Large-eddy simulations (LES) of a single-phase, turbulent flow in a 90° pipe bend are performed at three Reynolds numbers (5300, 27 000, and 45 000) to investigate the correlation between secondary flow motion and wall shear stresses, which is suspected to be a potential mechanism responsible for material erosion. The isothermal flows are validated against available experimental and numerical data first. The snapshot proper orthogonal decomposition (POD) is applied for the medium and high Reynolds number flows to identify the secondary flow motions and the oscillation of the Dean vortices that are found to cause swirl-switching. Distinguished frequencies of the POD time coefficients at Strouhal numbers of 0.25 and 0.28 are identified for Reynolds numbers at 27 000 and 45 000, respectively. Moreover, shear stress on the pipe wall and the associated power spectral density are obtained and shown to have the same oscillating frequency as the swirl-switching.

Supercritical carbon dioxide (sCO2) is an ideal working fluid for power generation in concentrated solar, fossil fuel, and nuclear power plants as it is a nontoxic, nonflammable, and low-cost fluid. The high-pressure (200–350 bar), high-temperature (550–750°C), and high-density fluid enables extremely compact and high efficiency turbomachinery designs. However, severe erosion has been observed under several occasions in the sCO2 power cycle test loops, particularly in the turbine nozzles and blades. Since the pipelines usually bear erosion and corrosion simultaneously in multiphase conditions, it was conjectured that erosion-corrosion (E-C) type of behavior, i.e., small oxide particles and impurities in sCO2, caused the erosion and failure of the blades. Recently, Finn and Doğan1 performed large-eddy simulations of turbulent flows over a turbine nozzle in real operating conditions and modeled solid particle collision on the turbine blades using one-way coupling. Based on the simulation results and the Finnie model,2 the erosion magnitudes on the turbine blade surface were estimated. It was found that the trailing edge erosion is roughly two orders of magnitude larger than that on the leading edge, consistent with the observations reported by Fleming et al.,3 suggesting that mechanical erosion, caused possibly by the small solid oxide particles spalled from the internal walls of the pipe, could be a source of severe erosion in the power generation systems.

However, erosion of the material was also observed even with a fairly pure sCO2 flowing through the power cycles.4 It is, thus, critical to identify the potential sources of erosion in this pure flow scenario with minimal particulates. The wall thinning has long been an important safety issue for complex pipeline systems including elements such as elbows and orifices,5 and the so-called flow accelerated corrosion (FAC) is the major cause for this phenomenon.6 FAC usually happens at the oxide layer of the carbon steel wall where the mass transfer of iron ions into the turbulent bulk flow occurs through a diffusive process,7,8 which could further promote the E-C type of erosion on a carbon steel wall. A wall thinning issue caused by the flow turbulence has been observed particularly in pipe bends and orifices.9–11 Hence, it is reasonable to speculate that FAC first contributes to the generation of small iron particulates at complex geometries (bend or orifice) in the pure sCO2 power cycles by thinning the carbon steel wall; then, the particles entrained in the flow are responsible for the erosion of the turbine parts. This speculation aligns well with the observation reported by Fleming and Kruizenga4 that high iron particles were detected in various areas on the flow path, and the numerical study of the mechanical erosion on turbine blades reported by Finn and Doğan.1 The current work focuses on the first half of the conjuncture to explore the potential cause of the erosion in relatively pure sCO2 systems.

It is noteworthy to mention the physical underlying of the mass transfer associated with FAC has not been fully understood due to complex geometry and the secondary flows generated by the curved pipe.5 Limited studies were published to investigate such phenomenon. Recently, Takano et al.12 experimentally studied how the swirling flow could impact the mass and momentum transfer of a pipe with elbow and orifice. It was found that the nonaxisymmetric flow structure is the major mechanism responsible for the increased mass transfer downstream of the pipe. Utanohara and Murase13 conducted experimental work on the influence of velocity and temperature on FAC in the pipe bend by measuring the FAC rate at various velocity and temperature settings. The results indicate that the combined effects are case dependent and particularly strong at the elbow. Some researchers argued the FAC may be attributed to large fluctuations of wall shear stress.11 Lister and Lang14 proposed that the shear stress is directly leading to the mechanical shearing of the outer oxide layer.

In the current work, it is hypothesized that the substantial shear stress on the pipe walls and its spatiotemporal variation due to turbulence and secondary flow patterns are the major causes of erosion in the pure sCO2 power systems. To test the hypothesis and to gain better understanding of the stress distributions on the inner walls of pipe bends, predictive numerical simulation of a single-phase, turbulent flows without heat transfer in pipe bends at representative conditions and flow Reynolds numbers are performed. Considerable work exploring the dynamics of laminar and turbulent flow through pipe bends has been conducted.15–17 Flow in curved pipes with bends involves flow turning that results in centrifugal forces on the fluid elements and corresponding pressure field to balance these forces. Formation of counter-rotating vortices, commonly termed as Dean vortices (Fig. 1), are observed as the fluid elements with higher velocity are forced to the outside of the bend and lower velocity elements are forced toward the inside.15,16 These curvature-induced swirling motions and their strength is described in terms of the Dean number, De, where De=D/2RcReD,ReD=DUb/ν is the Reynolds number, D is the diameter of the pipe, Rc is the radius of curvature of the bend measured to the pipe centerline, Ub is the bulk velocity, and ν is the kinematic viscosity of the fluid. Under unsteady flow conditions, the Dean vortices also exhibit transient behavior. The two counter-rotating swirling motions start to be dominant alternately over space and time. This unsteady behavior of Dean vortices is called swirl-switching.18 The transient behavior of swirl-switching, time scales related to oscillations of these vortices, and resultant variations in the shear stresses on the walls in turbulent flow have not been thoroughly investigated, especially at large Reynolds numbers.

FIG. 1.

Dean vortices on the cross section plane of a pipe after the bend. The mean flow is going out of the paper.

FIG. 1.

Dean vortices on the cross section plane of a pipe after the bend. The mean flow is going out of the paper.

Close modal

Rütten et al.16 investigated turbulent flows in a 90° pipe bend using large-eddy simulations at ReD of 5300–27 000. Their work indicated that the overall forces on the pipe walls showed a distinct peak in the power spectra related to the vortex shedding at the inner side of the bend. At large ReD, the power spectra also exhibited a frequency of oscillation much lower than the vortex shedding from flow separation. This was attributed to two Dean vortices whose strength varies in time dominating the flow field. This also suggests that local variation in the wall shear forces, especially after the bend, can be important in magnitude as well as time scales. Such variations in local shear stresses may be relevant to the erosion problem at very large ReD.

Furthermore, Hellström et al.17 used time-resolved stereoscopic particle image velocimetry (PIV) method to study turbulent flows at multiple ReD from 25 000 to ReD=115000 through a 90° bend. Snapshot proper orthogonal decomposition (POD) method was applied to identify the Dean motion and swirl-switching, as well as extracting the corresponding POD modes. The time coefficient of the first POD mode was found to be oscillating at the frequency, which is close to that of the overall force reported by Rütten et al.16 Although ReD of the two cases are not exactly the same, this leads to an open question as to whether the swirl-switching can cause the oscillation of the shear forces on the pipe downstream of the bend and be responsible for the material erosion observed in the power cycles.

More recently, Carlsson et al.19 performed large-eddy simulations of turbulent flows at ReD=3.4×104 through pipe bends with four different curvature ratios γ=0.16,0.25,0.35,and0.5, which is defined as γ=Rc/D. Two distinct types of swirl-switching were observed at high (Strouhal number St=0.50.6) and low (St0.01) frequencies, respectively. The Strouhal number is a nondimensional frequency and defined as St=fD/Ub. The authors argued that the high frequency switching roots from the bend pipe; while the low frequency signal results from the upstream turbulent pipe flows. However, Hufnagel et al.20 disagrees with this conclusion, where direct numerical simulations of flow at ReD = 11 700 in a bend pipe with γ=0.3 were conducted. Three dimensional POD analyses were applied and the traveling waves were identified to be the cause of the swirl-switching rather than the straight pipe section upstream of the pipe bend. One important distinction of the simulation settings between Carlsson et al.19 and Hufnagel et al.20 is that the recycling inflow boundary was used in the former study, while a synthetic eddy method was employed by the latter. Hufnagel et al.20 claims that the artificial frequency introduced by the recycling inflow boundary attributes to the connection between swirl-switching and upstream pipe flow established by Carlsson et al.19 Besides, Ikarashi et al.21 measured the velocity fields of very high ReD flows (1×1053×105) through a pipe with an short elbow (γ=1.0) using PIV. They inferred that the periodic velocity fluctuations are generated by the oscillation of flow structures downstream of the pipe bend. However, no shear stress/force data were reported in any of these recent publications.

In the present work, we simulate the turbulent flows at three different ReD of 5300, 27 000, and 45 000 through a 90° bend to first validate our results against those available data from experiments as well as other simulations and then investigate the nature of shear force variations at different locations after the pipe bend. In addition, the snapshot POD approach is applied to detect the dominant POD modes and any frequency associated with the swirl-switching. The power spectral density functions are computed for both shear force and POD modes to reveal the connections between them. The rest of the paper is arranged as follows. Section II describes the mathematical formulations and numerical scheme used to solve the governing equations. Section III briefly goes through the simulation setups including geometry and inlet boundary condition as well as grid resolutions. The results are presented and discussed in Sec. IV. Section V states the summary and conclusions of the study.

The governing equations, the dynamic Smagorinsky model, and the numerical solver used in this work are briefly described below.

In the present work, the standard mathematical formulations for the single-phase, incomprehensible large-eddy simulations (LES) are employed as

uj̃xj=0,
(1)
uĩt+uĩuj̃xj=p¯ρxi+xj(2ν¯Sij̃)qijrxj,
(2)

where

Sij̃=12(uĩuj+uj̃ui)13δijũkxk,
(3)

and ũi is the filtered velocity, μ¯ is the viscosity, and p¯ is the pressure. The subgrid scale unclosed transport terms in the momentum equation are grouped into the residual stress qijr. The dynamic Smagorinsky model by Mahesh et al.,22 Moin and Apte23 is used. The dynamic model has no adjustable constants, and thus allows evaluation of the predictive capability of the model. The subgrid scale stress, qijr is modeled as

qijr=(ũiũjuiuj̃)=2νtS̃ij13q2δij,
(4)

where the isotropic part of the subgrid stress q2 is absorbed into the pressure term, and the subgrid viscosity νt is given as

νt=CνΔ¯2Sij̃Sij̃.
(5)

Here, Δ¯ is the filter width and the coefficient Cμ is obtained using least squares approach following the dynamic procedure.24 

An energy-conserving, finite-volume scheme for unstructured, arbitrary shaped grid elements is used to solve the fluid-flow equations using a fractional step algorithm.22,23,25 The velocity and pressure are stored at the centroids of the volumes. The cell-centered velocities are advanced in a predictor step such that the kinetic energy is conserved. The predicted velocities are interpolated to the faces and then projected. Projection yields the pressure potential at the cell-centers, and its gradient is used to correct the cell and face-normal velocities. A novel discretization scheme for the pressure gradient was developed by Mahesh et al.22 to provide robustness without explicit numerical dissipation on grids with rapidly varying elements. This algorithm was found to be imperative to perform LES at high Reynolds number in complex flows. The overall algorithm is second-order accurate in space and time for uniform orthogonal grids. A numerical solver based on this approach was developed and shown to give very good results for both simple26 and complex geometries23 and is used in the present study.27,28 A thorough verification and validation of the algorithm was conducted29,30 to assess the accuracy of the numerical scheme for test cases involving two-dimensional decaying Taylor vortices, flow through a turbulent channel and duct flows,29 and particle-laden turbulent flows.23,26

The computational geometry and the boundary conditions are provided in detail in this section. Three different ReD flows were simulated at 5300, 27 000, and 45 000. Among them, the lowest ReD flow (5300) mostly serves as a validation case, whose results were compared with the published numerical and experimental data set. The results from the rest of the cases with higher ReD were analyzed thoroughly to investigate the characteristics of such flows. A simulation in a periodic pipe was performed at each ReD to generate proper inlet flow data for the pipe bend. The grid resolutions are also listed and discussed.

The computational setup used consists of a circular cross section pipe with a 90° bend as shown in Fig. 2. The pipe consists of a straight section of length 8D, followed by the bend section and another straight section of 8D length. The bend with the curvature ratio of 1 is investigated. Owing to the bend geometry an unstructured grid with predominantly hexagonal mesh is used for calculations. The so-called O-grid mesh is generated using a commercial software ANSYS ICEM-CFD.

FIG. 2.

Schematic diagram of the computational setup for the 90° pipe bend with curvature ratio Rc/D=1.

FIG. 2.

Schematic diagram of the computational setup for the 90° pipe bend with curvature ratio Rc/D=1.

Close modal

When conducting large-eddy simulations, it is important to impose consistent, continuity-satisfying fluctuations in the velocity field at the inlet31 to create proper turbulence structure and velocity fluctuations in downstream. The use of random fluctuations scaled with turbulence intensity values has been proposed. However, this approach can lead to convergence issues in a conservative numerical solver. As a result, a separate periodic pipe flow (length 5D as used in Rütten et al.,16 see Fig. 2) is simulated at the desired mass-flow rate and ReD using a body-force technique.31 The turbulent flow field in this cycling pipe was verified by comparing the first-and second-order turbulent statistics with those from straight pipe flow at the same ReD. Then, the instantaneous velocity field at the outlet face of the periodic pipe is stored at every time step and is used as the inlet flow condition for the pipe bend. This provides a fully developed, instantaneous, turbulent velocity profile at the inlet of the pipe bend. A convective outlet condition is enforced at the outlet. To avoid the possible artificial frequency associated with this process,20 the time span of the inflow data were chosen to be long enough (>60 flow-through times for the flow in the pipe bend). In this study, only an isothermal flow field is simulated and thus effects of temperature and density variations and properties of sCO2 are not included. According to Utanohara and Murase,13 both velocity and temperature have strong influence on the erosion rate, and the combined effects are rather complicated than a simple linear combination from the two aspects. Later study involves turbulent flow with heat transfer where these effects do become important.

The flow fields simulated are for three different ReD, 5300 and 27 000 similar to the work by Rütten et al.16 and 45 000 which is one of the experimental cases conducted by Hellström et al.17 The grid resolutions for the three cases are listed in Table I. Stretched grids are generated based on the hyperbolic tangent function. The grids in the wall normal (radial) direction have a minimum grid resolution near the wall in wall units of Δr+0.8 for ReD = 5300, less than 2 for ReD = 27 000 and 45 000. Nearly uniform grids are used in the axial direction with a resolution of Δz+20 for both ReD = 5300 and 27 000, while above 30 for ReD = 45 000. The maximum resolution in the circumferential direction is on the order of (rΔθ)+6 for ReD = 5300, around 20 for ReD = 27 000 and 31 for ReD = 45 000. Grid resolutions in the azimuthal direction used in this work for ReD=5000and27000 are finer than those used for LES by Rütten et al.16 (14.1 for ReD = 5300 and 41.9 for ReD = 27 000); ReD = 45 000 is comparable to the LES case (ReD = 37 600) by Chin et al.32 In the present wall-resolved large-eddy simulation, no-slip conditions are applied at walls.

TABLE I.

Grid resolutions in wall units for three different ReD. Δr represents the grid resolution in radial direction, Δrθ azimuthal direction and Δz the axial direction. The superscript + indicates they are in wall unit.

Δrmin+(Δrθ)max+Δzmin/max+
ReD = 5300 0.8 6.0 20/25 
ReD = 27 000 1.13 21.4 20/25 
ReD = 45 000 1.77 31.5 30/38 
Δrmin+(Δrθ)max+Δzmin/max+
ReD = 5300 0.8 6.0 20/25 
ReD = 27 000 1.13 21.4 20/25 
ReD = 45 000 1.77 31.5 30/38 

In this section, the simulated results are presented and discussed. The mean and fluctuating velocity profiles at multiple locations of both upstream and downstream of the pipe bend are shown and compared with available reference data. Then, snapshot POD method is applied to analyze the flows at ReD=27000and45000. Three most energetic modes at three xy-planes (z/D=1,3,and5) are visualized. The power spectral densities of the associated time coefficients are computed. Besides, the shear forces on the pipe wall were analyzed in the frequency domain to identify any distinguished frequencies, which are found to be identical to those for the swirl-switching phenomenon.

To validate the large-eddy simulation results in the present study, the inflow data obtained in the straight pipe section are analyzed first. The RMS velocity components at ReD=5300and27000 are compared with data from multiple references.16,33,34 Both the mean and fluctuation velocity profiles on multiple cross sections of the pipe bend are shown to demonstrate the effect from pipe bend to the upstream flow. Also, the velocity profiles at one diameter (1D) downstream of the pipe bend for ReD = 5300 are compared with data from both experimental35 and numerical16 studies. Finally, the pressure distribution on pipe wall and pipe center is reported and set side by side with the reference data.

Figure 3 provides the RMS velocity profiles in the straight pipe, which was employed to generate inflow data, together with data from multiple reference, where ReD = 5300 is illustrated in Fig. 3(a) and ReD = 27 000 in (b). The LES results by Rütten et al.16 and DNS results by Unger33 are given as well for comparison in Fig. 3(a). The data from two LES studies16,34 are illustrated in Fig. 3(b). All three components of the RMS velocity are matching well with the reference data, suggesting the good credibility of the current simulation settings. In addition, the mean velocity and RMS velocity profiles of ReD = 5300 at various upstream locations, e.g., −0D, −1D, and −2D as illustrated in Fig. 2, are plotted in Fig. 4. Here, “inner side” means the profile is plotted along the inner side line of the pipe, and likewise the “outer side” and “center” [see Fig. 4(g)]. It is observed that the mean velocity profiles before 1D of the bend all collapse together, and coincide with that of a fully developed straight pipe. This is consistent with what has been observed in Vester et al.36 that the flow is not influenced by the bend before one diameter of the pipe. The same characteristic can be obtained from RMS velocity profiles as well. The velocity profiles in Fig. 4 are in good agreement with those in Rütten et al.,16 where the same flow was simulated, though they are not included here.

FIG. 3.

RMS velocity profiles of the straight pipe flow for velocity components in axial (uax+), circumferential (uθ+), and radial (ur+) directions at (a) ReD = 5300 and (b) ReD = 27 000. The reference data are represented using symbols: LES results by Rütten et al.,16 ○ DNS data by Unger,33 and ▽ LES data (channel) by Schröder et al.34 

FIG. 3.

RMS velocity profiles of the straight pipe flow for velocity components in axial (uax+), circumferential (uθ+), and radial (ur+) directions at (a) ReD = 5300 and (b) ReD = 27 000. The reference data are represented using symbols: LES results by Rütten et al.,16 ○ DNS data by Unger,33 and ▽ LES data (channel) by Schröder et al.34 

Close modal
FIG. 4.

Mean axial velocity profiles nondimensionalized by bulk velocity for ReD = 5300 along the (a) inner side, (c) outer side, and (e) central in between of the pipe; and RMS velocity profiles in wall units along the outer side for velocity components in (b) axial, (d) radial, and (f) circumferential directions. The inner (red), central (blue), and outer (green) sides are illustrated in (g).

FIG. 4.

Mean axial velocity profiles nondimensionalized by bulk velocity for ReD = 5300 along the (a) inner side, (c) outer side, and (e) central in between of the pipe; and RMS velocity profiles in wall units along the outer side for velocity components in (b) axial, (d) radial, and (f) circumferential directions. The inner (red), central (blue), and outer (green) sides are illustrated in (g).

Close modal

The velocity profiles along the centerline at 1D downstream of the bend are shown in Fig. 5. The LES data are compared with PIV data from Brücker35 and the LES results by Rütten et al.16 It is observed that the mean velocity profile matches well with both experimental and numerical data. The maximum relative difference was estimated to be 7%. The profile of the axial component of RMS velocity in Fig. 5(b) also shows a good comparison with experimental and published numerical data. The RMS velocity near the pipe walls agrees well with the LES study by Rütten et al.16 

FIG. 5.

(a) Mean velocity profile normalized by bulk velocity along the central line at 1D downstream the bend of ReD = 5300 and (b) RMS velocity profile. ––––– present results, □ experimental data by Brücker,35 LES data by Rütten et al.16 

FIG. 5.

(a) Mean velocity profile normalized by bulk velocity along the central line at 1D downstream the bend of ReD = 5300 and (b) RMS velocity profile. ––––– present results, □ experimental data by Brücker,35 LES data by Rütten et al.16 

Close modal

Besides, the pressure distribution along the pipe wall and pipe center at ReD = 5300 is investigated and plotted in Fig. 6(a) together with the LES data from Rütten et al.37 The abscissa represents the nondimensional distance along the wall starting at 3 diameter before the bend. The ordinate describes the nondimensional pressure drop computed by Δpn=Δp/(ρ2U¯b2). The results again match well with the reference data.

FIG. 6.

(a) Nondimensional pressure drop along the inner, center and outer pipe wall at ReD = 5300. ▷ represents the LES data by Rütten et al.37 The two vertical solid lines indicate the curved pipe section. (b) Illustration of the inner wall (red), pipe center (blue) and outer wall (green).

FIG. 6.

(a) Nondimensional pressure drop along the inner, center and outer pipe wall at ReD = 5300. ▷ represents the LES data by Rütten et al.37 The two vertical solid lines indicate the curved pipe section. (b) Illustration of the inner wall (red), pipe center (blue) and outer wall (green).

Close modal

The flow field is visualized to provide an overall understanding of the flow features. Figures 7(a) and 7(b) show the contours of the mean velocity magnitude field nondimensionalized by the bulk velocity (|Um|=ux¯2+uy¯2+uz¯2Ub) on the symmetry plane at ReD=27000and45000, respectively. No significant difference of the mean velocity magnitude was observed between the two ReD flows. The high velocity region from the straight pipe first gets closer to the inner wall starting about 1D before the bend. Then after about 30° of curvature angle, it deviates from the inner wall, moves toward the outer wall and keeps this way downstream. This flow feature is consistent with what was reported by Sudo et al.38 

FIG. 7.

Contours of the nondimensional mean velocity magnitude field for (a) ReD = 27 000 and (b) ReD = 45 000. The contours of the nondimensional in-plane velocity magnitude overlaid with stream lines at z/D=1 are displayed on the second row (c and d); z/D=3 the third (e and f), where the left-hand side is for ReD = 27 000 and right-hand side ReD = 45 000.

FIG. 7.

Contours of the nondimensional mean velocity magnitude field for (a) ReD = 27 000 and (b) ReD = 45 000. The contours of the nondimensional in-plane velocity magnitude overlaid with stream lines at z/D=1 are displayed on the second row (c and d); z/D=3 the third (e and f), where the left-hand side is for ReD = 27 000 and right-hand side ReD = 45 000.

Close modal

Besides, the nondimensional in-plane mean velocity magnitude field (|Uxy|=ux¯2+uy¯2Ub) overlaid with streamlines is depicted at the cross sections of pipe at z/D=1and3 [see the dashed line in Fig. 7(a)]. Figures 7(c) and 7(d) are for z/D=1; Figs. 7(e) and 7(f)z/D=3, where the left-hand side is for ReD = 27 000 and right-hand side ReD = 45 000. Symmetric distribution of the in-plane mean velocity was observed as expected. Similar to the observation from mean velocity magnitude field, limited difference was spotted for the in-plane mean velocity distributions between the two ReD flows. A stagnation point is present at the centers of both inner (right-hand side in the figure) and outer (left-hand side) sides of the pipe. Two counter-rotating vortices are generated due to the centrifugal force, with a secondary vortex existing solely at z/D=1. The high velocity regions are mostly concentrated near the outer wall, where large magnitudes of shear forces are guaranteed. Since the velocity fields are time-averaged, the unsteady swirl-switching is not detected in these plots and a more detailed temporal analysis is needed.

In order to identify the swirl-switching in the turbulent flows through the pipe bends, POD is applied to extract the flow features. POD is capable of capturing the optimal modes from any field being examined bases on the energy content. In other words, the optimal basis vectors are computed that can best represent the given data. The snapshot POD method can greatly reduce the computational effort compared to the original POD approach. Hence, this method has been widely used in the fluid dynamics to detect the coherent structures and flow patterns.17,39 In this work, the mathematical formulations for the snapshot POD method described in Taira et al.40 are adopted and briefly discussed in the following. For any given vector field q(x,t) with the temporal mean q¯(x) removed, the remaining fluctuation part can be decomposed as

q(x,t)q¯(x)=jaj(t)ϕj(x),
(6)

where ϕj represents the POD modes, and aj the corresponding time coefficients. The POD mode is defined as a function of location only; while the temporal coefficient depends on time. In this way, the unsteady components of the vector field q(x,t) can be decomposed into temporal and spatial domains, respectively.

In this work, the velocity fields at three different locations: z/D=1,3,and5 after the bend are investigated using the snapshot POD method. The snapshots of the velocity fields at these locations are stored in a specific frequency that is much higher than the target frequency of the POD modes to avoid aliasing. Table II provides the detailed time interval and span for generating the data sets used to compute the POD modes. Figure 8 shows the relative and integrated energy levels for the first 20 modes at three cross sections for both high ReD flows. It shows that the first three modes are the most energetic ones, which takes over 50% of the total energy and are illustrated in Figs. 9 and 10. The modes are distinguished by black solid lines and are overlaid by the normalized streamwise velocity field that is pseudo-colored.

TABLE II.

Time interval and span for the POD analyses. δt is the time step for simulation, δtPOD is time interval for dumping data at specific locations and nPOD represents the total number of the data sets generated for POD analyses.

δtδtPODnPOD
ReD = 27 000 0.0035 0.875 4000 
ReD = 45 000 0.0022 0.44 6000 
δtδtPODnPOD
ReD = 27 000 0.0035 0.875 4000 
ReD = 45 000 0.0022 0.44 6000 
FIG. 8.

The relative and integrated energy levels for the first 20 modes. Columns from left to right are corresponding to the planes at z/D=1,3,&5. The first row (a, b, and c) is for ReD = 27 000 and second (d, e, and f) ReD = 45 000.

FIG. 8.

The relative and integrated energy levels for the first 20 modes. Columns from left to right are corresponding to the planes at z/D=1,3,&5. The first row (a, b, and c) is for ReD = 27 000 and second (d, e, and f) ReD = 45 000.

Close modal
FIG. 9.

The first three POD modes (modes 1–3 from left to right) pseudo-colored by streamwise velocity (uz/Ub) for ReD = 27 000 at z/D=1,3,&5. The modes at the first row are at 1D downstream of the bend (a, d, and g), second 3D (b, e, and h) and third 5D (c, f, and i).

FIG. 9.

The first three POD modes (modes 1–3 from left to right) pseudo-colored by streamwise velocity (uz/Ub) for ReD = 27 000 at z/D=1,3,&5. The modes at the first row are at 1D downstream of the bend (a, d, and g), second 3D (b, e, and h) and third 5D (c, f, and i).

Close modal
FIG. 10.

The first three POD modes (modes 1–3 from left to right) pseudo-colored by the normalized streamwise velocity (uz/Ub) for ReD = 45 000 at z/D=1,3,&5. The modes at the first row are at 1D downstream of the bend (a, d, and g), second 3D (b, e, and h) and third 5D (c, f, and i).

FIG. 10.

The first three POD modes (modes 1–3 from left to right) pseudo-colored by the normalized streamwise velocity (uz/Ub) for ReD = 45 000 at z/D=1,3,&5. The modes at the first row are at 1D downstream of the bend (a, d, and g), second 3D (b, e, and h) and third 5D (c, f, and i).

Close modal

The first mode is a single cell covering the whole pipe cross section regardless of mode number and ReD. This mode, consistent with what is reported by Hellström et al.17 for flows at ReD=25000and115000, represents the swirl-switching phenomenon. Since the modes extracted by POD are orthogonal, the single cell mode observed can represent the two counter-rotating vortices which share the same basis, as pointed out by Hellström et al.17 

The Dean-like vortex is recovered in the second and third modes, where two vortices can be identified, indicating that the Dean vortices have less energy and are weaker than the swirl-switching. In fact, the Dean-like motions could be seen as the transient period of the swirl-switching. It is noticed that the second mode at z/D=1 is not a typical Dean motion compared to that at z/D=3and5 for both ReD=27000and45000. This may be attributed to the fact that z/D=1 is too close to the bend. The Dean motion is still under development at this specific location. Another interesting observation is, the second and third modes are nearly perpendicular to each other at both z/D=3and5, demonstrating the bimodal characteristic of such type of flows. This feature was not present at z/D=1, possibly due to the incompleteness of the Dean motion development. Hellström et al.17 did not report such abnormal behavior because all of the data reported were at or after 5D of the bend. In addition, no substantial difference in the POD modes' structures was observed between ReD = 27 000 and ReD = 45 000.

Moreover, the power spectral density functions for the time coefficients aj are plotted in Figs. 11(a)–11(f), representing the temporal characteristics of the POD modes. The first row is for PSD at xy-plane of z/D=1, second z/D=3, and third z/D=5. The figures in the left column are for ReD = 27 000, right column ReD = 45 000. Aiming to reduce the noise and focus on the meaningful frequencies, the PSD of the time coefficients are filtered using Bartlett window with the degree of freedom of 60. The 95% confidence interval is included in each figure as reference. The abscissa of the figure is the Strouhal number. A very well pronounced peak at St =0.25 is observed for ReD = 27 000, and it shifts to St =0.28 at ReD = 45 000. Those distinguished peaks suggest that the motion of swirl-switching is oscillating at such frequencies. The sole dependency of the peaking frequency on ReD indicates that this oscillation happens over the whole pipe section after the bend, thus is not spatially dependent.

FIG. 11.

Power spectral density functions for the time coefficient of POD modes at the cross section of z/D=1 are in the first row (a and b), z/D=3 second (c and d) and z/D=5 third (e and f). The left column is for ReD = 27 000 and right ReD = 45 000.

FIG. 11.

Power spectral density functions for the time coefficient of POD modes at the cross section of z/D=1 are in the first row (a and b), z/D=3 second (c and d) and z/D=5 third (e and f). The left column is for ReD = 27 000 and right ReD = 45 000.

Close modal

The PSD for mode 2 has a low frequency peak at St0.01 at all locations regardless of ReD. Carlsson et al.19 claimed the root of this low frequency peak is the very-large-scale motions (VLSMs) created in the upstream pipe flow; while Hufnagel et al.20 argued it results from the artificial signal created by the recycling inflow data. As mode 2 exhibits the low frequency peak disregarding the location and ReD, the latter explanation seems plausible. However, with a closer look, the low frequency signal in mode 3 gradually becomes more intensified over the distance away from the bend. In addition, such tendency is more clearly observed at higher ReD flow (ReD = 45 000), where mode 3 has almost the same magnitude of low frequency peak as that for mode 2 at z/D=5. Such observation suggests the peaking low frequency signal may be related to both location and ReD at least for mode 3, which contradicts with the arguments from Hufnagel et al.20 but aligns with the conclusion by Carlsson et al.19 Last but not least, due to the very long time span for the inflow data used in the present work (>60 flow-through time), the low frequency is very unlikely to come from the recycling inlet condition.

One significant difference of the PSD between the two ReD cases is that modes 1 and 2 for ReD = 45 000 share the same peak at all three xy-planes; while the overlapping of the peak frequency is only heavily pronounced at z/D=5 for ReD = 27 000. Such different behavior has not been reported by other research, to the best of the authors' knowledge. One possible reason is that, the length of pipe for flow to achieve the fully developed Dean motions is depending on ReD. This is yet to be confirmed by further studies with even higher ReD flows.

The shear forces caused by the fluid motions are obtained everywhere on the pipe surface. The overall forces are analyzed in the frequency domain to identify any oscillation. The shear forces have been nondimensionalized by the stagnation pressure of the bulk velocity 12ρUb2 and the squared pipe diameter D2 following Rütten et al.16 It is observed in Figs. 12(a) and 12(b) that the PSD has the peak frequency at St =0.25 for ReD = 27 000 and St =0.28 for ReD = 45 000. These two frequencies are exactly matching with those of the POD modes. This is a strong evidence that the swirl-switching is highly correlated with the oscillation of the total shear force. This suggests that the turbulent flow induces oscillations of wall shear forces which can be a cause of the erosion, especially if the force magnitudes are large. For real sCO2 power plant, the operating conditions can lead to extreme high ReD (20 million), and hence much higher shear forces. These oscillating shear forces can remove material from the internal surface of pipe such as oxide scale formed as a result of a high-temperature reaction between CO2 and steel surface.41 In many instances, oxide layer growing on the steel surface is loosely adhered to the surface and prone to spallation under external forces such as oscillating shear forces demonstrated above or even due to internal growth stresses.42 Once the spalled oxide particles become entrained in the CO2 flow, they can accelerate erosion through hard particle impact on the internal surfaces of downstream components.1 In addition, changes in temperatures and thermal properties of the fluid can influence the forces and their spectra as well.13 

FIG. 12.

PSD of the overall shear force in X direction of (a) ReD = 27 000 and (b) ReD = 45 000.

FIG. 12.

PSD of the overall shear force in X direction of (a) ReD = 27 000 and (b) ReD = 45 000.

Close modal

Large-eddy simulations of turbulent flows through a 90° pipe bend with unit radius of curvatures (Rc/D=1) were performed using an unstructured grid solver. Flows at three different ReD (5300, 27 000, and 45 000) were simulated to understand the behavior of the flow field within the pipe bends, the dynamics of the Dean vortices, and variations in the shear stress. First, the solver was validated by comparing the mean and RMS velocity profiles before and after the bend with available experimental and numerical data. Then, the snapshot POD method was applied to detect the optimal modes and identify the swirl-switching motion. The associated time coefficients were analyzed in the frequency domain to obtain its oscillating frequency. In addition, the shear forces on the pipe were computed and the power spectral density function was calculated as well. It was found that the swirl-switching has the same frequency as that of the oscillation of the shear stress. This represents strong evidence that the swirl-switching could possibly be the cause of the erosion by contributing the oscillating shear forces on the pipe.

This work was performed in support of the U.S. Department of Energy's Fossil Energy Crosscutting Technology Research and Advanced Turbine Programs. The research was executed through NETL Research and Innovation Center's Advanced Alloy Development Field Work Proposal. This research was supported in part by an appointment (X.H.) to the NETL Research Participation Program sponsored by the U.S. Department of Energy and administered by the Oak Ridge Institute for Science and Education. The use of the NETL's supercomputer Joule is also acknowledged. This research was partially supported by the U.S. Department of Energy, Office of Science, Office of Biological and Environmental Research, Environmental System Science (ESS) Program. This contribution originates from the River Corridor Scientific Focus Area (SFA) project at Pacific Northwest National Laboratory (PNNL).

This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
J. R.
Finn
and
Ö. N.
Doğan
, “
Analyzing the potential for erosion in a supercritical CO2 turbine nozzle with large eddy simulation
,” in
Turbo Expo: Power for Land, Sea, and Air
(
American Society of Mechanical Engineers
,
2019
), Vol.
58585
, p.
V02DT47A019
.
2.
I.
Finnie
, “
Erosion of surfaces by solid particles
,”
Wear
3
,
87
103
(
1960
).
3.
D.
Fleming
,
A.
Kruizenga
,
J.
Pasch
,
T.
Conboy
, and
M.
Carlson
, “
Corrosion and erosion behavior in supercritical CO2 power cycles
,” in
ASME Turbo Expo 2014: Turbine Technical Conference and Exposition
(
American Society of Mechanical Engineers
,
2014
), p.
V03BT36A002
.
4.
D.
Fleming
and
A.
Kruizenga
, “
Identified corrosion and erosion mechanisms in sCO2 Brayton cycles
,” Report No. SAND2014-15546 (
Sandia National Laboratories
,
2014
).
5.
S.
Taguchi
,
Y.
Ikarashi
,
T.
Yamagata
,
N.
Fujisawa
, and
F.
Inada
, “
Mass and momentum transfer characteristics in 90° elbow under high Reynolds number
,”
Int. Commun. Heat Mass Transfer
90
,
103
110
(
2018
).
6.
L.
Zeng
,
G.
Chen
, and
H.
Chen
, “
Comparative study on flow-accelerated corrosion and erosion–corrosion at a 90° carbon steel bend
,”
Materials
13
,
1780
(
2020
).
7.
L. E.
Sanchez-Caldera
,
P.
Griffith
, and
E.
Rabinowicz
, “
The mechanism of corrosion–erosion in steam extraction lines of power stations
,”
J. Eng. Gas Turbines Power
110
,
180
184
(
1988
).
8.
K. M.
HWANG
,
T. E.
JIN
, and
K. H.
KIM
, “
Identification of the relationship between local velocity components and local wall thinning inside carbon steel piping
,”
J. Nucl. Sci. Technol.
46
,
469
478
(
2009
).
9.
W. H.
Ahmed
, “
Evaluation of the proximity effect on flow-accelerated corrosion
,”
Ann. Nucl. Energy
37
,
598
605
(
2010
).
10.
M.
El-Gammal
,
H.
Mazhar
,
J.
Cotton
,
C.
Shefski
,
J.
Pietralik
, and
C.
Ching
, “
The hydrodynamic effects of single-phase flow on flow accelerated corrosion in a 90-degree elbow
,”
Nucl. Eng. Des.
240
,
1589
1598
(
2010
).
11.
J. M.
Pietralik
and
C. S.
Schefski
, “
Flow and mass transfer in bends under flow-accelerated corrosion wall thinning conditions
,”
J. Eng. Gas Turbines Power
133
,
012902
(
2011
).
12.
T.
Takano
,
Y.
Ikarashi
,
K.
Uchiyama
,
T.
Yamagata
, and
N.
Fujisawa
, “
Influence of swirling flow on mass and momentum transfer downstream of a pipe with elbow and orifice
,”
Int. J. Heat Mass Transfer
92
,
394
402
(
2016
).
13.
Y.
Utanohara
and
M.
Murase
, “
Influence of flow velocity and temperature on flow accelerated corrosion rate at an elbow pipe
,”
Nucl. Eng. Des.
342
,
20
28
(
2019
).
14.
D. H.
Lister
and
L. C.
Lang
, “
A mechanistic model for predicting flow-assisted and general corrosion of carbon steel in reactor primary coolants
,” Proc. Int. Conf. Water Chem. Nucl. Reactor Systems. Avignon, France, 2002 (September
2002
), available at https://scholar.google.com/scholar_lookup?title=A%20Mechanistic%20Model%20for%20Predicting%20Flow-assisted%20and%20General%20Corrosion%20of%20Carbon%20Steel%20in%20Reactor%20Primary%20Coolants&author=DH.%20Lister&author=L.%20Lang&publication_year=2002.
15.
S.
Berger
,
L.
Talbot
, and
L.
Yao
, “
Flow in curved pipes
,”
Annu. Rev. Fluid Mech.
15
,
461
512
(
1983
).
16.
F.
Rütten
,
W.
Schröder
, and
M.
Meinke
, “
Large-eddy simulation of low frequency oscillations of the Dean vortices in turbulent pipe bend flows
,”
Phys. Fluids
17
,
035107
(
2005
).
17.
L. H.
Hellström
,
M. B.
Zlatinov
,
G.
Cao
, and
A. J.
Smits
, “
Turbulent pipe flow downstream of a bend
,”
J. Fluid Mech.
735
,
R7
(
2013
).
18.
M. J.
Tunstall
and
J. K.
Harvey
, “
On the effect of a sharp bend in a fully developed turbulent pipe-flow
,”
J. Fluid Mech.
34
,
595
608
(
1968
).
19.
C.
Carlsson
,
E.
Alenius
, and
L.
Fuchs
, “
Swirl switching in turbulent flow through 90°pipe bends
,”
Phys. Fluids
27
,
085112
(
2015
).
20.
L.
Hufnagel
,
J.
Canton
,
R.
Örlü
,
O.
Marin
,
E.
Merzari
, and
P.
Schlatter
, “
The three-dimensional structure of swirl-switching in bent pipe flow
,”
J. Fluid Mech.
835
,
86
101
(
2018
).
21.
Y.
Ikarashi
,
T.
Yamagata
,
F.
Yamagishi
, and
N.
Fujisawa
, “
Unsteady turbulence structure in and downstream of a short elbow at post-critical Reynolds numbers
,”
Nucl. Eng. Des.
364
,
110649
(
2020
).
22.
K.
Mahesh
,
G.
Constantinescu
,
S.
Apte
,
G.
Iaccarino
,
F.
Ham
, and
P.
Moin
, “
Large-eddy simulation of reacting turbulent flows in complex geometries
,”
J. Appl. Mech.
73
,
374
(
2006
).
23.
P.
Moin
and
S.
Apte
, “
Large-eddy simulation of realistic gas turbine-combustors
,”
AIAA J.
44
,
698
708
(
2006
).
24.
M.
Germano
,
U.
Piomelli
,
P.
Moin
, and
W.
Cabot
, “
A dynamic subgrid-scale eddy viscosity model
,”
Phys. Fluids A: Fluid Dyn.
3
,
1760
(
1991
).
25.
K.
Mahesh
,
G.
Constantinescu
, and
P.
Moin
, “
A numerical method for large-eddy simulation in complex geometries
,”
J. Comput. Phys.
197
,
215
240
(
2004
).
26.
S.
Apte
,
K.
Mahesh
,
P.
Moin
, and
J.
Oefelein
, “
Large-eddy simulation of swirling particle-laden flows in a coaxial-jet combustor
,”
Int. J. Multiphase Flow
29
,
1311
1331
(
2003
).
27.
S.
Apte
,
O.
Dogan
, and
C.
Ghodke
, “
Numerical investigation of potential erosion mechanisms in turbulent flow of sCO2 pipe bends
,” in
5th International Symposium—Supercritical CO₂ Power Cycles
(
2016
).
28.
X.
He
,
S. V.
Apte
, and
O. N.
Dogan
, “
Numerical simulations of supercritical CO2 flow through pipe bends: Identification of a potential cause of material erosion
,” in
6th International sCO₂ Power Cycles Symposium
(
2018
).
29.
E.
Shams
,
J.
Finn
, and
S.
Apte
, “
A numerical scheme for Euler-Lagrange simulation of bubbly flows in complex systems
,”
Int. J. Numer. Methods Fluids
67
,
1865
1898
(
2011
).
30.
E.
Shams
and
S. V.
Apte
, “
Prediction of small-scale cavitation in a high speed flow over an open cavity using large-eddy simulation
,”
J. Fluids Eng.
132
,
111301
(
2010
).
31.
C.
Pierce
and
P.
Moin
, “
Large eddy simulation of a confined coaxial jet with swirl and heat release
,”
AIAA Paper No. 2892
(
1998
).
32.
C.
Chin
,
H.
Ng
,
H.
Blackburn
,
J.
Monty
, and
A.
Ooi
, “
Turbulent pipe flow at re 1000: A comparison of wall-resolved large-eddy simulation, direct numerical simulation and hot-wire experiment
,”
Comput. Fluids
122
,
26
33
(
2015
).
33.
F.
Unger
, “
Numerische Simulation Turbulenter Rohrströmungen
,” Ph.D. thesis (
Lehrstuhl für Fluidmechanik, Technische Universität München
,
1994
).
34.
W.
Schröder
,
M.
Meinke
, and
A.
Schvorak
, “
Large-eddy simulations of accelerated pipe flow
,”
Comput. Fluid Dyn. J.
(submitted).
35.
C.
Brücker
, “
A time-recording DPIV-study of the swirl switching effect in a 90° bend flow
,” in
Proceedings of the Eight International Symposium on Flow Visualization
,
Sorrento, Italy
(
1998
).
36.
A. K.
Vester
,
R.
Örlü
, and
P. H.
Alfredsson
, “
Turbulent flows in curved pipes: Recent advances in experiments and simulations
,”
Appl. Mech. Rev.
68
,
050802
(
2016
).
37.
F.
Rütten
,
M.
Meinke
, and
W.
Schröder
, “
Large-eddy simulations of 90°pipe bend flows
,”
J. Turbul.
2
,
N3
(
2001
).
38.
K.
Sudo
,
M.
Sumida
, and
H.
Hibara
, “
Experimental investigation on turbulent flow through a circular-sectioned 180 degrees bend
,”
Exp. Fluids
28
,
51
57
(
2000
).
39.
A. K.
Vester
,
R.
Örlü
, and
P. H.
Alfredsson
, “
POD analysis of the turbulent flow downstream a mild and sharp bend
,”
Exp. Fluids
56
,
15
(
2015
).
40.
K.
Taira
,
S. L.
Brunton
,
S. T. M.
Dawson
,
C. W.
Rowley
,
T.
Colonius
,
B. J.
McKeon
,
O. T.
Schmidt
,
S.
Gordeyev
,
V.
Theofilis
, and
L. S.
Ukeiley
, “
Modal analysis of fluid flows: An overview
,”
AIAA J.
55
,
4013
4041
(
2017
).
41.
G. R.
Holcomb
,
C.
Carney
, and
N.
Doğan
, “
Oxidation of alloys for energy applications in supercritical CO2 and H2O
,”
Corros. Sci.
109
,
22
35
(
2016
).
42.
M.
Schütze
,
P.
Tortorelli
, and
I.
Wright
, “
Development of a comprehensive oxide scale failure diagram
,”
Oxid. Met.
73
,
389
418
(
2010
).