mean value of the roughness height equal to 0.01 m a value consistent with the 84th percentile (d 84 ) of bed material in the confluence [Rhoads et al., 2009], a common metric of roughness scale. The 50th percentile (d 50 ) of bed material in the confluence was 4 mm. [23] For both test cases, inflow conditions corresponding to fully developed turbulent channel flow with resolved turbulent fluctuations are applied. These conditions conform with those in the field experiment where the channel reaches upstream of the confluence are long and straight. The inlet section is situated about two channel widths upstream of the confluence apex in both tributaries (Figure 2c). For each inflow section, the turbulent fluctuations (zero mean velocity) are obtained from a precursor well-resolved LES simulation of fully developed turbulent flow in a straight channel conducted at a physical channel Reynolds number of 20,000. This low Reynolds number was used because LES without wall functions is computationally expensive at the field Reynolds number. The LES fluctuations in the two incoming streams were scaled such that average value of the turbulent intensity away from the wall surfaces matched the value measured in the field experiment (5 7%). The scaled fluctuations were then added to the mean velocity profile obtained from a RANS simulation of fully developed turbulent flow for the Reynolds number and geometrical conditions in the corresponding incoming stream (Table 1). The total velocity fields (RANS mean plus LES fluctuations) were stored in a file and then fed in a time-accurate manner through the two inflow sections. The presence of a minimum amount of resolved turbulence in the incoming flow is critical because of the perturbations this turbulence induces in the shear layers on the two sides of the confluence apex [see also Chang et al., 2006, 2007]. [24] A convective boundary condition is used at the outflow of the domain. The outflow section is situated at about two channel widths downstream of Section C (Figure 2c) to minimize effects on flow within the confluence hydrodynamic zone. The free surface is modeled as a shear-free rigid lid. Given that the Froude numbers defined with the average depth and velocity in the incoming and downstream channels is less than 0.3 in both test cases (Table 1), this approximation is acceptable. Though the model does not directly simulate variations in water surface elevation (e.g., superelevation within the confluence), the model does account for such effects in an approximate way by predicting non-uniform pressure distributions on the rigid-lid surface. [25] Transport of mass in the vicinity of the MI is investigated by solving an advection-diffusion equation for a passive (conserved) tracer similar to the approach used by Bradbrook et al. [2000b] and Biron and Lane [2008]. A tracer with a normalized concentration CSS = 1 is introduced continuously at all flow depths in two small regions situated about 4D upstream of the junction corner and close to the two banks that form the junction corner. The concentration of the tracer in the two incoming streams is CSS = 0. The flux of tracer is set equal to zero at the channel bottom, free surface, and the banks. The time step is 0.1 D/U. Statistics were collected over an interval of 900D/U after the flow and tracer transport reached a statistically steady state (500D/U) and were checked for convergence. [26] Estimates of the eddy (vortex) diameter are obtained by first calculating the area of high (positive or negative depending on the sense of rotation of the eddy) along-theaxis vorticity component within a plane perpendicular to the axis of the vortex. The value of this area, along with the assumption that each eddy has nearly circular shape, is then used to calculate the vortex diameter. 4. Effect of Momentum Ratio on Instantaneous and Mean Flow Structure [27] Values of the momentum and velocity ratios of the two tributaries close to unity as well as the formation of a zone of flow stagnation in the vicinity of the upstream junction corner in Case 1 favor the development of a strong wake mode that allows: 1) the Kelvin Helmholtz instability inside the separated shear layers to grow fast and develop into strongly coherent tube-like structures; and 2) the vortex tubes inside the two separated shear layers to interact and shed large eddies into the MI. Generally this shedding happens when the boundary layers on one or both sides of the confluence apex separate upstream of the junction corner and the size of the stagnation zone is large. The streamlined shape of the junction corner region in Case 2 (Figure 2b) reduces the size of the stagnation zone forming between the two separated shear layers and the confluence apex and constrains the development of large-scale eddies inside the separated shear layer on the Kaskaskia River side, which detaches from the channel bank upstream of the confluence apex (Figure 4b). By contrast, in Case 1 both separated shear layers contain strongly coherent eddies and the size of the stagnation zone is larger than in Case 2 at all flow depths. The larger size of the stagnation zone in Case 1 is primarily due to upstream separation of the boundary layer on the CS side and its movement away from the adjacent bank. [28] The proximity of the two separated shear layers and the small angle between these separated shear layers explain why the shapes of the billows shed downstream of the 7of21

8 Figure 4. Distribution of the vertical vorticity, w z D/U, in the instantaneous flow in a horizontal surface situated 0.1 D below the free surface. (a) Case 1; (b) Case 2. The black dash-dot line visualizes the shear layer forming at the east bank due to the strong curvature of the inner bank and the presence of a submerged bar of deposited sediment in Case 2. The green dashed line visualizes the SSL forming in Case 1 at the west bank due to the presence of a submerged bar of failed bank material near Section A3 (see also Figure 2a). The black dashed line follows the centerline of the Copper Slough (CS) stream until it intersects the MI. Figure is redrawn from Miyawaki et al. [2010]. stagnation zone in Case 1 are elongated compared to the form of billows observed in the wake of bluff bodies (e.g., circular cylinders). The mean velocity distributions show that, close to the free surface, the differential shear across the MI is around U downstream of Section A1. In other words, a weak Kelvin-Helmholtz mode is present over part of the MI in Case 1. By contrast, the most upstream part of the MI in Case 2 resembles a vorticity sheet (Figure 4b). Under these conditions, the wake mode is negligible and the perturbations of the vorticity sheet by the incoming turbulence on the two sides of the MI are not sufficient to generate large-scale vortices (Kelvin-Helmholtz billows) even though the differential shear across the vorticity sheet is large. Large-scale billows are only first observed around Section A1 for Case 2 (see also Figure 19a). [29] Compared to Case 1 where the MI is located within the central part of the downstream channel, the large momentum ratio in Case 2 and deflection of shallow flow near the east bank (flow depth less than 10% of the mean flow depth in the downstream channel) by a submerged bar shift the MI toward the west bank. The position of the MI in both cases is in good agreement with the location inferred from field data [Rhoads and Sukhodolov, 2001, 2008; Sukhodolov and Rhoads, 2001]. Near Section C, the trajectories of the MI eddies in Case 2 are deflected by the west bank. Around Section C, these eddies attain a diameter of about 10 times the mean flow depth in the downstream channel. Analysis of the temporal evolution of the instantaneous flow fields shows that, on average, two vortex-pairing events occur among neighboring Kelvin-Helmholtz billows between Section A2, where the width of the MI is less than two times the mean flow depth, and Section C. The width of the MI is estimated in an approximate way by determining the width of the region where MI eddies visualized using both the out-of-plane vorticity and a passive scalar introduced at the confluence apex (see section 8) are convected in the instantaneous flow fields. Although the average diameter of the MI eddies around Section C is considerably smaller in Case 1 than in Case 2, shallow flow effects become significant in Case 1 as the average diameter of the eddies is close to four times the mean flow depth. [30] The dynamics and coherence of the MI eddies at and downstream of Section C in Case 1 are strongly affected by their interaction with: 1) the large-scale turbulence originating in the region where the incoming flow in Kaskaskia River is advected over a submerged block of failed bank material along the west bank near Section A3 (Figures 2a and 4a); and 2) the narrow shear layer that develops due to an abrupt lateral decrease in streamwise velocity toward the east bank downstream of Section A1 induced by sharp curvature of this bank (Figure 2a). The development of inner bank shear layers is a general characteristic of flow in sharply curved open channels with flat or deformed beds [Constantinescu et al., 2011b]. The core of high momentum fluid parallel to the banks of the Copper Slough upstream of the region of high bank curvature does not change direction fast enough for this core to follow the bank line. A shear layer forms between the core of high momentum fluid and fluid moving slowly in the same direction near the inner bank. [31] The water level in Case 2 is sufficiently low such that the flow does not submerge the block of failed material around Section A3 (Figures 2b and 4b). This is why largescale energetic eddies are absent on the Kaskaskia River side of the downstream channel in Case 2. The flow separates as it is deflected laterally by the submerged bar along the inner bank between Sections A and C. A shear layer of positive vorticity forms at about the same location as the shear layer induced by abrupt bank curvature in Case 1. However, its effect on the structure of the flow within the MI is negligible because the MI is displaced toward the west bank in Case 2. [32] The formation of strong SOV cells at the KRCS confluence is expected due to the large angle between the Copper Slough and the downstream channel. The system of SOV cells predicted in the two cases is depicted in Figure 5 using the Q criterion [Hunt et al., 1988]. The quantity Q is the second invariant of the (resolved in DES) velocity 8of21

9 Figure 5. Visualization of the main vortical structures in the mean flow using a Q iso-surface (Q = 10). (a) Case 1; (b) Case 2. The 3-D ribbons visualize the helical motion of the particles inside the SOV cells. The red segments visualize the approximate position of the MI close to the free surface in Sections A1, A and C. Figure is redrawn from Miyawaki et al. [2010] and Constantinescu et al. [2011a]. gradient tensor (Q = 0.5 u i / x j u j / x i D 2 /U 2 ) and represents the balance between the rotation rate and the strain rate. The same value of Q (=10) was used in all analyses to allow relative estimation of the size and coherence of the coherent structures in the two cases. The primary SOV cells on the two sides of the MI are denoted SVE1 (East side) and SVW1 (West side). Their axes follow the boundaries of the MI. The secondary SOV cells are denoted SVE2 and SVW2. The cores of the SOV cells in the mean flow are fairly circular and are regions of high streamwise vorticity (see Figure 6 for Section A1 in Case 1). The high-vorticity fluid inside their cores is transported downstream, but the streamwise velocity in the core of the SOV cells is smaller that that of the surrounding flow by about 20 30%. This difference is especially evident in the streamwise velocity distributions for the instantaneous flow fields (Figure 6). The same is true of the streamwise velocity distributions for the mean flow (Figure 6), but the difference between streamwise velocities within and outside of the vorticity cores is smaller because of large-scale temporal oscillations of the SOV cells. [33] Important differences are observed in the system of SOV cells in the two cases. In Case 1, the flow in the Copper Slough approaches the MI at a large angle close to the upstream junction corner and requires a certain distance to deflect the MI toward the west bank. Meanwhile, flow in the Kaskaskia River is initially aligned with the downstream channel. The adverse pressure gradients needed to decelerate the transverse component (relative to the MI) of the flow on the Copper Slough side of the MI close to the upstream junction corner are much larger than the requisite gradients on the Kaskaskia River side of the MI. Thus, the circulation of SVE1 is 2 3 times larger than that of SVW1 close to Section A3. The difference between the circulations of SVE1 and SVW1 decreases until Section A1 where the circulations of the two SOV cells are close to equal. This change in relative circulation strength is not surprising because the momentum of the two incoming streams and the angles between the MI and the two incoming streams downstream of Section A2 are about the same. The coherence of the secondary SOV cells on both sides of the MI is considerably smaller than that of the primary SOV cells. [34] In Case 2, the coherence of the SOV cells is strongly dependent on the velocity and momentum of the two incoming streams. The core of high velocities on the Copper Slough side rapidly loses transverse momentum as it approaches the MI. Strong adverse pressure gradients are generated in the transverse direction relative to the MI and Figure 6. Flow structure at Section A1 for Case 1 as visualized by: (a) streamwise-oriented vorticity, w s D/U, mean flow; (b) streamwise velocity, u s /U, mean flow; (c) streamwise velocity, u s /U, instantaneous flow. A Q isosurface (Q = 10) is used to visualize the cores of the clockwise rotating (dash line) and counter-clockwise rotating (solid line) SOV cells (see also Figure 5a). MW denotes the local width of the MI. The dotted rectangle in frame a shows the extent of region visualized in Figure 10. Mean streamwise flow is toward the viewer. Frame a is taken from Constantinescu et al. [2011a]. 9of21

10 Figure 7. Distribution of the mean pressure fluctuations, 10 3 p 2 =r 2 U 4, at Sections A1, A and C for Case 1. A Q isosurface (Q = 10) is used to visualize the cores of the clockwise rotating (dash line) and counter-clockwise rotating (solid line) SOV cells (see also Figure 5a). MW denotes the local width of the MI. The arrows point toward the location of the pressure peaks. Two distinct peaks are observed in the region where the core of SVE1 is subject to bimodal oscillations at Sections A1 and C. Even at Section A, where only one peak is observed, the region of high p 2 =r 2 U 4 induced by the oscillations of SVE1 is very elongated in the spanwise direction. Mean streamwise flow is toward the viewer. induce the formation of strongly coherent SOV cells on the high-momentum side. The SOV cells on the Kaskaskia River side are quite weak even in the region between Sections A3 and A where the core of large velocities within the Kaskaskia River approaches the MI at a high angle. The ratio between the circulations of SVE1 and SVW1 is close to 3.5 in this region. Downstream of Section A1, the core of high velocity flow on the Kaskaskia River side becomes relatively well aligned with the MI. The low transverse momentum of flow on the Kaskaskia River side of the MI induces a rapid decay of the circulation and coherence of SVW1. Thus, the circulation of SVE1 is about eight times larger than that of SVW1 at Section A. Another important difference with the structure of the SOV system observed in Case 1 is that the circulation of SVE1 is larger than that of SVE2 only until Section A1. Downstream of Section A, the coherence of SVE1 decays faster than that of SVE2. Meanwhile, SVE2 moves into the deepest part of the section. [35] Thus, the differences in the value of the momentum ratio and, to a certain extent, the bathymetry between the two cases are responsible for the change in the primary mechanism responsible for the formation and growth of the quasi 2-D eddies within the mixing interface, the changes in the position of the interface and the coherence of the SOV cells on the two sides of the interface. [36] Analysis of the distributions of the turbulent kinetic energy, k, and mean pressure fluctuations, p 2, in the two cases shows that the turbulence is strongly amplified inside the cores of the most coherent SOV cells compared to the levels observed in the surrounding turbulent flow, especially over the region where the two incoming streams collide. In particular, for Case 1 the largest values of k and p 2 are observed within and adjacent to the core of SVE1. The turbulence intensity in this region is 4 5 times larger than that predicted inside the MI and at least two times larger than values in the region occupied by the primary SOV cell on the other side of the MI (Figure 7). The levels of turbulence amplification inside the secondary SOV cells (e.g., SVE2) with respect to the background levels are negligible. As discussed in the next section, the reason for the large amplification of the turbulence within and near the core of SVE1 is directly related to the oscillatory motion of SVE1. [37] Although a well-defined scour hole exists for both cases, the model did not predict separation of flow from the channel bed as it moves through the zone of scour. This prediction is consistent with the field data, which also did not document flow separation from the bed within the scour hole [Kenworthy and Rhoads, 1995; Rhoads, 1996; Rhoads and Kenworthy, 1998; Rhoads and Sukhodolov, 2001]. The lack of separation is attributable to the small adverse pressure gradients along paths of the flow streamlines, which are nearly parallel to the side slopes of the scour hole. Thus, the rate of change in bed elevation along these flow paths is too small to produce flow separation. 5. Dynamics of the SOV Cells [38] The simulations show that in the region where the transverse (relative to the MI) pressure gradients are strongest, the cores of the primary SOVs are often subject to large-scale aperiodic bimodal oscillations, similar to the ones observed for necklace vortices in junction flows. To determine the bimodal nature of these oscillations, the approach of Devenport and Simpson [1990] is used to analyze histograms of the velocity components in the directions (vertical or transverse in the case of SOV cells) that do not coincide with the axis of the vortex inside the region where the core of that vortex oscillates. If the histograms show a double-peaked shape, then the oscillations are bimodal, meaning the core of the vortex oscillates between two extreme modes associated with the two peaks in the velocity histograms. The switch from one mode to the other does not take place at a definite frequency. Analysis here focuses on the structure of the flow around the SOV vortices subject to 10 of 21

11 Figure 8. Probability-density function of the vertical component of the instantaneous velocity, v/u, for Case 1 at a point situated in the Kaskaskia River upstream of the confluence (mid-depth and mid span). bimodal oscillations and the mechanisms that are responsible for transition from one mode to the other. [39] For Case 1, a peculiar characteristic of the region of high amplification of p 2 within the region where the core of SVE1 oscillates is that it contains two distinct zones of maximum p 2 (e.g., see Sections A1 and C in Figure 7). Even in Section A where the region contains only one maximum, the shape of this zone is fairly elliptical a result expected when the distance between the centers of two separate circular patches of high p 2 becomes smaller than the mean radii of the patches. The reason for this variation of p 2 inside the horseshoe vortex region is that the core of SVE1, rather than randomly moving around its mean flow (timeaveraged) position, is subject to large-scale bimodal oscillations toward and away the MI, similar to those observed for the necklace vortices in turbulent junction flows [e.g., Devenport and Simpson, 1990; Simpson, 2001]. Eddy resolving simulations conducted for bluff bodies mounted on a flat bed [e.g., Koken and Constantinescu, 2009; Kirkil and Constantinescu, 2009b] and on scoured deformed bed [e.g., Kirkil et al., 2009; Koken and Constantinescu, 2008; Kirkil and Constantinescu, 2010] have shown that the twin maxima in the spatial distribution of p 2 in the region where the primary necklace vortex oscillates are a strong indication that these oscillations are bimodal. Thus, the validity of a similar explanation for the dynamics of SVE1 has to be explored further through analysis of the velocity histograms. [40] In a turbulent flow without large-scale coherent vortices, the histograms of the velocity fluctuations in a certain direction contain only one sharp peak around the mean value of the velocity component and their shapes are fairly symmetric. The velocity histogram calculated at a point within the Kaskaskia River upstream of the junction where the open channel flow is fully developed and the mean value of the vertical velocity is close to zero exhibits such a shape (Figure 8). Even when the flow contains large scale eddies that pass quasi-regularly through a certain region (e.g., MI quasi 2-D eddies), the histograms at points situated within Figure 9. Probability-density functions of the vertical component of the instantaneous velocity, v/u, for Case 1 at a point situated close to the axis of SVE1 in the mean flow. (a) Section A1; (b) Section A; (c) Section C. The peaks associated with the bank mode (BM) and the interface mode (IM) are shown with arrows. The double-peaked histograms of the instantaneous velocity confirm the bimodal nature of the large-scale oscillations of SVE1 for Case of 21

12 Figure 10. Velocity vectors and streamwise-oriented vorticity, w s D/U, in Section A1 (Case 1) at two time instances when: (a) SVE1 is in the interface mode; (b) SVE1 is in the bank mode. The dotted rectangle in Figure 6a shows the extent of region visualized in Figure 10. The shaded patch shows the change in position and shape of the core of SVE1 as it switches between the two modes associated with the bimodal oscillations. The clockwise-rotating bottom attached vortex, BAW, is shown using a dash line. Mean streamwise flow is toward the viewer. this region contain only one peak, but the decay away from the mean value is not as sharp as that for flows without large eddies. [41] The histograms at points situated inside the core of SVE1 are bimodal at all sections within the confluence hydrodynamic zone (e.g., see Figure 9). The separation between the two peaks associated with the predominant states, or modes, decays past Section A1 in the downstream direction, suggesting that the average amplitude of the bimodal oscillations decays gradually past the region where the high velocity cores of the two incoming streams converge. For Case 1, this region is situated around the location (between Sections A2 and A1) where the centerlines of the two incoming tributary channels intersect the MI. Within this region, the two converging streams lose the largest amount of transverse momentum as flow approaches the MI and rapidly changes its direction from the orientation of the tributary to the orientation of the downstream channel. To produce this change in direction, large adverse pressure gradients are generated in the transverse direction in the vicinity of the two sides of the MI, which results in formation of strong SOV cells and the generation of bimodal oscillations for the most coherent of these vortices. Histograms at points situated inside the core of SVW1 display two closely spaced peaks, which suggest SVW1 is subject to weak bimodal oscillations. This weak bimodality is consistent with the fairly circular shape of the region of high p 2 values present around the location of SVW1 in Section A (Figure 7). At all sections, the values of p 2 inside the region where SVW1 oscillates are smaller than those inside the corresponding region for SVE1. [42] The mechanisms responsible for the formation of the primary and secondary SOV cells and the generation of bimodal oscillations in regions of strong adverse pressure gradients are, in many respects, similar to those responsible for the formation and unsteady dynamics of necklace vortices generated within the horseshoe-vortex system that develops at the base of a surface-mounted bluff body. In both cases the bimodal oscillations of the vortex is the main reason for the large amplification of the turbulence (e.g., up to one order of magnitude for p 2 and Reynolds stresses) compared to the background levels and for the oscillating vortex to induce high bed shear stresses beneath it. The oscillations involve movement of the vortices toward and away from a boundary and are driven by the adverse pressure gradients in the direction perpendicular to that boundary. In the case of junction flows, the relevant boundary is the upstream no-slip surface of a bed-mounted obstacle. This boundary is fixed in time and space and is also independent of the flow. In the case of a river confluence, the boundary is the MI, the position of which is determined by the confluence geometry and flow conditions. The boundary is dynamic and contains large-scale eddies that can interact with the core of SVE1 when the vortex gets close to the MI. In the case of junction flows, the primary and secondary necklace vortices in front of the obstacle follow approximately the junction line between the bed and the obstacle, while the legs of the vortices tend to align with the incoming flow direction. In the case of river confluences, the SOV cells are parallel to the internal flow boundary, but as the MI extends past the region where the two streams converge, the SOV cells remain parallel to the MI over their entire length. While for junction flows the main flow direction is, in most cases, perpendicular to the boundary, or upstream face of the obstacle, at stream confluences the main flow direction is oriented obliquely in relation to the boundary, in this case the MI. Thus, at confluences only the transverse component (with respect to the local orientation of the MI) of the incoming flow contributes to the generation of SOV cells. [43] The bimodal character of the oscillation of SOV cells can be discerned by examining the structure of the flow at 12 of 21

13 Figure 11. Flow structure in Section A1 for Case 2. (a) streamwise-oriented vorticity, w s D/U, mean flow; (b) mean pressure fluctuations, 10 3 p 2 =r 2 U 4. The solid lines show the cores of the SOV cells as visualized by a Q isosurface (Q = 10, see also Figure 5b). MW denotes the local width of the MI. The black arrows point toward the two regions of amplification of the mean pressure fluctuations induced by the bimodal oscillations of SVE1. Mean streamwise flow is toward the viewer. Section A1 at two different times (Figure 10). In the first instance, which is representative of times when SVE1 is in what is referred to here as the interface mode (IM), SVE1 is situated close to the MI and does not exchange vorticity with SVE2. The core of SVE1 is fairly circular and its interaction with the bottom-attached boundary layer induces the formation of a bottom-attached cell called BAW that separates SVE1 from SVE2 and rotates in opposite direction. During the transition to the other preferred state, which is referred to as the bank mode (BM), the circulation of SVE1 increases. The core of SVE1 moves toward BAW and starts extracting vorticity of the same sign from the core of SVE2. At Section A1, where the energy of the bimodal oscillations is the largest, the transition to the bank mode generally results in the complete merging of SVE2 and SVE1. The core of SVE1 increases in size and assumes a more elliptical shape. A strong jet-like flow oriented parallel to the channel bottom develops beneath SVE1. As the intensity of the jet-like flow decreases, SVE1 moves closer to the MI, initiating the transition to the interface mode. [44] The two modes are somewhat similar to two modes observed in junction flows: the zero-flow mode, in which the main necklace vortex is fairly circular and close to the upstream face of the obstacle and the back flow mode, in which the core of the necklace vortex acquires an elliptical shape and is situated away from the obstacle. Devenport and Simpson [1990] have shown that the transition to the zeroflow mode is generally triggered by the injection of a patch of high-vorticity and low-momentum fluid moving toward the obstacle first into the downflow parallel to the face of the obstacle and from there into the necklace vortex. The transition to the back flow mode generally occurs when a patch of low-vorticity and high-momentum fluid from the upper part of the channel is injected into the downflow. [45] The injection of patches of high and low vorticity from the two incoming streams moving toward the MI modulates the coherence of the SOV cells in a manner similar to that observed for junction flows. However, in the case of a stream confluence the mechanisms triggering the transition to the two modes are more complex than those for junction flows. Because secondary flow within the MI involves movement of fluid toward the bed and away from the bed toward the two banks, the presence of a strongly coherent SOV cell along the MI results in the injection of near-bed diverging flow into the cell core along with MI eddies advected laterally by this flow. This mechanism appears to be equally, if not more, effective in starting the transition to the bank mode than the mechanism related to injection of eddies from the incoming channel. Moreover, the interactions between quasi-vertical MI eddies and the streamwise oriented SOV cell results in the generation of 3D turbulent eddies, which contribute strongly to the increase of the turbulent kinetic energy. Analysis of the instantaneous flow fields shows that SVE1 spends significantly more time in the interface mode compared to the bank mode. Similar to the case of junction flows, the transition from one mode to the other does not happen quasi periodically. Rather, it happens at unevenly spaced time intervals. The average frequency associated with the switching in Case 1 (0.035Hz) is several times larger than the average frequency corresponding to passage of the MI eddies in Case 1 (0.16Hz and 0.23Hz). It is also is consistent with the frequency of a prominent peak in the transverse velocity spectra ( 0.04Hz) based on velocity time series measurements (300 s time series available at selected locations) obtained within or on the margins of the MI in the field experiment corresponding to Case 1 [Rhoads and Sukhodolov, 2004]. [46] Bimodal oscillations of the core of SVE1 are also observed in Case 2. The centerline of the Copper Slough channel containing high momentum fluid intersects the MI between Sections A1 and A, where the coherence of SVE1 is the largest. Because the MI is in the Kelvin-Helmholtz mode, the largest values of p 2 are observed within the MI rather than the region occupied by the most coherent SOV cells, as was the case for Case 1. Still, a clear amplification of p 2 is observed around the region where the core of SVE1 oscillates. The regions of amplification of p 2 induced by these oscillations between Sections A1 and A contains two peaks (e.g., see Figure 11 for Section A1), which suggests that the large scale oscillations have a bimodal character a conclusion confirmed by the velocity histograms, which are double peaked between Sections A1 and A (e.g., see Figure 12a). In contrast to Case 1, bimodal oscillations of SVE1 are not observed downstream of Section B. For example, the histogram at Section C in Figure 12b contains only one distinct peak. In terms of the flow structure during the interface mode and the bank mode, the largest differences with Case 1 are observed during the bank mode. Though SVE1 extracts vorticity from the core of SVE2 during the time SVE1 is in the bank mode, no local merging of the two vortices is evident. The lateral displacement of the core of SVE1 as it switches between the interface mode and the bank mode is, on average, smaller than for Case 1. The overall lower energy and amplitude of the bimodal oscillations in Case 2 may also be due to the different structure of the MI. Because the coherence of the Kelvin Helmholtz billows is much greater than that of the eddies advected within a MI in which the wake mode is dominant, the SOV cells on the two sides of the MI are less efficient in 13 of 21

14 Figure 12. Probability-density functions of the vertical component of the instantaneous velocity, v/u, for Case 2 at a point situated close to the axis of SVE1 in the mean flow in a given cross section. (a) Section A; (b) Section C. The peaks associated with the bank mode (BM) and the interface mode (IM) are shown with arrows. The histograms of the instantaneous velocity show that the large-scale oscillations of SVE1 are bimodal at Section A but not at Section C. extracting turbulent eddies from the MI region a mechanism that can trigger the transition to bank mode. 6. Effect of Large-Scale Coherent Structures on Turbulence Statistics [47] Variation in the structure of the MI, the coherence of the SOV cells, and the spatial extent of bimodal oscillations of the SOV cells explain important qualitative and quantitative differences in the patterns of the nondimensional depth-averaged turbulent kinetic energy, ^k=u 2 for the two test cases (Figure 13). In both cases, the amplification of ^k is primarily due to the passage of coherent turbulent structures in a certain flow region and oscillations of large-scale coherent structures that are much lower in frequency than velocity fluctuations associated with small-scale turbulence. In Case 1, the convection of strong vortical eddies inside the separated shear layers and their interaction that results in the formation of the large-scale vortices for a MI in wake mode account for the large values of ^k around the upstream junction corner and within the recirculation zone. Outside of this region, large values of ^k occur along the entire length of SVE1 and, in particular, between Sections A2 and A1 where the amplitude of the bimodal oscillations is greatest. The distribution of the TKE in these cross sections (Figure 9 in Constantinescu et al. [2011a]) show the peak values occur in the region that corresponds to the position of SVE1 in the mean flow. However, amplification of ^k in the region situated between the centerline of the MI and SVE1 is due primarily to the passage of the MI eddies and their interactions with the cores of the SOV cells (see discussion in section 5). The levels of ^k within the regions occupied by SVW1 and SVE2 are comparable and are about % smaller than the magnitude of ^k values inside the region occupied by SVE1. The same is true for the separated shear layer induced by the submerged bar of failed bank material at the west bank over which strong turbulent eddies are convected (see also Figure 14). Past Section B, the region of amplified depth-averaged turbulent kinetic energy (^k > 0.01 U 2 ) relative to levels in the incoming streams occupies more than half of the channel section. The cores of the SOV cells are subject to large-scale instabilities that result in the break up of the downstream part of the core into a succession of streamwise-oriented streaks of highly vortical fluid (Figure 14). As they are convected downstream, these streaks are stretched and change orientation due to interactions with the MI eddies and with the 3-D eddies originating in the separated shear layer along the west bank. [48] Comparison of the eddy content of the instantaneous flow in the two cases shows that the coherence of the MI eddies is significantly greater in Case 2 (Figure 15) than in Case 1 (Figure 14), not only close to the upstream junction corner, but also farther downstream due to merging of corotating billows. This enhanced coherence is the main reason why in Case 2 the largest values of ^k are observed toward the inner-bank side of the MI rather than the region occupied by Figure 13. Distribution of the depth-averaged turbulent kinetic energy,10 4 ^k/u 2. (a) Case 1; (b) Case 2. The dashed lines visualize the boundaries of the MI in the mean flow. 14 of 21

15 Figure 14. Visualization of the 3 D vortical structure of the instantaneous flow for Case 1 using a Q isosurface (Q = 10). The two black solid lines indicate the approximate position of the MI. The red arrows point toward vertically oriented quasi-2d MI eddies. The dash-dot lines delimit the shear layer forming at the east bank due to the strong curvature of the inner bank and the separated shear layer at the west bank produced by flow moving over a submerged bar of failed bank material near Section A3 (see also Figure 2a). the SOV cells (Figure 13b), a finding consistent with field experiments [Rhoads and Sukhodolov, 2008]. Such a pattern is expected for confluences where the Kelvin-Helmholtz mode is strong. In contrast to Case 1 where large values of ^k were observed over the whole length of SVE1, the levels of ^k inside SVE1 in Case 2 are larger than those in the surrounding flow only within the region (situated approximately between Sections A2 and B) where the cell is subject to bimodal oscillations. Still, in a given cross-section the values of ^k in the region occupied by SVE1 are less than half those inside the MI. The levels of amplification of ^k inside SVE2 and SVW1 are negligible. Past Section B, the growth of instabilities propagating along the cores of the main SOV cells and vortex breakdown result in the formation of a region populated by highly 3-D energetic eddies whose orientation is variable in time (e.g., see Figure 15). These eddies interact and stretch the billows within the MI. 7. Implications of Large-Scale Coherent Structures for Bed Friction Velocity and Sediment Entrainment [49] The flows investigated in this study were not transport effective, at least for bed material. The extant bed morphology was produced by flows much larger than those for which the velocity data were obtained, but which had similar momentum ratios to the measured flows. Extensive field work at this confluence over the years has documented the influence of high momentum and low momentum ratio transport-effective events on bed morphology at the confluence [see Kenworthy and Rhoads, 1995; Rhoads, 1996; Rhoads et al., 2009]. The location of the scour hole in the center of the confluence for Case 1 (Mr 1) corresponds to morphologic conditions produced by transport effective flows with Mr 1, whereas the shift in position of the scour hole toward the outer (west) bank and the development of a pronounced bar along the inner (east bank) for Case 2 (Mr > 1) conforms to the pattern of bed morphology produced by events with Mr >1[Rhoads et al., 2009] (Figure 2). Thus, the changes morphology at the site between Cases 1 and 2 can be linked to changes in momentum ratio, or equivalently velocity ratio, and the Mr values of the low flows examined here correspond to Mr values of formative flows that produced the extant morphologic conditions. Moreover, field data for transport effective stages, when bed forms are present, indicate that flow structure for such events is characterized by pronounced SOVs on each side of a mixing interface [Rhoads, 1996] the same pattern documented at low flow stages and predicted by the simulations. As discussed below, simulations for the two test cases provide insight into predicted patterns of bed friction velocity for two different momentum ratios and suggest that mechanisms responsible for sediment entrainment within the confluence may differ for these two cases. [50] The most striking feature of the nondimensional distributions of the bed friction velocity is that for both cases the largest predicted values of u t /U occur beneath the SOV cells (Figure 16). While this result is not surprising for Case 1 as the MI is in wake mode and the highest predicted levels of turbulence are generated by SVE1, the results for Case 2 suggest that the SOV cells may be a primary mechanism for entraining sediment within the confluence hydrodynamic zone at high-angle asymmetrical confluences, regardless of the structure of the MI. In Case 2, however, the predicted instantaneous values of u t /U beneath the strongly coherent billows (Figure 15) are at many times comparable to those beneath the strongly coherent SOV cells, which is not the case for Case 1 [Constantinescu et al., 2011a]. The time of passage of a large Kelvin-Helmholtz billow over a certain location inside the MI is of the order of seconds downstream of Section A1. This duration is much longer than the minimum time interval required for sediment particles with d 84 < 1 cm to be entrained in a region where the bed friction velocity is larger than the threshold value for sediment entrainment [e.g., Celik et al., 2010]. Thus, if the Figure 15. Visualization of the 3D vortical structure of the instantaneous flow for Case 2 using a Q isosurface (Q = 10). The red arrows point toward vertically oriented quasi-2d eddies within the MI. The two black solid lines indicate the approximate position of the MI. Some of the neighboring MI eddies are in the process of merging together (e.g., see second and fifth red arrow). The blue and green ribbons visualize the paths of particles drawn into the core of SVE2. 15 of 21

16 Figure 16. Distribution of the magnitude of the bed friction velocity, u t /U, in the mean flow. (a) Case 1; (b) Case 2. The solid lines visualize the boundaries of the MI in the mean flow. The dash-dotted lines indicate the positions of the axes of the main SOV cells. The approximate position of the scour hole is visualized using two isobath (red dashed) lines. The two isobath lines correspond to levels situated 0.5D and 1D above point where the maximum scour depth is reached in Case 1 and to levels situated 0.35D and 0.7D above the point where the maximum scour depth is reached in Case 2. Figure is redrawn from Miyawaki et al. [2010]. Kelvin-Helmholtz mode is strong inside the MI, the quasi 2-D MI eddies may play a major role in sediment entrainment and transport. Comparison of the predicted distributions of u t /U in Figures 16a and 16b show that the size of the region characterized by high values of u t /U beneath the MI and thus the capacity of the mixing interface eddies to entrain sediment increases as the Kelvin-Helmholtz mode becomes dominant (e.g., for high values of the momentum ratio). [51] The predicted pattern of u t /U varies between the two cases in relation to the strength of the SOV cells. For a momentum ratio close to one, large u t /U are induced beneath the SOV cells on both sides of the MI, whereas for a high momentum ratio, large u t /U occur only beneath the SOV cells forming on the high-momentum side of the confluence (Figure 16). The streamwise locations where the strongly coherent SOV cells induce the largest u t /U values are centered on the region where the high speed cores of the two incoming streams collide (between Sections A2 and A1 for Case 1 and between Sections A1 and A for Case 2) and where, if present, the bimodal oscillations are most energetic. Despite the fact that the mean velocity in the Copper Slough is comparable in the two cases and the circulation of SVE1 is largest in Case 1, the peak values of u t /U beneath SVE1 are larger by 20 30% in Case 2 because the flow is much shallower and the cores of the SOV cells are situated closer to the bed than in Case 1. [52] Details of the bed morphology in the two cases account for other regions of high amplification of u t /U. For example, pronounced acceleration of the flow past the failed bar of bank material near the west bank explains the patch of high u t /U values upstream of Section A3 in Case 1. The patch of high u t /U values observed in Case 2 downstream of Section B within the region of greatest confluence scour is induced by high streamwise velocity fluid in the vicinity of the bed (Figure 17). This submerged core of high streamwise velocity fluid likely forms through redistribution of high, near-surface streamwise momentum by the strong SOV cells on the high momentum side of the MI. The overall pattern of bed friction velocity for Case 2 is generally consistent with the pattern of near-bed turbulent kinetic energy, which shifts from a narrow vertical band along the MI upstream to a horizontal pattern extending across the bed downstream as secondary flow advects streamwise momentum downward [Rhoads and Sukhodolov, 2008]. The downstream increase in the width of the zone of high bed friction velocity predicted by the numerical model is consistent with this pattern of the turbulent kinetic energy near the bed, which is related to bed shear stress [Biron et al., 2004]. [53] The spatial patterns of bed friction velocity for Cases 1 and 2 suggest that, if these patterns are sustained during transport-effective events, scour should extend close to the junction apex (Figure 16), whereas in both cases the zone of maximum scour is located downstream of the apex (Figure 2). Evaluation of the reason for this discrepancy requires velocity measurements during high-stage, channelshaping events, which are difficult to obtain at the KRCS confluence with ADVs due to dangerous flow conditions and the short duration (a few hours) of such events. Velocity measurements obtained at this confluence with an electromagnetic current meter during a transport-effective event with Mr > 1 indicate that the core of high velocity shifts toward the bed immediately downstream of the confluence as the strength of secondary circulation increases with increasing flow velocities and momentum is advected from the surface toward the bed [Rhoads, 1996]. More pronounced acceleration of flow through the confluence, enhanced helical motion, enlargement of the zone of flow stagnation, and a spatial lag in helical cell development and oscillation may Figure 17. Distribution of the mean streamwise velocity, u s /U, in Section C for Case 2. Mean streamwise flow is toward the viewer. The core of SVE2 is visualized using a Q isosurface (Q = 10). 16 of 21

Coherent structure dynamics and sediment erosion mechanisms around an in-stream rectangular cylinder at low and moderate angles of attack

Coherent structure dynamics and sediment erosion mechanisms around an in-stream rectangular cylinder at low and moderate angles of attack WATER RESOURCES RESEARCH, VOL. 47, W12532, doi:10.1029/2011wr010586, 2011 Coherent structure dynamics and sediment erosion mechanisms around an in-stream rectangular cylinder at low and moderate angles

More information

Flow and turbulence structure around an in stream rectangular cylinder with scour hole

Flow and turbulence structure around an in stream rectangular cylinder with scour hole WATER RESOURCES RESEARCH, VOL. 46,, doi:10.1029/2010wr009336, 2010 Flow and turbulence structure around an in stream rectangular cylinder with scour hole Gokhan Kirkil 1,2 and George Constantinescu 1 Received

More information

Periodic planes v i+1 Top wall u i. Inlet. U m y. Jet hole. Figure 2. Schematic of computational domain.

Periodic planes v i+1 Top wall u i. Inlet. U m y. Jet hole. Figure 2. Schematic of computational domain. Flow Characterization of Inclined Jet in Cross Flow for Thin Film Cooling via Large Eddy Simulation Naqavi, I.Z. 1, Savory, E. 2 and Martinuzzi, R. J. 3 1,2 The Univ. of Western Ontario, Dept. of Mech.

More information

2014 Quinn William Lewis

2014 Quinn William Lewis 2014 Quinn William Lewis RATES AND PATTERNS OF TEMPERATURE MIXING AT A SMALL STREAM CONFLUENCE UNDER VARIABLE INCOMING FLOW CONDITIONS BY QUINN WILLIAM LEWIS THESIS Submitted in partial fulfillment of

More information

B-1. Attachment B-1. Evaluation of AdH Model Simplifications in Conowingo Reservoir Sediment Transport Modeling

B-1. Attachment B-1. Evaluation of AdH Model Simplifications in Conowingo Reservoir Sediment Transport Modeling Attachment B-1 Evaluation of AdH Model Simplifications in Conowingo Reservoir Sediment Transport Modeling 1 October 2012 Lower Susquehanna River Watershed Assessment Evaluation of AdH Model Simplifications

More information

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD N. Jarrin 1, A. Revell 1, R. Prosser 1 and D. Laurence 1,2 1 School of MACE, the University of Manchester,

More information

5. Secondary Current and Spiral Flow

5. Secondary Current and Spiral Flow 5. Secondary Current and Spiral Flow The curve of constant velocity for rectangular and triangular cross-section obtained by Nikuradse are shown in Figures and 2. In all cases the velocities at the corners

More information

Module 3: Velocity Measurement Lecture 15: Processing velocity vectors. The Lecture Contains: Data Analysis from Velocity Vectors

Module 3: Velocity Measurement Lecture 15: Processing velocity vectors. The Lecture Contains: Data Analysis from Velocity Vectors The Lecture Contains: Data Analysis from Velocity Vectors Velocity Differentials Vorticity and Circulation RMS Velocity Drag Coefficient Streamlines Turbulent Kinetic Energy Budget file:///g /optical_measurement/lecture15/15_1.htm[5/7/2012

More information

Turbulent eddies in the RANS/LES transition region

Turbulent eddies in the RANS/LES transition region Turbulent eddies in the RANS/LES transition region Ugo Piomelli Senthil Radhakrishnan Giuseppe De Prisco University of Maryland College Park, MD, USA Research sponsored by the ONR and AFOSR Outline Motivation

More information

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders A. Jugal M. Panchal, B. A M Lakdawala 2 A. M. Tech student, Mechanical Engineering Department, Institute

More information

Validation 3. Laminar Flow Around a Circular Cylinder

Validation 3. Laminar Flow Around a Circular Cylinder Validation 3. Laminar Flow Around a Circular Cylinder 3.1 Introduction Steady and unsteady laminar flow behind a circular cylinder, representing flow around bluff bodies, has been subjected to numerous

More information

1. Introduction, tensors, kinematics

1. Introduction, tensors, kinematics 1. Introduction, tensors, kinematics Content: Introduction to fluids, Cartesian tensors, vector algebra using tensor notation, operators in tensor form, Eulerian and Lagrangian description of scalar and

More information

Turbulence Laboratory

Turbulence Laboratory Objective: CE 319F Elementary Mechanics of Fluids Department of Civil, Architectural and Environmental Engineering The University of Texas at Austin Turbulence Laboratory The objective of this laboratory

More information

Boundary-Layer Theory

Boundary-Layer Theory Hermann Schlichting Klaus Gersten Boundary-Layer Theory With contributions from Egon Krause and Herbert Oertel Jr. Translated by Katherine Mayes 8th Revised and Enlarged Edition With 287 Figures and 22

More information

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Turbulent Boundary Layers & Turbulence Models. Lecture 09 Turbulent Boundary Layers & Turbulence Models Lecture 09 The turbulent boundary layer In turbulent flow, the boundary layer is defined as the thin region on the surface of a body in which viscous effects

More information

7. Basics of Turbulent Flow Figure 1.

7. Basics of Turbulent Flow Figure 1. 1 7. Basics of Turbulent Flow Whether a flow is laminar or turbulent depends of the relative importance of fluid friction (viscosity) and flow inertia. The ratio of inertial to viscous forces is the Reynolds

More information

THE HYDRAULIC PERFORMANCE OF ORIENTED SPUR DIKE IMPLEMENTATION IN OPEN CHANNEL

THE HYDRAULIC PERFORMANCE OF ORIENTED SPUR DIKE IMPLEMENTATION IN OPEN CHANNEL Tenth International Water Technology Conference, IWTC10 2006, Alexandria, Egypt 281 THE HYDRAULIC PERFORMANCE OF ORIENTED SPUR DIKE IMPLEMENTATION IN OPEN CHANNEL Karima Attia 1 and Gamal El Saied 2 1

More information

Numerical Investigation of Vortex Induced Vibration of Two Cylinders in Side by Side Arrangement

Numerical Investigation of Vortex Induced Vibration of Two Cylinders in Side by Side Arrangement Numerical Investigation of Vortex Induced Vibration of Two Cylinders in Side by Side Arrangement Sourav Kumar Kar a, 1,, Harshit Mishra a, 2, Rishitosh Ranjan b, 3 Undergraduate Student a, Assitant Proffessor

More information

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing. Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing. Thus, it is very important to form both a conceptual understanding and a quantitative

More information

Detailed Outline, M E 521: Foundations of Fluid Mechanics I

Detailed Outline, M E 521: Foundations of Fluid Mechanics I Detailed Outline, M E 521: Foundations of Fluid Mechanics I I. Introduction and Review A. Notation 1. Vectors 2. Second-order tensors 3. Volume vs. velocity 4. Del operator B. Chapter 1: Review of Basic

More information

Visualization of flow pattern over or around immersed objects in open channel flow.

Visualization of flow pattern over or around immersed objects in open channel flow. EXPERIMENT SEVEN: FLOW VISUALIZATION AND ANALYSIS I OBJECTIVE OF THE EXPERIMENT: Visualization of flow pattern over or around immersed objects in open channel flow. II THEORY AND EQUATION: Open channel:

More information

NPTEL Quiz Hydraulics

NPTEL Quiz Hydraulics Introduction NPTEL Quiz Hydraulics 1. An ideal fluid is a. One which obeys Newton s law of viscosity b. Frictionless and incompressible c. Very viscous d. Frictionless and compressible 2. The unit of kinematic

More information

APPLICATION OF 3D NUMERICAL MODELS IN CONFLUENCE HYDRODYNAMICS MODELLING

APPLICATION OF 3D NUMERICAL MODELS IN CONFLUENCE HYDRODYNAMICS MODELLING XIX International Conference on Water Resources CMWR 2012 University of Illinois at Urbana-Champaign June 17-22, 2012 APPLICATION OF 3D NUMERICAL MODELS IN CONFLUENCE HYDRODYNAMICS MODELLING Dejana Đorđević

More information

Two-dimensional model problem to explain counter-rotating vortex pair formation in a transverse jet

Two-dimensional model problem to explain counter-rotating vortex pair formation in a transverse jet PHYSICS OF FLUIDS 18, 085103 2006 Two-dimensional model problem to explain counter-rotating vortex pair formation in a transverse jet Suman Muppidi and Krishnan Mahesh Aerospace Engineering & Mechanics,

More information

Simulation of Three-Dimensional Flow Field around Unconventional Bridge Piers

Simulation of Three-Dimensional Flow Field around Unconventional Bridge Piers Simulation of Three-Dimensional Flow Field around Unconventional Bridge Piers Adnan Ismael 1, Hamid Hussein 2, Mohammed Tareq 3, Mustafa gunal 4 1 Technical Institute/ Mosul-Iraq, 2 Technical College,

More information

Turbulent boundary layer

Turbulent boundary layer Turbulent boundary layer 0. Are they so different from laminar flows? 1. Three main effects of a solid wall 2. Statistical description: equations & results 3. Mean velocity field: classical asymptotic

More information

Fluid Dynamics Exercises and questions for the course

Fluid Dynamics Exercises and questions for the course Fluid Dynamics Exercises and questions for the course January 15, 2014 A two dimensional flow field characterised by the following velocity components in polar coordinates is called a free vortex: u r

More information

Effects of Free-Stream Vorticity on the Blasius Boundary Layer

Effects of Free-Stream Vorticity on the Blasius Boundary Layer 17 th Australasian Fluid Mechanics Conference Auckland, New Zealand 5-9 December 2010 Effects of Free-Stream Vorticity on the Boundary Layer D.A. Pook, J.H. Watmuff School of Aerospace, Mechanical & Manufacturing

More information

Vortex shedding from slender surface mounted pyramids

Vortex shedding from slender surface mounted pyramids Vortex shedding from slender surface mounted pyramids M. J. Morrison 1, R. J. Martinuzzi 3, E. Savory 1, G. A. Kopp 2 1 Department of Mechanical and Materials Engineering, University of Western Ontario,

More information

Numerical Methods in Aerodynamics. Turbulence Modeling. Lecture 5: Turbulence modeling

Numerical Methods in Aerodynamics. Turbulence Modeling. Lecture 5: Turbulence modeling Turbulence Modeling Niels N. Sørensen Professor MSO, Ph.D. Department of Civil Engineering, Alborg University & Wind Energy Department, Risø National Laboratory Technical University of Denmark 1 Outline

More information

Flow Patterns in an Open Channel Confluence with Increasingly Dominant Tributary Inflow

Flow Patterns in an Open Channel Confluence with Increasingly Dominant Tributary Inflow Water 2015, 7, 4724-4751; doi:10.3390/w7094724 Article OPEN ACCESS water ISSN 2073-4441 www.mdpi.com/journal/water Flow Patterns in an Open Channel Confluence with Increasingly Dominant Tributary Inflow

More information

Numerical study of the effects of trailing-edge bluntness on highly turbulent hydro-foil flows

Numerical study of the effects of trailing-edge bluntness on highly turbulent hydro-foil flows Numerical study of the effects of trailing-edge bluntness on highly turbulent hydro-foil flows T. Do L. Chen J. Tu B. Anderson 7 November 2005 Abstract Flow-induced noise from fully submerged lifting bodies

More information

Chapter 5 Phenomena of laminar-turbulent boundary layer transition (including free shear layers)

Chapter 5 Phenomena of laminar-turbulent boundary layer transition (including free shear layers) Chapter 5 Phenomena of laminar-turbulent boundary layer transition (including free shear layers) T-S Leu May. 3, 2018 Chapter 5: Phenomena of laminar-turbulent boundary layer transition (including free

More information

THE FLUVIAL DYNAMICS OF CONFLUENT MEANDER BENDS JAMES RILEY DISSERTATION

THE FLUVIAL DYNAMICS OF CONFLUENT MEANDER BENDS JAMES RILEY DISSERTATION THE FLUVIAL DYNAMICS OF CONFLUENT MEANDER BENDS BY JAMES RILEY DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Geography in the Graduate College

More information

Computation of Incompressible Flows: SIMPLE and related Algorithms

Computation of Incompressible Flows: SIMPLE and related Algorithms Computation of Incompressible Flows: SIMPLE and related Algorithms Milovan Perić CoMeT Continuum Mechanics Technologies GmbH milovan@continuummechanicstechnologies.de SIMPLE-Algorithm I - - - Consider

More information

Flow and Bed Topography in a 180 Curved Channel

Flow and Bed Topography in a 180 Curved Channel Flow and Bed Topography in a 180 Curved Channel Jae Wook Jung 1, Sei Eui Yoon 2 Abstract The characteristics of flow and bed topography has been analyzed by changing the bed materials in a 180-degree,

More information

UNIT II CONVECTION HEAT TRANSFER

UNIT II CONVECTION HEAT TRANSFER UNIT II CONVECTION HEAT TRANSFER Convection is the mode of heat transfer between a surface and a fluid moving over it. The energy transfer in convection is predominately due to the bulk motion of the fluid

More information

Quasi-three dimensional computations for flows and bed variations in curved channel with gently sloped outer bank

Quasi-three dimensional computations for flows and bed variations in curved channel with gently sloped outer bank River Sedimentation Wieprecht et al. (Eds) 2017 Taylor & Francis Group, London, ISBN 978-1-138-02945-3 Quasi-three dimensional computations for flows and bed variations in curved channel with gently sloped

More information

Direct Numerical Simulations of Transitional Flow in Turbomachinery

Direct Numerical Simulations of Transitional Flow in Turbomachinery Direct Numerical Simulations of Transitional Flow in Turbomachinery J.G. Wissink and W. Rodi Institute for Hydromechanics University of Karlsruhe Unsteady transitional flow over turbine blades Periodic

More information

Basic Fluid Mechanics

Basic Fluid Mechanics Basic Fluid Mechanics Chapter 6A: Internal Incompressible Viscous Flow 4/16/2018 C6A: Internal Incompressible Viscous Flow 1 6.1 Introduction For the present chapter we will limit our study to incompressible

More information

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION Joongnyon Kim, Kyoungyoun Kim, Hyung Jin Sung Department of Mechanical Engineering, Korea Advanced Institute of Science

More information

NUMERICAL INVESTIGATION OF THE FLOW OVER A GOLF BALL IN THE SUBCRITICAL AND SUPERCRITICAL REGIMES

NUMERICAL INVESTIGATION OF THE FLOW OVER A GOLF BALL IN THE SUBCRITICAL AND SUPERCRITICAL REGIMES NUMERICAL INVESTIGATION OF THE FLOW OVER A GOLF BALL IN THE SUBCRITICAL AND SUPERCRITICAL REGIMES Clinton Smith 1, Nikolaos Beratlis 2, Elias Balaras 2, Kyle Squires 1, and Masaya Tsunoda 3 ABSTRACT Direct

More information

Contents. I Introduction 1. Preface. xiii

Contents. I Introduction 1. Preface. xiii Contents Preface xiii I Introduction 1 1 Continuous matter 3 1.1 Molecules................................ 4 1.2 The continuum approximation.................... 6 1.3 Newtonian mechanics.........................

More information

Simulation of Flow around a Surface-mounted Square-section Cylinder of Aspect Ratio Four

Simulation of Flow around a Surface-mounted Square-section Cylinder of Aspect Ratio Four Simulation of Flow around a Surface-mounted Square-section Cylinder of Aspect Ratio Four You Qin Wang 1, Peter L. Jackson 2 and Jueyi Sui 2 1 High Performance Computing Laboratory, College of Science and

More information

On the influence of bed permeability on flow in the leeside of coarse-grained bedforms

On the influence of bed permeability on flow in the leeside of coarse-grained bedforms On the influence of bed permeability on flow in the leeside of coarse-grained bedforms G. Blois (1), J. L. Best (1), G. H. Sambrook Smith (2), R. J. Hardy (3) 1 University of Illinois, Urbana-Champaign,

More information

On the transient modelling of impinging jets heat transfer. A practical approach

On the transient modelling of impinging jets heat transfer. A practical approach Turbulence, Heat and Mass Transfer 7 2012 Begell House, Inc. On the transient modelling of impinging jets heat transfer. A practical approach M. Bovo 1,2 and L. Davidson 1 1 Dept. of Applied Mechanics,

More information

PROPERTIES OF THE FLOW AROUND TWO ROTATING CIRCULAR CYLINDERS IN SIDE-BY-SIDE ARRANGEMENT WITH DIFFERENT ROTATION TYPES

PROPERTIES OF THE FLOW AROUND TWO ROTATING CIRCULAR CYLINDERS IN SIDE-BY-SIDE ARRANGEMENT WITH DIFFERENT ROTATION TYPES THERMAL SCIENCE, Year, Vol. 8, No. 5, pp. 87-9 87 PROPERTIES OF THE FLOW AROUND TWO ROTATING CIRCULAR CYLINDERS IN SIDE-BY-SIDE ARRANGEMENT WITH DIFFERENT ROTATION TYPES by Cheng-Xu TU, a,b Fu-Bin BAO

More information

meters, we can re-arrange this expression to give

meters, we can re-arrange this expression to give Turbulence When the Reynolds number becomes sufficiently large, the non-linear term (u ) u in the momentum equation inevitably becomes comparable to other important terms and the flow becomes more complicated.

More information

EFFECT OF CHANNEL BENDS ON TRANSVERSE MIXING

EFFECT OF CHANNEL BENDS ON TRANSVERSE MIXING NIJOTECH VOL. 10. NO. 1 SEPTEMBER 1986 ENGMANN 57 EFFECT OF CHANNEL BENDS ON TRANSVERSE MIXING BY E. O. ENGMANN ABSTRACT Velocity and tracer concentration measurements made in a meandering channel are

More information

Active Control of Separated Cascade Flow

Active Control of Separated Cascade Flow Chapter 5 Active Control of Separated Cascade Flow In this chapter, the possibility of active control using a synthetic jet applied to an unconventional axial stator-rotor arrangement is investigated.

More information

Channel Pattern. Channel Pattern, Meanders, and Confluences. Description of Channel Pattern. Bridge (2003)

Channel Pattern. Channel Pattern, Meanders, and Confluences. Description of Channel Pattern. Bridge (2003) Channel Pattern Channel Pattern, Meanders, and Confluences Outline Description of channel pattern Alternate bars Channel pattern continua and evolution Controls of channel pattern Description of Channel

More information

DYNAMICS OF CONTROLLED BOUNDARY LAYER SEPARATION

DYNAMICS OF CONTROLLED BOUNDARY LAYER SEPARATION p.1 DYNAMICS OF CONTROLLED BOUNDARY LAYER SEPARATION Václav Uruba, Martin Knob Institute of Thermomechanics, AS CR, v. v. i., Praha Abstract: The results of experimental study on a boundary layer separation

More information

INTRODUCTION OBJECTIVES

INTRODUCTION OBJECTIVES INTRODUCTION The transport of particles in laminar and turbulent flows has numerous applications in engineering, biological and environmental systems. The deposition of aerosol particles in channels and

More information

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM 206 9th International Conference on Developments in esystems Engineering A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM Hayder Al-Jelawy, Stefan Kaczmarczyk

More information

Simulation Study on the Generation and Distortion Process of the Geomagnetic Field in Earth-like Conditions

Simulation Study on the Generation and Distortion Process of the Geomagnetic Field in Earth-like Conditions Chapter 1 Earth Science Simulation Study on the Generation and Distortion Process of the Geomagnetic Field in Earth-like Conditions Project Representative Yozo Hamano Authors Ataru Sakuraba Yusuke Oishi

More information

LES of turbulent shear flow and pressure driven flow on shallow continental shelves.

LES of turbulent shear flow and pressure driven flow on shallow continental shelves. LES of turbulent shear flow and pressure driven flow on shallow continental shelves. Guillaume Martinat,CCPO - Old Dominion University Chester Grosch, CCPO - Old Dominion University Ying Xu, Michigan State

More information

Introduction to Turbulence AEEM Why study turbulent flows?

Introduction to Turbulence AEEM Why study turbulent flows? Introduction to Turbulence AEEM 7063-003 Dr. Peter J. Disimile UC-FEST Department of Aerospace Engineering Peter.disimile@uc.edu Intro to Turbulence: C1A Why 1 Most flows encountered in engineering and

More information

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration 1 Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration A. Henze, C. Glatzer, M. Meinke, W. Schröder Institute of Aerodynamics, RWTH Aachen University, Germany March 21,

More information

International Conference on Energy Efficient Technologies For Automobiles (EETA 15) Journal of Chemical and Pharmaceutical Sciences ISSN:

International Conference on Energy Efficient Technologies For Automobiles (EETA 15) Journal of Chemical and Pharmaceutical Sciences ISSN: HEAT TRANSFER ENHANCEMENT WITH PRESSURE LOSS REDUCTION IN COMPACT HEAT EXCHANGERS USING VORTEX GENERATORS Viswajith M V*, Gireesh Kumaran Thampi, James Varghese Department of Mechanical Engineering, School

More information

Fluid Mechanics Prof. T.I. Eldho Department of Civil Engineering Indian Institute of Technology, Bombay. Lecture - 17 Laminar and Turbulent flows

Fluid Mechanics Prof. T.I. Eldho Department of Civil Engineering Indian Institute of Technology, Bombay. Lecture - 17 Laminar and Turbulent flows Fluid Mechanics Prof. T.I. Eldho Department of Civil Engineering Indian Institute of Technology, Bombay Lecture - 17 Laminar and Turbulent flows Welcome back to the video course on fluid mechanics. In

More information

Vortex Induced Vibrations

Vortex Induced Vibrations Vortex Induced Vibrations By: Abhiroop Jayanthi Indian Institute of Technology, Delhi Some Questions! What is VIV? What are the details of a steady approach flow past a stationary cylinder? How and why

More information

Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition

Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition C. Pozrikidis m Springer Contents Preface v 1 Introduction to Kinematics 1 1.1 Fluids and solids 1 1.2 Fluid parcels and flow

More information

Due Tuesday, November 23 nd, 12:00 midnight

Due Tuesday, November 23 nd, 12:00 midnight Due Tuesday, November 23 nd, 12:00 midnight This challenging but very rewarding homework is considering the finite element analysis of advection-diffusion and incompressible fluid flow problems. Problem

More information

Flow over ripples: KEY features ripple size independent of flow depth l ~ 1000d deceleration in leeside topographic acceleration over stoss flow

Flow over ripples: KEY features ripple size independent of flow depth l ~ 1000d deceleration in leeside topographic acceleration over stoss flow Ripples and dunes Flow over ripples: KEY features ripple size independent of flow depth l ~ 1000d deceleration in leeside topographic acceleration over stoss flow separation in leeside shear layer development

More information

A Novel Method for Prevention of Scouring Around Bridge Piers

A Novel Method for Prevention of Scouring Around Bridge Piers International Journal of Ecological Science and Environmental Engineering 2015; 2(2): 11-16 Published online April 10, 2015 (http://www.aascit.org/journal/ijesee) ISSN: 2375-3854 A Novel Method for Prevention

More information

Before we consider two canonical turbulent flows we need a general description of turbulence.

Before we consider two canonical turbulent flows we need a general description of turbulence. Chapter 2 Canonical Turbulent Flows Before we consider two canonical turbulent flows we need a general description of turbulence. 2.1 A Brief Introduction to Turbulence One way of looking at turbulent

More information

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman 1. Characteristics of the Roughness Sub layer With well understood caveats, the time averaged statistics of flow in the atmospheric

More information

Numerical Investigation of the Fluid Flow around and Past a Circular Cylinder by Ansys Simulation

Numerical Investigation of the Fluid Flow around and Past a Circular Cylinder by Ansys Simulation , pp.49-58 http://dx.doi.org/10.1457/ijast.016.9.06 Numerical Investigation of the Fluid Flow around and Past a Circular Cylinder by Ansys Simulation Mojtaba Daneshi Department of Mechanical Engineering,

More information

Principles of Convection

Principles of Convection Principles of Convection Point Conduction & convection are similar both require the presence of a material medium. But convection requires the presence of fluid motion. Heat transfer through the: Solid

More information

Towards the prediction of free-forming meander formation using 3D computational fluid dynamics

Towards the prediction of free-forming meander formation using 3D computational fluid dynamics Wasserbaukolloquium 2006: Strömungssimulation im Wasserbau 31 Dresdner Wasserbauliche Mitteilungen Heft 32 Towards the prediction of free-forming meander formation using 3D computational fluid dynamics

More information

White Paper FINAL REPORT AN EVALUATION OF THE HYDRODYNAMICS MECHANISMS WHICH DRIVE THE PERFORMANCE OF THE WESTFALL STATIC MIXER.

White Paper FINAL REPORT AN EVALUATION OF THE HYDRODYNAMICS MECHANISMS WHICH DRIVE THE PERFORMANCE OF THE WESTFALL STATIC MIXER. White Paper FINAL REPORT AN EVALUATION OF THE HYDRODYNAMICS MECHANISMS WHICH DRIVE THE PERFORMANCE OF THE WESTFALL STATIC MIXER Prepared by: Dr. Thomas J. Gieseke NUWCDIVNPT - Code 8233 March 29, 1999

More information

CFD Study of Flow Over Parallel Ridges with Varying Height and Spacing

CFD Study of Flow Over Parallel Ridges with Varying Height and Spacing Proceedings of the World Congress on Engineering 21 Vol II WCE 21, June 3 - July 2, 21, London, U.K. CFD Study of Flow Over Parallel Ridges with Varying Height and Spacing Lee Chin Yik, Salim Mohamed Salim,

More information

Explicit algebraic Reynolds stress models for internal flows

Explicit algebraic Reynolds stress models for internal flows 5. Double Circular Arc (DCA) cascade blade flow, problem statement The second test case deals with a DCA compressor cascade, which is considered a severe challenge for the CFD codes, due to the presence

More information

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh Fluid Mechanics Chapter 9 Surface Resistance Dr. Amer Khalil Ababneh Wind tunnel used for testing flow over models. Introduction Resistances exerted by surfaces are a result of viscous stresses which create

More information

Chapter 5. Experimental Results - Dynamics

Chapter 5. Experimental Results - Dynamics Chapter 5 Experimental Results - Dynamics 16 5.1 Overview The dynamics of the mean flow fields described in Chapter 4 will be presented in this chapter using mainly two tools: the power spectral density

More information

EXCITATION OF GÖRTLER-INSTABILITY MODES IN CONCAVE-WALL BOUNDARY LAYER BY LONGITUDINAL FREESTREAM VORTICES

EXCITATION OF GÖRTLER-INSTABILITY MODES IN CONCAVE-WALL BOUNDARY LAYER BY LONGITUDINAL FREESTREAM VORTICES ICMAR 2014 EXCITATION OF GÖRTLER-INSTABILITY MODES IN CONCAVE-WALL BOUNDARY LAYER BY LONGITUDINAL FREESTREAM VORTICES Introduction A.V. Ivanov, Y.S. Kachanov, D.A. Mischenko Khristianovich Institute of

More information

ROLE OF THE VERTICAL PRESSURE GRADIENT IN WAVE BOUNDARY LAYERS

ROLE OF THE VERTICAL PRESSURE GRADIENT IN WAVE BOUNDARY LAYERS ROLE OF THE VERTICAL PRESSURE GRADIENT IN WAVE BOUNDARY LAYERS Karsten Lindegård Jensen 1, B. Mutlu Sumer 1, Giovanna Vittori 2 and Paolo Blondeaux 2 The pressure field in an oscillatory boundary layer

More information

BLUFF-BODY AERODYNAMICS

BLUFF-BODY AERODYNAMICS International Advanced School on WIND-EXCITED AND AEROELASTIC VIBRATIONS OF STRUCTURES Genoa, Italy, June 12-16, 2000 BLUFF-BODY AERODYNAMICS Lecture Notes by Guido Buresti Department of Aerospace Engineering

More information

Department of Mechanical Engineering

Department of Mechanical Engineering Department of Mechanical Engineering AMEE401 / AUTO400 Aerodynamics Instructor: Marios M. Fyrillas Email: eng.fm@fit.ac.cy HOMEWORK ASSIGNMENT #2 QUESTION 1 Clearly there are two mechanisms responsible

More information

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION 7.1 THE NAVIER-STOKES EQUATIONS Under the assumption of a Newtonian stress-rate-of-strain constitutive equation and a linear, thermally conductive medium,

More information

LES ANALYSIS ON CYLINDER CASCADE FLOW BASED ON ENERGY RATIO COEFFICIENT

LES ANALYSIS ON CYLINDER CASCADE FLOW BASED ON ENERGY RATIO COEFFICIENT 2th International Conference on Heat Transfer, Fluid Mechanics and Thermodynamics ANALYSIS ON CYLINDER CASCADE FLOW BASED ON ENERGY RATIO COEFFICIENT Wang T.*, Gao S.F., Liu Y.W., Lu Z.H. and Hu H.P. *Author

More information

Module 2: External Flows Lecture 12: Flow Over Curved Surfaces. The Lecture Contains: Description of Flow past a Circular Cylinder

Module 2: External Flows Lecture 12: Flow Over Curved Surfaces. The Lecture Contains: Description of Flow past a Circular Cylinder The Lecture Contains: Description of Flow past a Circular Cylinder Experimental Results for Circular Cylinder Flow file:///d /Web%20Course%20(Ganesh%20Rana)/Dr.%20gautam%20biswas/Final/convective_heat_and_mass_transfer/lecture12/12_1.htm[12/24/2014

More information

Physical Properties of Fluids

Physical Properties of Fluids Physical Properties of Fluids Viscosity: Resistance to relative motion between adjacent layers of fluid. Dynamic Viscosity:generally represented as µ. A flat plate moved slowly with a velocity V parallel

More information

Numerical Studies of Supersonic Jet Impingement on a Flat Plate

Numerical Studies of Supersonic Jet Impingement on a Flat Plate Numerical Studies of Supersonic Jet Impingement on a Flat Plate Overset Grid Symposium Dayton, OH Michael R. Brown Principal Engineer, Kratos/Digital Fusion Solutions Inc., Huntsville, AL. October 18,

More information

FLUID MECHANICS. Chapter 9 Flow over Immersed Bodies

FLUID MECHANICS. Chapter 9 Flow over Immersed Bodies FLUID MECHANICS Chapter 9 Flow over Immersed Bodies CHAP 9. FLOW OVER IMMERSED BODIES CONTENTS 9.1 General External Flow Characteristics 9.3 Drag 9.4 Lift 9.1 General External Flow Characteristics 9.1.1

More information

2.3 The Turbulent Flat Plate Boundary Layer

2.3 The Turbulent Flat Plate Boundary Layer Canonical Turbulent Flows 19 2.3 The Turbulent Flat Plate Boundary Layer The turbulent flat plate boundary layer (BL) is a particular case of the general class of flows known as boundary layer flows. The

More information

Chuichi Arakawa Graduate School of Interdisciplinary Information Studies, the University of Tokyo. Chuichi Arakawa

Chuichi Arakawa Graduate School of Interdisciplinary Information Studies, the University of Tokyo. Chuichi Arakawa Direct Numerical Simulations of Fundamental Turbulent Flows with the Largest Grid Numbers in the World and its Application of Modeling for Engineering Turbulent Flows Project Representative Chuichi Arakawa

More information

Self-Excited Vibration in Hydraulic Ball Check Valve

Self-Excited Vibration in Hydraulic Ball Check Valve Self-Excited Vibration in Hydraulic Ball Check Valve L. Grinis, V. Haslavsky, U. Tzadka Abstract This paper describes an experimental, theoretical model and numerical study of concentrated vortex flow

More information

NUMERICAL SIMULATION OF THE FLOW AROUND A SQUARE CYLINDER USING THE VORTEX METHOD

NUMERICAL SIMULATION OF THE FLOW AROUND A SQUARE CYLINDER USING THE VORTEX METHOD NUMERICAL SIMULATION OF THE FLOW AROUND A SQUARE CYLINDER USING THE VORTEX METHOD V. G. Guedes a, G. C. R. Bodstein b, and M. H. Hirata c a Centro de Pesquisas de Energia Elétrica Departamento de Tecnologias

More information

Calculation of the separation streamlines of barchans and transverse dunes

Calculation of the separation streamlines of barchans and transverse dunes Calculation of the separation streamlines of barchans and transverse dunes H. J. Herrmann a,b, J. S. Andrade Jr. b, V. Schatz a, G. Sauermann a and E. J. R. Parteli a a ICP, University of Stuttgart, Pfaffenwaldring

More information

Preliminary Study of the Turbulence Structure in Supersonic Boundary Layers using DNS Data

Preliminary Study of the Turbulence Structure in Supersonic Boundary Layers using DNS Data 35th AIAA Fluid Dynamics Conference, June 6 9, 2005/Toronto,Canada Preliminary Study of the Turbulence Structure in Supersonic Boundary Layers using DNS Data Ellen M. Taylor, M. Pino Martín and Alexander

More information

Dual Vortex Structure Shedding from Low Aspect Ratio, Surface-mounted Pyramids

Dual Vortex Structure Shedding from Low Aspect Ratio, Surface-mounted Pyramids Dual Vortex Structure Shedding from Low Aspect Ratio, Surface-mounted Pyramids Robert J. Martinuzzi Department of Mechanical and Manufacturing Engineering Schulich School of Engineering University of Calgary

More information

CFD STUDY OF MASS TRANSFER IN SPACER FILLED MEMBRANE MODULE

CFD STUDY OF MASS TRANSFER IN SPACER FILLED MEMBRANE MODULE GANIT J. Bangladesh Math. Soc. (ISSN 1606-3694) 31 (2011) 33-41 CFD STUDY OF MASS TRANSFER IN SPACER FILLED MEMBRANE MODULE Sharmina Hussain Department of Mathematics and Natural Science BRAC University,

More information

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson, lada

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson,   lada ublication 97/ An ntroduction to Turbulence Models Lars Davidson http://www.tfd.chalmers.se/ lada Department of Thermo and Fluid Dynamics CHALMERS UNVERSTY OF TECHNOLOGY Göteborg Sweden November 3 Nomenclature

More information

Numerical simulation of three-dimensional, time-averaged flow

Numerical simulation of three-dimensional, time-averaged flow WATER RESOURCES RESEARCH, VOL. 36, NO. 9, PAGES 2731-2746, SEPTEMBER 2000 Numerical simulation of three-dimensional, time-averaged flow structure at river channel confluences K. F. Bradbrook Mort McDonald

More information

SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS

SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS ICAS 2000 CONGRESS SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS H Yao, R K Cooper, and S Raghunathan School of Aeronautical Engineering The Queen s University of Belfast, Belfast BT7 1NN,

More information

HIGH RESOLUTION NUMERICAL SIMULATIONS OF LOCK- EXCHANGE GRAVITY-DRIVEN FLOWS

HIGH RESOLUTION NUMERICAL SIMULATIONS OF LOCK- EXCHANGE GRAVITY-DRIVEN FLOWS HIGH RESOLUTION NUMERICAL SIMULATIONS OF LOCK- EXCHANGE GRAVITY-DRIVEN FLOWS by Seng Keat Ooi, George Constantinescu, and Larry J. Weber IIHR Technical Report No. 450 IIHR Hydroscience & Engineering College

More information

MATHEMATICAL MODELING OF FLUVIAL SEDIMENT DELIVERY, NEKA RIVER, IRAN. S.E. Kermani H. Golmaee M.Z. Ahmadi

MATHEMATICAL MODELING OF FLUVIAL SEDIMENT DELIVERY, NEKA RIVER, IRAN. S.E. Kermani H. Golmaee M.Z. Ahmadi JOURNAL OF ENVIRONMENTAL HYDROLOGY The Electronic Journal of the International Association for Environmental Hydrology On the World Wide Web at http://www.hydroweb.com VOLUME 16 2008 MATHEMATICAL MODELING

More information

THE EFFECT OF THICKNESS OF PILLAR IN THE CHANNEL BEND TO CHANGES THE COEFFICIENT OF SUPERELEVATION

THE EFFECT OF THICKNESS OF PILLAR IN THE CHANNEL BEND TO CHANGES THE COEFFICIENT OF SUPERELEVATION Journal Engineering Science and Technology Vol. 11, No. 5 (2016) 745-754 School Engineering, Taylor s University THE EFFECT OF THICKNESS OF PILLAR IN THE CHANNEL BEND TO CHANGES THE COEFFICIENT OF SUPERELEVATION

More information

Vortex wake and energy transitions of an oscillating cylinder at low Reynolds number

Vortex wake and energy transitions of an oscillating cylinder at low Reynolds number ANZIAM J. 46 (E) ppc181 C195, 2005 C181 Vortex wake and energy transitions of an oscillating cylinder at low Reynolds number B. Stewart J. Leontini K. Hourigan M. C. Thompson (Received 25 October 2004,

More information