Anda di halaman 1dari 11

Proceedings of the ASME 2012 International Mechanical Engineering Congress & Exposition IMECE2012 November 9-15, 2012, Houston,

Texas, USA

IMECE2012-85203
RACK LEVEL MODELING OF AIR FLOW THROUGH PERFORATED TILE IN A DATA CENTER
Vaibhav K. Arghode G.W. Woodruff School of Mechanical Engineering Georgia Institute of Technology Atlanta, Georgia, USA Yogendra Joshi G.W. Woodruff School of Mechanical Engineering Georgia Institute of Technology Atlanta, Georgia, USA Pramod Kumar Department of Mechanical Engineering Indian Institute of Science Bangalore, Karnataka, India Thomas S. Weiss Triad Floors Inc. Denver, Colorado, USA

Gary Meyer Triad Floors Inc. Denver, Colorado, USA ABSTRACT Effective air flow distribution through perforated tiles is required to efficiently cool servers in a raised floor data center. We present detailed computational fluid dynamics (CFD) modeling of air flow through a perforated tile and its entrance to the adjacent server rack. The realistic geometrical details of the perforated tile, as well as of the rack are included in the model. Generally models for air flow through perforated tiles specify a step pressure loss across the tile surface, or porous jump model based on the tile porosity. An improvement to this includes a momentum source specification above the tile to simulate the acceleration of the air flow through the pores, or body force model. In both of these models geometrical details of tile such as pore locations and shapes are not included. More details increase the grid size as well as the computational time. However, the grid refinement can be controlled to achieve balance between the accuracy and computational time. We compared the results from CFD using geometrical resolution with the porous jump and body force model solution as well as with the measured flow field using Particle Image Velocimetry (PIV) experiments. We observe that including tile geometrical details gives better results as compared to elimination of tile geometrical details and specifying physical models across and above the tile surface. A modification to the body force model is also suggested and improved results were achieved. Key words: High density rack; Perforated tile; Air flow distribution; Geometrical resolution; Porous jump model; Body force model

INTRODUCTION In recent past, many investigations have been reported on modeling of data center air flows [2-11]. Some of the investigations only focused on the under floor region (plenum) modeling [3-5] and they imposed a step pressure loss across the tile surface, or porous jump model thereby eliminating any geometrical details of the tile from the model. The pressure loss factor (K=P/0.5Vin2) was calculated based on the tile porosity (open area ratio) [1]. The focus of these works was to predict the air flow rate emerging from the plenum. The suggested models include a 0D model where a uniformly pressurized plenum was considered [3] or 1D model where the effects of under floor pressure variation as well as frictional resistance in plenum was considered. These simplified model results in manageable grid size and reduced computational time for modeling large data centers. Such investigations are helpful for understanding the overall air supply pattern in the data center with objective of supplying the air flow rate as per the requirement of the adjacent rack. It was also suggested that the air flow design problem is based on the under floor flow features and supplying the required amount of air is sufficient to effective cool the adjacent server rack [2]. However experimental investigation of air flow emerging from perforated tile and entering the rack suggests that even though the supplied air flow rate matches the rack requirement, issues such as air bypass and room air entrainment may be present [13-14]. Hence modeling of above

Copyright 2012 by ASME

floor system is deemed useful and has been considered in investigation such as [8-11]. In a more comprehensive investigation, both under floor and above floor regions have been investigated together [6]. In this investigation the two regions were connected by perforated tiles where a step pressure loss is specified through the tile surface (porous jump model). A review of the computational investigations pertaining to data center air flows could be found in [7]. An improvement in the tile flow model was suggested in [10-11]. However, in these investigations the focus was limited to the above floor surface only where designated air flow rates were specified at the tile inlet. It was mentioned that in the porous jump model (which considers uniform velocity based on total open area of tile), the momentum rise of air due to pressure loss across the tile surface is not captured and hence in these investigations additional body force (momentum source) was specified above the tile surface to simulate higher momentum of the air flow above the tile surface (body force model). This model showed improved results as compared to the porous jump model [10-11]. Other strategies such as modeling the tile with an equivalent reduced single open area to simulate higher momentum or multiple openings were also suggested [10]. Multiple openings with some tile geometrical details (much larger pore length scales as compared to the actual pore size) was also investigated and fair agreement with measurements was obtained [10]. In this case higher air entrainment and faster mixing was observed as compared to the body force model which eliminated pore geometrical details. This suggests that further inclusion of finer geometrical details can be helpful to closely capture the flow pattern above the floor and at rack entry. Hence, in the present work we report the effect of including more resolved tile geometrical details on the solution accuracy at the expense of increased computational effort. Note that a balance between the solution accuracy and computational effort needs to be considered while investigating data center air flows. We also present a modification to the body force model which could be useful to simulate the flow field with much reduced computational efforts and it also shows promise for further improvements to develop simplified models which will reduce the computational time while still maintaining acceptable accuracy. COMPUTATIONAL SETUP Figure 1 shows the server simulators, rack and the perforated tile used in the experimental investigation [13-14]. The 42U (1U = 4.45 cm) rack houses four server simulators each having height of 10U (see Figure 1(a)). There is a gap of about 2.54 cm (1 inch)

between the bottom of the rack and the solid tile surface below the rack (see Figure 1(b)). Server simulators have a front plastic grill with porosity of about 37% [12]. Each server simulator has four fans mounted on a plate. The fan-plate is located at a distance of 25.4 cm (10 inch) from the front of server simulator. The fans are centered and spaced 5.08 cm (2 inch) apart from fan outer edge. Each fan has an outer diameter of 16.83 cm (6.625 inch) and 8.26 cm (3.25 inch) hub diameter. Each server simulator has fan speed dial settings to set desired flow rates through the rack. Further details of the server simulator could be found in [12]. The tile used for the investigation has dampers on the bottom and a perforated plate on the tile top surface. The dampers have multiple rectangular openings (see Figure 1(d)) and the tile top has openings of different shapes with length scales of about 1.27 cm (0.5 inch) (see Figure 1(c)). The distance between the dampers and the tile top is about 3.81 cm (1.5 inch). Figure 2 shows the computational domain and modeled geometrical tile details. The computational domain considers only one tile and one rack as shown in the Figure 2(a). All the geometrical features are modeled as orthogonal shapes. Circular fans with hub are modeled as square openings with square hub. The rack geometry after the fan plate is not included in the model and the fans are considered as exhaust fans with target mass flow boundary condition. In this case the pressure at the fan surface is adjusted so as to reach a pre-assigned mass flow rate through the fan surface. The swirl generated by the fans is not included in the model. Figure 3 shows the dimensional details of the computational domain. The grill at the rack inlet is modeled as thin rectangular openings of about 1.27 cm (0.5 inch) height. The impermeable strips adjacent to the openings have height of 2.54 cm (1 inch) hence giving an overall porosity of 33% (see Figure 3(d)). All four server simulators are separated by impermeable walls as shown in Figure 3(a). The gap of 2.54 cm (1 inch) between the bottom of the rack and solid tile below it is also considered in the model and pressure inlet boundary condition is specified at the inlet of the gap (see Figure 3(a)). Plenum is included in the model by specifying mass flow rate boundary condition below the tile surface with symmetry boundary condition on the plenum walls (see Figure 2(a), 3(a), 3(b)). For the present investigation the plenum is fairly deep (~ 0.9 m or 3 ft) and hence the effect of directional flow under the tile surface is expected to be less in the present case [2]. Moreover, two down flow CRAC units were used which were located on either sides of the cold aisle (along the length). This would further

Copyright 2012 by ASME

suppress the directional flow characteristics in the plenum, see [13-14] for data center layout. Hence, mass flow condition is specified below the tile surface and not on the sides of the modeled plenum. The aisle top is specified as pressure outlet and the aisle center is specified as pressure inlet. The aisle center is the aisle surface facing the rack inlet. The sides of the aisle are specified as pressure inlet with some resistance to the air flow (pressure loss coefficient, K = 10). This condition was used due to the anticipated resistance to the flow entering the computational domain because of the presence of adjacent tiles. Refer [13-14] for further details on tile layout for the experimental investigation. The value of K was appropriately adjusted to simulate the flow field features obtained by experiments. However, further investigation is required to arrive at an appropriate value for K based on the physical set-up and operating conditions. The modeled tile geometry is shown in Figure 2(b-f). The tile top surface is modeled as 900 square pores of 1.27 cm (0.5 inch) side. Note that the tile top surface has pore shapes which have curved edges, however in our investigation we observe that the curved pores could be modeled as orthogonal shapes with same hydraulic diameter [15]. The spacing between two square pores on the tile top is about 0.64 cm (0.25 inch). The porosity of the modeled tile top surface is about 39%. The damper openings are modeled as 36 rectangular openings, each with size of 3.18 cm (1.25 inch) 13.34 cm (5.25 inch). The distance between two damper openings is 0.64 cm (0.25 inch) along the longitudinal direction and 3.18 cm (1.25 inch) in the transverse direction. The porosity of damper openings is 41%. The tile top and the dampers are spaced 3.81 cm (1.5 inch) apart (see Figure 3(a)). The computational domain is meshed with full rectilinear grid. The region near the perforated tile is appropriately refined such that each square pore on the tile top has 8 8 cells. In this case grid refinement was performed using hanging-node approach. Simulations suggest that having 8 8 cells in each pore gives reasonably close result with a grid independent solution [15]. Note that appropriate balance between grid refinement and solution accuracy is sought here. For the case using geometrical details the grid size is about 4.2 million cells. Without grid refinement, for the cases using simplified models such as porous jump or body force models the grid size is about 1 million cells. The isothermal flow field was solved using a steady state, finite volume based method and SIMPLE algorithm was used for pressure velocity coupling [16]. Symmetry with only one-half the geometry was used to reduce the computational time. k- realizable

turbulence model was used with inlet turbulence intensity of 5%. For convective term discretization, second order upwind scheme was used. Convergence was assumed when the normalized residuals for all the variables were less than 1E-04 and the selected velocity monitors have stabilized within variation of 1%. Commercial CFD software FLUENT was used for the simulations. The results are compared with the experiments [13-14], and with the simplified models specifying porous jump boundary condition as well as body force source above the tile surface. Porous jump model: Porous jump model specifies step pressure loss across the tile surface [2-9]. Generally, the magnitude of pressure loss factor is calculated based on the porosity of the tile. The pressure loss specification has a form similar to that given below in Equation (1-2) [1]. (1) = = , = , =

+ 1

(2)

It may be noted that use of this model results in no alteration in the velocity field across the tile surface. However, there is acceleration of air though the pores resulting in increased momentum flux, attempted to be captured by the body force model. It may be noted that the tile has two surfaces in series in the path of air flow. The bottom surface has tile dampers and top surface has the perforations as shown in Figure 2(b-f). The pressure drop across theses surfaces, as obtained by correlation (equation 1-2), is added to arrive at the final pressure drop. Note that in actual case the flow is not uniform after passing through the dampers and this may affect the overall pressure drop from the tile. For the present case the equivalent K across the two tile surfaces is 16.5. The equivalent porosity of the two layers is calculated as 31%. The value of K across the rack inlet (front grill of server simulators) is 14.3 corresponding to the grill porosity of 33%. Body force model: In this model a momentum source is specified above the tile surface to simulate the acceleration of air passing through the pores in the tile (see equation (3-5)) [10-11]. The magnitude of the momentum source was calculated based on the difference of momentum through the tile pores as compared to the momentum of air flow in case of uniform velocity as in the case in porous jump model (see Equation (3)).

Copyright 2012 by ASME

The momentum of air flow through the pores was calculated assuming top-hat velocity profile through the pores (see Equation (4)). = =

3 4

1.224, 2594 1.224, 2594

0.754, 1598 1.224, 2594

60% 100%

(3) (4)

= (5) , " " = =

Modified Body force model: The body force model suggested previously considers a top hat velocity profile at the tile pores while calculating the excess momentum of air above the tile surface. However, even higher momentum of the jets is expected as in reality there will be a velocity profile present at the tile pores instead of a top-hat profile. Hence, we suggest a modification to the body force model. In the modified body force model the momentum rise across the tile surface is calculated based on the pressure loss as obtained from the correlations (see Equation (6)).
= =

= + 1

(6)

Once the momentum averaged velocity (Vpore) through the tile surface is calculated the momentum source term can be calculated as suggested by the body force model (see Equation (5)). Note that the modified body force model will have higher magnitude of momentum source as compared to the body force model. RESULTS AND DISCUSSION Table 1 summarizes the cases studied. The air flow rate across the rack corresponds to a high power density rack with 15 kW of heat dissipation and bulk air temperature rise of 10 C across the rack. The tile flow rate is varied from 0% upto 100% of the rack air flow rate and the values are listed in Table 1. Further details of the experimental conditions could be found in [13-14]. Table 1. Different cases investigated. Tile flow Tile flow / Rack flow rate (m3/s, Rack flow rate (m3/s, CFM) CFM) 1.224, 2594 0, 0 0% 1.224, 2594 0.234, 496 20%

Case # 1 2

Figure 4 shows the pressure loss obtained across the tile surface from CFD simulations. In this case the tile geometrical details are included in the model as shown in Figure 2(b-f). The correlation used for comparison is given in Equation (2) corresponding to equivalent tile porosity of 31%. Figure 4 suggests a good comparison between the correlation and CFD simulations. However, note that we assume addition of pressure loss across the two tile surfaces based on the correlations and in actual case this may not represent the actual pressure loss. Value of K across the tile for cases 2, 3 and 4 are also very close to each other suggesting that the pressure loss factor does not vary significantly with respect to flow rates investigated here. Note that case 1 is not included, as there is no flow through the tile in this case. Figure 5-8 shows the comparison between experimental (PIV) and computational (CFD) results for cases 1-4 (see Table 1). Pathline plots are used for visualization of the flow field. In the CFD simulations, massless particles released from the lines at boundaries (tile, aisle center and aisle top, see Figure 5-8) and are tracked and plotted. The plots correspond to the symmetry plane positioned along the height of the rack. Note that in the CFD model considering geometrical resolution, the modeled geometry of the grill at the rack inlet is included. In the porous jump model pressure loss is specified across the grill surface. Figure 5 shows the results for case 1 (see Table 1). In this case there is no flow through the tile. From the figure we observe that the flow field measured by PIV and calculated using CFD are similar and entrainment of air from aisle top as well as the aisle center is present in both the experimental and computational results. The top region of the rack inlet shows air entrainment from the aisle top (see figure 5(a)). Figure 5(b-e) suggests that the resolution of grill geometry (rack inlet) does not significantly influence the computational solution, as compared to using the porous jump model across the grill for the present case. This also suggests that it may be beneficial to use porous jump boundary condition across the rack inlet to save computational effort for this case. Figure 6 shows the flow field for case 2 (see Table 1). In this case the flow rate through the tile is about 20% of the air flow rate through the rack. Due to insufficient supply of air from the perforated tile, the rack also entrains the air from aisle top as well as aisle center as shown in Figure 6(a). The air supplied from the tile reaches only the lower portion of the rack, while near the top the rack entrains air from aisle top. The middle portion of the rack mostly entrains air

Copyright 2012 by ASME

from the aisle center. These flow field features are also captured by all the computational models (see Figure 6(b-e)). The body force model and modified body force model show that the tile air flow reaches higher along the rack height, as compared to the porous jump model (see Figure 6(b) and 6(c)). Inclusion of body force source above the tile surface increases the momentum of bulk air in the region above the tile surface, and hence results in higher reach of air to the rack inlet. The momentum source region is between the yellow line and tile surface (see Figure 6(c-d)). The result obtained with geometrical resolution is very similar to the result obtained by the body force and modified body force models. The flow field for case 3 (60% of rack air flow through tile, see Table 1) is shown in Figure 7. From Figure 7(a) we note that even though the air flow rate through the tile is lower than the rack air requirement, a portion of the supplied air escapes from the aisle top. We can also observe delayed entry of the tile air at the lower portion of rack inlet (see bottom right corner of Figure 7(a)). This was attributed to the higher momentum of air emerging from the tile [13-14]. Air entrainment from the aisle center could also be noticed for this case (see bottom left corner of Figure 7(a)). The porous jump model is not able to capture these phenomena and the air flow from the tile reaches only up to a certain height of rack (see Figure 7(b)). Porous jump model also suggests entrainment of air from aisle top which is not observed experimentally. The porous jump model is also not able to capture the air entrainment from aisle center near the tile surface (bottom left corner) and delayed entry of air in the lower portion of rack inlet (bottom right corner). When momentum source is included above the tile surface (body force model) the tile air flow reaches higher as compared to the porous jump model (see Figure 7(c)). However air entrainment from the aisle top could be observed for the body force model too, which is not present in the experimental results. Body force model shows minor improvement in the computed flow field. In this case, small entrainment of air from the aisle center near tile (bottom left corner) and slight delay of entry of air in lower portion of rack could be observed (bottom right corner). The modified body force model shows further improvement in the results possibly due to higher magnitude of momentum source specification above the tile surface (see Figure 7(d)) as compared to the body force model. It may be noted that even with the modified body force model, the flow field is significantly different from the experimental results. The flow field with the model that included the geometrical details of the tile is shown in Figure 7(e). From the figure we observe that the flow field features are closer to the experimental results. A portion of air

flow is observed to escape from the aisle top. This model also captures the delayed entry of air to the lower portion of the rack inlet. The air entrainment from aisle center could also be observed from Figure 7(e). Figure 8 shows the flow field features for case 4 (see Table 1). In this case, the air flow rate through the tile is equal to the air flow rate through the rack. Due to higher momentum of air jet, significant portion of the tile air is observed to bypass the rack through the aisle top (see Figure 8(a), refer to [14] for the discussion). The air entrainment through the aisle center and delayed entry of tile air to the rack could be prominently observed. The porous jump model (see Figure 8(b)) shows the tile air reaching near the aisle top, however, the prominent flow field features measured experimentally, could not be captured by this model. The body force model, as well as modified body force model while showing improved prediction, could not compare closely with the experimental results (see Figure 8(c-d). Figure 8(e) shows the flow field obtained by resolving the tile geometry, which is closer to the experimental results. This model is able to capture significant bypass of the tile air through the aisle top, as well as entrainment of air though aisle center near the tile (bottom left corner). The delayed entry of the air near the bottom portion of the rack could be clearly seen from bottom right corner of Figure 8(e). This suggests that simplified geometrical resolution shows promise for better solution accuracy and further improvements in the body force, or development of other simplified models is required to capture the flow field closely. CONCLUSIONS Experimentally measured flow field in data center cold aisles were compared with various computational modeling approaches. The simplified flow models through floor tiles, such as porous jump and body force models suggests fair comparison with the experimental results at lower tile air flow rates. However, at higher air flow rates, required for high power density racks, these simplified models do not capture the flow field features accurately. A modified body force model is suggested, which calculates the momentum rise based on the prescribed pressure loss across the tile surface. While this model shows improvement in the results, it still does not compare well with the experimental results at higher tile air flow rates. Including the geometrical details of the tile provides better comparison with the experimental results. This model is able to capture the prominent flow features, such as bypass of tile air though the aisle top, as well as entrainment of air from aisle center and delayed entry of air to the rack inlet. The

Copyright 2012 by ASME

present investigation suggests that a fairly coarse grid could be used, with more geometrical details to achieve better solution accuracy as compared to specifying simplified models across the tile surface. Improvements in the simplified models show scope for future exploration to achieve a better balance between the desired solution accuracy and reduced computational time. ACKNOWLEDGMENTS This research is supported by Triad Floors, Inc. and the Consortium for Energy Efficient Thermal Management (CEETHERM). The support is gratefully acknowledged. REFERENCES [1] Freid, E., Idelchik, I. E., Flow Resistance, A Design Guide for Engineers, Hemisphere, New York, 1989. [2] Patankar, S. V., Airflow and Cooling in a Data Center, Journal of Heat Transfer, 2010, Vol. 132, pp. 073001-1-17. [3] Kang, S., Schmidt, R., Kelkar, K. M., Radmehr, A., Patankar, S. V., A Methodology for the Design of Perforated Tiles in Raised Floor data Centers using Computational Flow Analysis, IEEE Transactions on Components and Packaging Technologies, 2001, Vol. 24, pp. 177-183. [4] Karki, K. C., Patankar, S. V., Airflow Distribution Through Perforated Tiles in Raised-Floor Data Centers, Building and Environment, 2006, Vol. 41, pp. 734-744. [5] Schmidt, R. R., Karki, K. C., Kelkar, K. M., Radmehr, A., Patankar, S. V., Measurements and Predictions of the Flow Distribution through Perforated Tiles in Raised-Floor Data Centers, International Electronic Packaging Technical Conference and Exhibition, July 8-13, 2001, Kauai, Hawaii, USA. [6] Rambo, J., Joshi, Y., Convective Transport Process in Data Centers, Numerical Heat Transfer Part A, 2006, Vol. 49, pp. 923-945. [7] Rambo, J., Joshi, Y., Modeling of Data Center Airflow and Heat Transfer: State of the Art and Future Trends, Distributed Parallel Databases, 2007, Vol. 21, pp. 193-225. [8] Cruz, E., Joshi, Y., Iyengar, M., Schmidt, R., Comparison of Numerical Modeling to Experimental Data in a Small Data center Test Cell, International Electronic Packaging Technical Conference and Exhibition, July 19-23, 2009, San Francisco, California, USA. [9] Iyengar, M., Schmidt, R. R., Hamann, H., VanGilder, J., Comparision between Numerical and Experimental Temperature Distributions in a Small Data Center Test Cell, International Electronic

Packaging Technical Conference and Exhibition, July 8-12, 2007, Vancouver, British Columbia, Canada. [10] Abdelmaksoud, W. A., Khalifa, H. E., Dang, T. Q., Elhadidi, B., Schmidt, R. R., Iyengar, M., Experimental and Computational Study of Perforated Floor Tile in Data Centers, Intersociety Conference on Thermal Phenomena, Jun 2-5, 2010, Las Vegas, USA. [11] Abdelmaksoud, W. A., Khalifa, H. E., Dang, T. Q., Schmidt, R. R., Iyengar, M., Improved CFD Modeling of a Small Data Center Test Cell, Intersociety Conference on Thermal Phenomena, Jun 2-5, 2010, Las Vegas, USA. [12] Nelson, G., Development of an Experimentally-Validated Compact Model of a Server Rack. MS Thesis - 2007, Georgia Institute of Technology, GA, USA. [13] Kumar, P., Joshi, Y., Experimental Investigations on the Effect of Perforated Tile Air Jet Velocity on Server Air Distribution in a High Density Data Center, Intersociety Conference on Thermal Phenomena), Jun 2-5, 2010, Las Vegas, USA. [14] Kumar, P., Sundaralingam, V., Joshi, Y., Dynamics of Cold Aisle Air Distribution in a Raised Floor Data Center, Thermal Issues in Emerging Technologies, Dec 19-22, 2010, Cairo, Egypt. [15] Arghode, V. K., Joshi, Y., Modeling Strategies for Air flow Through Perforated Tile in a Data Center, IEEE Transactions on Components, Packaging and Manufacturing Technology, In Preparation 2012 [16] Patankar, S. V., Numerical Heat Transfer and Fluid Flow, Hemisphere, New York, 1980.

Copyright 2012 by ASME

(c) Tile top

(a) Server simulators

(b) Rack

(d) Dampers

Figure 1. (a-b) Rack and (c-d) tile details under investigation.

Copyright 2012 by ASME

(b) Tile top

(c) Mesh for each square pore in tile top

(d) Dampers

(e) Mesh for each opening in dampers

(a) Computational domain (f) Perforated tile with dampers Figure 2. (a) Computational domain and (b-f) the modeled tile.

Copyright 2012 by ASME

Side View Pressure outlet

12inch Wall Server 18inch Simulator Fan 1 Wall Server 18inch Simulator Fan 2 Wall Server 18inch Simulator Fan 3 Wall Server 18inch Simulator Fan 4 1inch Wall 5inch 1ft Pressure inlet

Front View (aisle) Pressure outlet


3inch 6inch

Front View (fan)

Front View (grill) 0.5inch open area 1inch wall

Grill Pressure inlet

Pressure inlet (K=10)

hub wall fan

Aisle

Aisle
3inch

1.5inch

Tile top Dampers Plenum Symmetry

Tile top Dampers Plenum Symmetry 24inch Mass flow inlet

3.25inch

18inch

24inch Mass flow inlet

(a) Side view

(b) Front view (c) Fans (rack exit) (d) Grill (rack inlet) Figure 3. Details of the computational domain and set-up.
20 16.1 16 16.6 16.7 16.5

K = P/(0.5Vin2 )

12 8 4 0 Case 2 Case 3 Case 4 Correlation

Figure 4. Pressure drop across the tile obtained by CFD with tile geometrical resolution.

Copyright 2012 by ASME

Velocity (a) PIV (b) Porous jump (c) Body force (d) Modified (e) Geometrical (m/s) body force resolution Figure 5. Comparison of PIV and CFD results for case 1 (rack flow = 1.224 m3/s, 2594 CFM, tile flow = 0 m3/s, 0 CFM, 0% of rack flow).

Velocity (a) PIV (b) Porous jump (c) Body force (d) Modified body (e) Geometrical (m/s) force resolution Figure 6. Comparison of PIV and CFD results for case 2 (rack flow = 1.224 m3/s, 2594 CFM, tile flow = 0.234 m3/s, 496 CFM, 60% of rack flow).

10

Copyright 2012 by ASME

Velocity (a) PIV (b) Porous jump (c) Body force (d) Modified body (d) Geometrical (m/s) model model force resolution Figure 7. Comparison of PIV and CFD results for case 3 (rack flow = 1.224 m3/s, 2594 CFM, tile flow = 0.754 m3/s, 1598 CFM, 60% of rack flow).

Velocity (a) PIV (b) Porous jump (c) Body force (d) Modified body (e) Geometrical (m/s) force resolution Figure 8. Comparison of PIV and CFD results for case 4 (rack flow = 1.224 m3/s, 2594 CFM, tile flow = 1.224 m3/s, 2594 CFM, 100% of rack flow).

11

Copyright 2012 by ASME

Anda mungkin juga menyukai