Sie sind auf Seite 1von 11

Nuclear Engineering and Design 239 (2009) 888898

Contents lists available at ScienceDirect

Nuclear Engineering and Design


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

Non-intrusive experimental investigation of ow behavior inside a 5 5 rod bundle with spacer grids using PIV and MIR
Elvis E. Dominguez-Ontiveros, Yassin A. Hassan
Department of Nuclear Engineering, Texas A&M University, College Station, TX 77843, USA

a r t i c l e

i n f o

a b s t r a c t
The validity of the simulation results from computational uid dynamics (CFD) is still under scrutiny. Some existing CFD closure models for complex ow produce results that are generally recognized as being inaccurate. Development of improved models for complex ow simulation requires an improved understanding of the detailed ow structure evolution with dynamic interaction of the ow multi-scales. Thus, the goal of this work is to contribute to a better understanding of presupposed and existent events that could affect the safety of nuclear power plants. The fundamental phenomena of uid ow in rod bundles with spacer grids can be elucidated by using state-of-the-art measurement techniques. This study aims to develop an experimental data base with high spatial and temporal resolution of uid ow velocity inside a 5 5 rod bundles with spacer grids. The full-eld detailed data base is intended to validate CFD codes at various temporal-spatial scales. Measurements are carried out using dynamic particle image velocimetry (DPIV) technique inside an optically transparent rod bundle utilizing the matching index of refraction (MIR) approach. This work presents full eld velocity vectors and turbulence statistics for a rod bundle under single phase ow conditions. 2009 Elsevier B.V. All rights reserved.

Article history: Received 30 November 2008 Received in revised form 25 January 2009 Accepted 26 January 2009

1. Introduction Fluid ow around circular cylinders is one of the classical problems of uid mechanics and has been well studied because of its common occurrence in many forms and in different applications. Cylinder-like structures can be found both alone and in groups in the designs for heat exchangers, cooling systems for nuclear power plants, offshore structures, power lines, struts, grids, screens, and cables, in both single and multi-phase ows. A complete understanding of the uid dynamics for the ow around a circular cylinder includes such fundamental subjects as the boundary layer, separation, the free shear layer, the wake, and the dynamics of vortices. The ow eld of multiple-cylinder congurations involves complex interactions between the shear layers, vortices and Karman vortex streets (Zdravkovich,1987). The problem is further complicated by the large number of congurations encountered in practice, resulting in different ow patterns, and by the effect of their interactions. One of the applications of paramount importance in this study is uid ow in fuel rod bundles of water nuclear reactors. In this type of nuclear reactor, optimum heat removal from the surface of fuel elements is the subject of many studies for researchers in order to determine reactor thermal margin and safety. In these reactors, the

Corresponding author. E-mail address: Hassan@ne.tamu.edu (Y.A. Hassan). 0029-5493/$ see front matter 2009 Elsevier B.V. All rights reserved. doi:10.1016/j.nucengdes.2009.01.009

spacer grids which support the fuel assembly are used as an effective mixing device by attaching various types of ow deectors. Several recent works are focused on the development of numerical simulations that predict the complex behavior of uid ow close to grid spacers and between fuel assemblies. The validity of the produced results from computational uid dynamics (CFD) is still under scrutiny for several applications in real practical cases. Moreover, the existing models for multiphase ows produce results that are generally recognized as unreliable. Development of better models for multiphase simulation requires an improved understanding of the ow evolution with dynamic interaction of the ow multi-scales. Therefore, experimental data is urgently needed for validation of the CFD models. In order to establish reliable design and performance criteria for tube bundle models, better velocity data is needed. Simonin and Barcouda (1988) conducted experiments using laser Doppler anemometry (LDA) within a specic tube arrangement. Although they gathered some velocity data in cross-ow over a tube bundle, the data points were limited and detailed velocity distributions, or whole ow eld data was not available due to the nature of the LDA technique. Chang et al. (2008) used 2D-LDA measurements in a 5 5 rod bundle array scaled to be 2.6 times larger than the actual bundle size. This work focused on the performance and mixing characteristics of two kinds of spacers with turbulence enhancement vanes. The used spacer grids were of a typical split and swirl type for pressurized water reactors (PWR). The experiments were performed at a condition of Reynolds number of Re = 48,000 and pressure of 1.5 bar.

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

889

Fig. 1. Schematic diagram of the experimental hydraulic loop and test section.

They reported the velocity at various axial positions ranging from one hydraulic diameter to 16 hydraulic diameters. The reported accuracy of the velocity measurements is about 1.6%. However, the use of laser Doppler velocimetry (LDV) methods was restricted by the existence of invisible regions in fuel rod bundles and insufcient spatial resolution for the narrow gaps in rod bundles. Ikeda and Hoshi (2006) developed a miniaturized LDV system embedded in a fuel cladding. The rod-embedded ber LDV can be inserted in an arbitrary grid cell instead of a fuel rod without disturbing the ow. They obtained ow velocity data in a 5 5 rod bundle typical of a PWR fuel assembly. Measurements were carried out for a central rod with a pitch-to-diameter ratio P/D = 1.326 with a hydraulic diameter, Dh = 9.75 mm. The Re number used was 57,000. The reported uncertainty in velocity data was 2.1%. In their work, the conditioning grid was without mixing vanes. They showed results of axial and cross ow velocity. The cross ow data indicated that positions close to the grid had velocities 40% greater than positions farther from the grid. The velocity uctuations also increased as much as 100% at positions far from the grid compared to positions close to the grid. In the case of the axial-ow velocity, the mean ow velocity at X = 3.1Dh was smaller than at X = 20.5Dh and uctuations were greater closer to the grid. It is found that for this type of grid

the ow recovers after a distance of X/Dh = 10 for axial ow velocity. For the case of cross ow situations the uctuations remain constant after X/Dh = 10 but the mean value of the velocity component decreases with distance from the grid. Mean cross ow velocity and turbulence in the fuel bundle are bigger near the spacer grid; recovery of mean ow velocity and a reduction in turbulence were observed downstream. These ow behaviors were assumed to be mainly due to the mixing vanes. The results indicate that the mixing vane effect has a strong inuence to around X = 10Dh downstream of the spacer grid. Conner et al. (2005) presents the experimental results of a 5 5 fuel bundle with spacer grids. Several spacer types typical of PWRs were tested in an air loop with fully heated rods. The work focuses on heat transfer measurements using a specially design thermocouple holder than can be moved axially inside the bundle. The Re number tested ranged from 15,000 to 37,000 based on the hydraulic diameter, Dh = 11.77 mm. The results were calculated based on temperature measurements at discrete positions inside the bundle. Conner et al. showed an improvement in the heat transfer after the spacer-grids but did not provide any explanation about the probable mechanism behind these improvements. Yang and Chung (1998) have studied the inuence of the spacer grids on the turbulent mixing within square sub-channel geome-

Fig. 2. Test section showing the optical arrangement and PIV system.

890

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

Fig. 3. Measurement system.

try. They analyzed the generation and decay of turbulent energy from LDA measurement results. Neti et al. (1982) measured the lateral velocity proles within sub channels using 2D-LDA and they claimed that the magnitudes were less than 1% of the averaged axial velocity. However, detailed secondary ow features were unclear. Rehme (1987) performed measurements using hot wire anemometry (HWA) in order to conrm the existence of lateral ow pulsations between the sub-channels. It is concluded that the thermal mixing in sub-channel geometry is mostly inuenced by almost periodic macro scale ow pulsations caused by the ow instability rather than the secondary ow from the Reynolds stress gradients. Regarding computational work, several attempts have been made to model the thermal hydraulic behavior of fuel rod bundles. Tzanos (2004) simulated the ow in a 7 7 rod bundle typical of a PWR. The spacers were modeled as plates with no thickness and the blockage ratio of the spacer was used as controlling parameter. The simulations were made using various turbulence models with symmetric and periodic boundary conditions. Tzanos concluded that far from components that cause signicant ow deections the agreement of mean velocity predictions with measurements was good. However, near such components the discrepancy between velocity predictions and measurements could be signicant. These discrepancies were attributed to shortcomings of the k models. Even in rod bundles without ow deectors, the turbulence predictions of standard k models showed signicant discrepancy with measurements. Ikeno and Kajishima (2006) used large Eddy simulation (LES) coupled with the immersed boundary method in an effort to represent the effect of ow geometry and compared some of their results with particle image velocimetry (PIV) focusing on developing swirl. The Re number for this investigations was Re = 4100. The

results indicated the effects of: the mixing-vanes which caused the swirl with large-scale uctuation and enhancement of heat transfer. The produced vortices in the rod gap caused enthalpy mixing enhancement between channels. The model successfully explained the mechanism of decaying swirl in the rod bundle: a developing vortex in the rod gap and the decreasing wake behind the mixing-vanes promoted the decay of swirl more strongly than in a pipe. Lee and Choi (2007) focused on the computational analysis of turbulence intensities, maximum surface temperatures of the rod bundle, heat transfer coefcients and pressure drops of four kinds of mixing vanes. The authors used a 17 17 rod bundle since they considered that vortex size effect may not be represented correctly with a smaller rod array. Caraghiaur and Anglart (2007) measured the axial velocity in three different sub-channels of 5 5 rod bundle with spacers using LDV. The results were compared with CFD predictions using a k and SSG turbulence models. It should be noted that the spacer did not have any turbulence enhancement device such as mixing vanes. In this study both turbulence models over-predicted the rate of the velocity downstream of the spacer. In addition, the root mean square values of the axial velocity did not agree with the CFD prediction. Caraghiaur and Anglart concluded that the pressure drop over the spacer was under-predicted by 2030% in comparison with CFD code calculation. There was a reasonable agreement between axial velocity distributions downstream of the spacer with a larger CFD model i.e. with half of the bundle geometry to remove the inuence of symmetry conditions imposed on sub-channel boundaries in the CFD predictions. Verication and validation (V&V) are the primary means to assess the accuracy and reliability of computational simulations. V&V methods and procedures have fundamentally improved the

Fig. 4. Cross sectional conguration of the rod bundle assembly.

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

891

Fig. 5. A 5 5 spacer grid segment lled with FEP tubes emulating a fuel-bundle. (a) Pipes immersed in water. (b) Pipes immersed in air.

credibility of simulations in several elds, such as aerospace industry. The eld of nuclear reactor safety has placed great emphasis on developing validation benchmarks. Many of these validations benchmarks are closely related to the overall behavior for determination of velocity, temperature, and void fraction among other parameters. However, detailed information of local velocity and temperature in full-eld are not available for CFD validation.

The importance of computer simulations in the design and performance assessment of engineered systems has increased dramatically during the last three decades. To have justied condence in this evolving eld, there should be major improvements in the transparency and maturity of the computer code used, the clarity of the physics included or excluded in the modeling, and the comprehensiveness of the uncertainty assessment performed. CFD codes

Fig. 6. Test section showing the optical transparency achieved inside the bundle by matching the refractive index of the FEP rods and water.

892

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

Fig. 9. Top view of rod bundle showing the planes selected for the PIV measurements.

Fig. 7. Horizontal laser sheet passing across the rod bundle without distortion due to a good matching index of refraction.

carried out using dynamic particle image velocimetry (DPIV) inside an optically transparent rod bundle utilizing the matching index of refraction (MIR) approach (Hassan and Dominguez-Ontiveros, 2008). This work presents results showing full eld velocity vectors and turbulence statistics for the bundle under single phase ow conditions. 2. Experimental set-up The experiments were conducted in the test loop of the optical multi-phase ow research laboratory (OMRL) at Texas A&M university (TAMU) allowing for test pressures up to 3 atm. A schematic of the hydraulic ow loop is shown in Fig. 1. The ow loop consisted of a variable speed pump, storage tank, ow meter, ow straightener, mesh, and test section. Water was used as the working uid, and it owed from the bottom of the test section in an upward direction. The ow rate was controlled by the rotational speed of the pump

should be applied with care to the ows where dominant physics are claried. Thus, the ultimate goal of this work is to contribute to a better understanding of presupposed and existent events that could affect the safety of nuclear power plants by using state-of-theart measurement techniques that may elucidate the fundamental physics of uid ow in rod bundle with spacer grids. In particular, this work concerns the development of an experimental data base with high spatial and temporal resolution of ow measurements inside 5 5 rod bundles with spacer grids. The data base is intended to validate CFD codes at various scales. Measurements are

Fig. 8. Multi-scale measurements.

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

893

Fig. 11. Average velocity elds. Fig. 10. Typical instantaneous velocity elds.

using a frequency controller and a control valve in the main loop in order to assure reproducibility of ow rates and constant torque in the pump impeller. The uid passed through a series of lters that limited the size of suspended particles in the uid to 50 m during experiments. The instrumentation and control stages for the hydraulic loop contained turbine type ow meters and rotameters. Pressure transducers were installed along the test section to measure the local differential pressures across the spacer grids and the loop pressure. The data was collected using a data acquisition unit NI-SCXI with the 1000/1600/1314 modules for signal conditioning, analog to digital conversion, and transmission. A dynamic PIV system was used to measure full velocity elds. The system was comprised of a high repetition Nd:YaG twin laser, optics, high accuracy pulse generator, motorized position stages, and high speed cameras as shown in Fig. 2. The maximum laser pulse frequency was 20,000 pulses per second with a 527 nm wavelength and maximum energy of 10 mJ per pulse. Three high speed cameras with up to 120,000 fps were used to accurately capture the uid behavior in the test section. A laser sheet with about 1 mm thickness was produced using a set of optical lenses and mirrors which impinged perpendicular to the lateral face of the channel.

Polymer seeding particles with diameter distribution of 69 m and specic gravity of 1.05 were added to the uid in a volume concentration of 1.8e2% for the PIV measurements. Data was obtained at various planes of the test section. Cross-ow measurements were obtained by using a second laser sheet to form a horizontal plane parallel to the spacer grid span. The horizontal laser sheet with a 2.5 mm thickness was shaped using a 50% mirror arrangement as shown in Fig. 3. A third high speed camera was placed in the test section with an inclination of 55 in order to capture the data obtained using the horizontal laser sheet. 2.1. Test section For observation of the ow structure the test section housed a square-arrayed 5 5 rod bundle. The test assembly consisted of 24 fuel rods with a diameter of D = 10.25 mm placed inside the egg-crated spaces of the spacer subsection and one square rod with a side length of 12.5 mm as shown in Fig. 4. The rods were arranged in a square array with a P/D = 1.26. This 5 5 rod bundle with spacer grids was installed in the aluminum square duct. Both ends of the rod bundle were xed with the same type of spacer grids. The square duct was made of an aluminum vertical channel with dimensions 101.6 mm 101.6 mm 1828 mm. The channel had four observation windows in order to have optical

894

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

Fig. 13. Typical instantaneous velocity elds. Fig. 12. Average velocity elds with calculated streamlines.

access to the channels interior faces. A set of four spacer grids were located inside the channel with a distance between them of 508 mm. However, the grid located at the end of the bundle on the downstream side was 250 mm from the edge of the previous spacer. The special shallow plastic rods are located inside the spacer grid in order to simulate the fuel rods. The rods were fabricated using a uorinated ethylene-propylene (FEP) plastic with dimensions of 10.25 mm OD 10 mm ID 1270 mm long with a 0.07 mm tolerance in all dimensions (Dominguez-Ontiveros and Hassan, 2007). These dimensions followed closely the dimensions of the real fuel rods used in the UltraowTM spacers (Kraemer et al., 1995). The rods were sealed at each end with a conical aluminum cap. The rods were lled with a solution of 5% chlorine in water. This solution kept the refractive index of water close to its original value of 1.33 and prevented the growth of bacteria and algae in the uid, which could alter the optical transparency of the stagnant solution inside the rods. The measurements were taken around the second spacer grid, with the rst spacer grid serving as a conditioning grid for the measurement zone. The spacer grids had mixing devices of the swirl type attached to the edge of the grid. The velocity components and their uctuations were measured in the test section at several locations upstream and downstream

the grid by particle image velocimetry. The local turbulence quantities were integrated over the uid section to obtain the value of turbulence in different sections. This value could then be used as boundary condition for CFD simulation. The ow velocities were measured using varying axial elevations from 0.1 to 6 hydraulic diameters, with the top edge of the spacer-grid as the reference point. 2.2. Matching index of refraction Optical ow studies of complex geometries such as fuel bundles were shown to be possible by matching the index of refraction of the solid test section with that of the working uid. When light passes, through materials with a different index of refraction, the light bends at the interface of the materials causing distortion or hidden areas. However, by matching the refractive index of the materials, the bending of the light was minimized or eliminated, making the eld of interest accessible and the test section optically transparent. MIR techniques have been used in past studies (e.g., Hassan and Dominguez-Ontiveros, 2008). Although matching the refractive index of the solid and liquid was not difcult in itself, however, other concerns such as uid price, viscosity, density, toxicity, ammability, compatibility, etc. complicated the selection process. In this work, optical transparency was achieved by match-

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

895

the FEP shallow rods were fabricated with a thin wall for transparency and lled with a solution of water and chlorine to enhance rod strength and prevent bacteria growing. 3. Results The measurements for axial velocity at the investigated region were performed at several planes using a multi-scale approach. This approach consisted of performing measurements on various viewing areas of the test section. The scales were selected based on the physical constrictions and the desired resolution of the velocity elds. Fig. 8 shows the different scales used in this work. The rst scale examined the ow behavior in the planes that cover the ve rods and its respective sub-channels which are denoted by the yellow dashed line. At this scale the interaction between sub-channels and general characteristics of the ow were captured. The equivalent viewing area of the cameras for this scale was 70 mm 60 mm. The second scale is denoted by the area marked with a solid red line in Fig. 8. This scale captured the ow characteristics of a central sub-channel with an equivalent camera viewing area of 15 mm 12 mm. At this scale the ow evolution in the sub-channel could be observed. The third scale is denoted by the red dashed line in Fig. 8 and it has an equivalent viewing area of 5 mm 4 mm. This scale was intended to resolve the local ow behavior, such as boundary layer thickness, turbulence energy production, turbulent energy dissipation, eddy size, coherent structures identication, among other important parameters of interest. The data collection process included the measurement of every scale at various planes inside the 5 5 bundle. The selected planes were parallel to the frontal channels window and they varied in distance from the wall to the center of the bundle by changing the position of the optical set-up and cameras using automatic motorized positioning stages as shown in Fig. 9. Fig. 10 shows the PIV measurement result of the axial velocity vectors in the investigating region downstream and upstream the spacer grid. The velocity elds show the whole velocity distribution in the region from 0 to 1.2Dh for both cases. 3.1. Velocity eld around spacer
Fig. 14. Average velocity elds.

ing the index of refraction. The immersed plastic rods have a similar index of refraction as water. The plastic rods are made from FEP polymer with a refractive index of 1.338, and lled with a solution of water and chlorine with a refractive index of 1.333. By matching the index of refraction of both materials optical transparency was achieved. The working uid for the tests was water, therefore no correction is necessary and the transparency of the test section was assured. Fig. 5 shows an emulated hot-rod (red color) embedded in the central rod position. The rod is optically accessible when the bundle was submerged in water. The matched refractive index of water and the FEP pipes allowed for the visualization of the central rod without considerable disturbances from the rod-layers surrounding it. Fig. 6 presents another pictorial of the achieved transparency in the setup. The vertical channel was partially lled with water to demonstrate the immediate clearness and accessibility of any point inside the grid spacer. The unobstructed path of the blue laser light sheet as it passed through the rod bundle is demonstrated in Fig. 7. It was observed that light was not signicantly distorted demonstrating the optical transparency obtained by this technique. As indicated by Hassan and Dominguez-Ontiveros (2008), the wall-thickness of the cylindrical rod must be chosen to achieve an optimal compromise between rod strength and optical transparency. Consequently,

Several different techniques were used to analyze the behavior of the channel ow. Velocity statistics were obtained from PIV measurements in the xy plane. These velocity statistics were determined by ensemble averaging over the number of velocity elds using Eq. (1),
N

Ai (X) A=
i=1

(1)

where A is the quantity being averaged, X is the position vector, and N is the total number of realizations. The result from Eq. (1) was a two-dimensional average vector eld. The uctuating velocity elds were obtained using the Reynolds decomposition for each vector in each instantaneous velocity eld using the following expression, u=U+u (2)

where U is the mean velocity and u is the velocity uctuation. The mean streamwise velocity determined in this fashion was subtracted from each instantaneous velocity eld to obtain the uctuating velocity elds using Eqs. (3) and (4). These uctuating velocity elds were then ensemble averaged to obtain the streamwise and normal intensities as well as the Reynolds shear stresses

896

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

Fig. 15. Turbulence intensities and Reynolds stresses below and above the spacer grid for Re = 6500.

using Eq. (5). ui (x, y) = 1 N 1 N


N

[ui (x, y) U(x, y)]


i=1 N

(3)

vi (x, y) =

[ui (x, y) U(x, y)]


i=1 N

(4)

u v (x, y) =

1 N

ui (x, y)
i=1

vi (x, y)

(5)

Instantaneous velocity vectors are shown in Fig. 10. The swirl was caused by the ow around the mixing-vane. The swirl contains large-scale uctuation, which affected the downstream turbulence.

Structure and intensity of the downstream turbulence are important for the promotion of heat transfer on the rod surface. As shown in Fig. 10, the velocity eld along the rod surface was captured reasonably by the PIV measurements. The ow elds are presented for both positions: downstream and upstream the spacer grid. These instantaneous velocity elds show non-uniform direction due to the swirling motion of the uid after having passed through the spacer. There is a change in the maximum velocity attained in the channel of 43% with respect to the maximum measured downstream from the spacer. To obtain the mean value the data was averaged from more than 45,000 instantaneous vectors. Axial evolutions of the mean streamwise velocity are shown in Fig. 11. As shown in Fig. 11 the ow was accelerated through the sudden narrowing channel spacer. The position of local maximum velocity was located at the center of the channel with a distance from the spacer

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

897

edge X = 6Dh . The averaged data shows a uniform direction prole of the velocity eld after having passed the spacer which elucidates the effect of averaging. On the other hand, the velocity prole downstream of the spacer grid showed a similar prole to that found in shear ow along no-slip wall conditions. One of the advantages of the DPIV obtained measurements was the direct comparison of instantaneous data with CFD which was previously done using an averaging approach. The plane at a distance of X = 0.3Dh intersected with the mixing vanes, where the ow was the wake ow behind the mixing-vane, a local acceleration in the narrow gap between the rod and the tip of mixing-vane is shown in Fig. 12. Streamtraces were calculated from the instantaneous and averaged velocity elds in an effort to visualize the ow behavior below and above the grid. Fig. 12 shows the streamtraces of the velocity averaged velocity elds. It was observed that the deection of the ow was caused by the mixing vanes. The results showed the complex inuence of the spacer grid in the sub-channel formed between the wall surfaces of two adjacent rods. The results of 2D velocity elds showed a considerable reduction in the axial velocity component of the velocity vector at the center of the channel. This change in magnitude is compensated by an increase in the radial component of the velocity vector. Flow visualization demonstrated swirling immediately above the spacer grid. Figs. 13 and 14 show the instantaneous and averaged velocity elds respectively downstream and upstream the spacers in the second plane measured. As previously explained in this section, the PIV measures were performed at various planes using motorized positioning stages. The results shown in Figs. 13 and 14 correspond to the plane 2 illustrated in Fig. 9. Since the rods are optically transparent when the water is owing through the channel, the instantaneous velocity elds could be obtained. It was observed that measured vectors in this plane had strong direction changes caused by the mixing vanes compared to the uniform vector direction seen downstream from the grid. Once again, the averaged velocity elds are illustrated in Fig. 14. Axial evolutions of turbulence intensities and Reynolds stresses are shown in Fig. 15. The turbulence intensity was based on the deviation from the mean velocity. Vortex cores were identied using the Galilean decomposition approach. In the Galilean transformation the total velocity is represented as the sum of a constant convection velocity, Uc , plus the deviation as shown in Eq. (6) u = Uc + uc (6)

Fig. 16. Vorticity modication below and above the spacer grid for a Re = 6500. The velocity eld spans for about 1Dh from the edge of the spacer.

ter was divided by the spacer edge. Turbulence models based on a developed turbulent ow were fundamentally inapplicable to these ows.

3.2. Vorticity modication Vorticity could be calculated using the average, the instantaneous, or the uctuating velocity eld. One of the advantages of PIV measurements is that it offered information about the vorticity eld from the spatial distribution of velocity. In this case, since measurements were carried out in the xy plane, only one component (spanwise) of the vorticity vector could be obtained. Therefore, the expression for instantaneous spanwise vorticity is v u y x

In an effort to locate vortex cores, Adrian et al. (2000), selected different fractions of the centerline velocity as the convection velocity implemented in the Galilean transformation technique. The resulted vortex cores and length scale are illustrated in Fig. 15 with red circles. The averaged uctuating velocity vectors are overlapped in the contour plot showing the magnitude of the turbulence intensities. This gure shows the reduction in vortex cores as a direct consequence of the spacer grids inuence on the ow behavior. The swirl decreased with distance from the mixing-vanes. Predicting this decrease rate is important to evaluate heat transfer at the rod surface. The mean axial velocity decreased with an increase in turbulence when passing the spacer grid. The axial ow recovered after several diameters from the grid by settling down the ow eld with distance from the grid. The results indicate that the ow eld around the spacer included a large-scale unsteadiness and an undeveloped turbulent ow. The unsteadiness was caused by a separation above the mixing vane attached to the spacer. The separation caused a rapid increase of the pressure drop against inclination of the mixing vanes. The undeveloped turbulence appears in the inlet of the spacer, where the developed main ow in the sub-channel cen-

z =

(7)

The derivative calculation was carried out using central difference scheme. Substituting the uctuating components of the velocity in Eq. (7), the uctuating spanwise vorticity was obtained. Likewise, the substitution of the average velocity components yielded the average vorticity eld. In turbulent ows, the existence of a high average vorticity is well known. This high level of vorticity was caused by the extension or stretching of vortex laments (Tennekes and Lumley, 1974). Vortex stretching, has been regarded as the cause of the high rate of dissipation associated with turbulent motions. The effect of the spacer grid was a modication of the boundary layer caused by the mixing vanes which seemed to produce an attenuation of the vorticity magnitude. Although vorticity was always present in the ow, there seemed to be a counteracting phenomenon that suppressed the stretching in the ow. Fig. 16 shows the average vorticity elds obtained for single phase ow below and above the spacer grid. It was observed that the close to the wall there is a pronounced decrease in the vorticity as the ow passes through the spacer grid.

898

E.E. Dominguez-Ontiveros, Y.A. Hassan / Nuclear Engineering and Design 239 (2009) 888898

4. Conclusions The effect of the spacer grid was a modication of the boundary layer caused by the mixing vanes which produced an attenuation of the vorticity magnitude. Although vorticity was always present in the ow, there seemed to be a counteracting phenomenon that suppressed the stretching in the ow. There is a change in the maximum velocity attained in the channel of 43% with respect to the maximum measured downstream of the spacer. The averaged data showed a uniform direction prole of the velocity eld after passing the spacer which elucidates the effect of averaging on the results. In the other hand, the velocity prole downstream of the spacer grid shows a similar prole to those found in shear ow along a no-slip wall. One of the advantages of the DPIV obtained measurements is the direct comparison of instantaneous data with CFD which was previously done using an averaging approach. The results indicated that the ow eld around the spacer included a large-scale unsteadiness and an undeveloped turbulent ow. The unsteadiness was caused by a separation above the mixing vane attached on the spacer. The undeveloped turbulence appeared in the inlet of the spacer, where the developed main ow in the sub-channel center was divided by the spacer edge. Turbulence models based on a developed turbulent ow are fundamentally inapplicable to these ows. Successful use of PIV in addition to matched index of refraction was used for full eld velocity elds in a complex geometry without ow disturbance. The results showed that turbulence evolution can be observed using the DPIV technique. Good quality data with high spatial and temporal resolution open the possibility of CFD benchmarking and validation among important quantication of the turbulent structures present in the ow above spacer grids. The experimental results from this work may be useful for the development of a turbulence model for applications such as sub-channel geometries, where the ow is highly non-isotropic.

References
Adrian, R.J., Christensen, K.T., Liu, Z., 2000. Analysis and interpretation of instantaneous turbulent velocity elds. Experiments in Fluids 29, 275290. Caraghiaur, D., Anglart, H., 2007. Measurements and CFD Predictions of Velocity, Turbulence Intensity and Pressure Development in BWR Fuel Rod Assembly with Spacers. Proceedings of the 12th International Meeting on Nuclear Reactor Thermal Hydraulics (NURETH12), Pittsburgh, PA, September 30October 4. Chang, S.K., Moon, S.K., Baek, W.P., Choi, Y.D., 2008. Phenomenological investigations on the turbulent ow structures in a rod bundle array with mixing devices. Nuclear Engineering and Design 238, 600609. Conner, M.E., Smith, L.D., Holloway, M.V., Beasley, D.E., 2005. Heat Transfer Coefcient Testing in Nuclear Fuel Rod Bundles with Mixing Vane Grids, Water Reactor Fuel Performance Meeting, Kyoto, Japan, October 26. Dominguez-Ontiveros, E.E., Hassan, Y.A., 2007. Jet ow mixing measurements in the vicinity of rod bundles using PIV. Transactions of the American Nuclear Society Winter Meeting, Washington, DC, November 1115. Hassan, Y.A., Dominguez-Ontiveros, E.E., 2008. Flow visualization in a pebble bed reactor experiment using PIV and refractive index matching techniques. Nuclear Engineering and Design 238, 30803085. Ikeda, K., Hoshi, M., 2006. Development of rod-embedded ber LDV to measure velocity in fuel rod bundles. Journal of Nuclear Science and Technology 43, 150158. Ikeno, T., Kajishima, T., 2006. Decay of swirling turbulent ow in rod-bundle. Journal of Fluid Science and Technology 1, 3647. Kraemer, W., Proebstle, G., Uebelhack, W., Keheley, T., Tsuda, K., Kato, S., 1995. The ULTRAFLOW spaceran advanced feature of ATRIUM fuel assemblies for boiling water reactors. Nuclear Engineering and Design 154, 1721. Lee, C.M., Choi, Y.D., 2007. Comparison of thermo-hydraulic performances of large scale vortex ow (LSVF) and small scale vortex ow (SSVF) mixing vanes in 17 17 nuclear rod bundle. Nuclear Engineering and Design 237, 23222331. Neti, S., Eichhorn, R., Hahn, O.J., 1982. Laser Doppler measurements of ow in a rod bundle. Nuclear Engineering and Design 74, 105116. Rehme, K., 1987. The structure of turbulent ow through rod bundles. Nuclear Engineering and Design 74, 105154. Simonin, O., Barcouda, M., 1988. Measurements and Prediction of Turbulent Flow Entering a Staggered Tube Bundle. Proceedings of the 4th international symposium on applications of laser anemometry to uid mechanics, Lisbon, Portugal, July. Tennekes, H., Lumley, J.L., 1974. A First Course in Turbulence. MIT Press, Cambridge, Massachusetts. Tzanos, C., 2004. Computational uid dynamics for the analysis of light water reactor ows. Nuclear Technology 147, 181190. Yang, S.K., Chung, M.K., 1998. Turbulent ow through spacer grids in rod bundles. Journal of Fluids Engineering ASME 120, 786791. Zdravkovich, M.M., 1987. The effects of interface between circular cylinders in cross ow. Journal of Fluids and Structures 1, 239261.

Das könnte Ihnen auch gefallen