Sunteți pe pagina 1din 8

Proceedings of ASME Turbo Expo 2008: Power for Land, Sea and Air GT2008

June 9-13, 2008, Berlin, Germany

GT2008-51349
A MEANLINE MODEL FOR IMPELLER FLOW RECIRCULATION
Xuwen Qiu, David Japikse, Mark Anderson, Concepts NREC 217 Billings Farm Road ABSTRACT
Flow recirculation at the impeller inlet and outlet is an important feature that affects impeller performance, especially the power consumption at a very low flow rate. Although the mechanisms for this flow phenomenon have been studied, a practical model is needed for meanline modeling of impeller off-design performance. In this paper, a meanline recirculation model is proposed. At the inlet, the recirculation zone acts as area blockage to relieve the large incidence of the active flow at a low flow rate. The size of the blockage is estimated through a critical area ratio of an artificial inlet diffuser from the inlet to throat. The intensity of the reverse flow can then be calculated by assuming a linear velocity profile of meridional velocity in the recirculation zone. At the impeller outlet, a recirculation zone near the suction surface is established to balance the velocity difference on the pressure and suction sides of the blade. The size and the intensity of the outlet recirculation zone is assumed related to blade loading, which can be evaluated based on flow turning and Coriolis force. A few validation cases are presented showing a good comparison between test data and prediction by the model.

& = mass flow rate m

R, r = radius P = power p = static pressure s = pitch at the blade exit; s = 2R2 / Z t = thickness at the blade trailing edge U = blade rotating speed W = the relative velocity W = relative tangential velocity Subscripts: 1: impeller inlet 2: impeller exit h: hub t: tip c: critical condition p: pressure side of the blade s: suction side of the blade

INTRODUCTION NOMENCLATURE
= flow angle, positive if same as rotation direction b = blade metal angle
= impeller passage inclination angle = radius ratio (hub to tip, etc) ht = hub to tip radius ratio Flow recirculation at reduced flow rate can significantly affect impeller performance, especially power consumption. Failing to consider flow recirculation in a meanline model will lead to incorrect flow characteristics for power and efficiency at the reduced flow rate and modern design practice increasingly includes far off-design considerations. In his power balancing analysis, Stepanoff [1] found that a large portion of power could not be accounted for with the known sources of power losses at the reduced flow rate, and therefore concluded that there must be another source of power loss besides the disk, mechanical and leakage power losses, and Euler work. We now know that this extra power loss is largely due to flow recirculation. At a reduced flow rate, a recirculating zone may form at the impeller inlet and outlet (Figure 1). This flow absorbs energy when it is inside the impeller passage and loses it when the flow gets pushed out, causing additional work for the impeller. A few attempts have been made in modeling this complicated flow phenomenon and its effect. Neumann [2] presented a complicated scheme based on principle of least resistance which also, questionably, assumes that the inlet and outlet
Copyright 2008 by Concepts ETI, Inc.

ct = Rc / / Rt , radius ratio
= C / U , ratio of acquired swirl and rotation speed

= density = angular velocity of the blade rotation A = flow area AR = area ratio B = blockage b = passage width C = absolute velocity C = absolute tangential velocity Cm = meridional velocity (on Z-R plane) m = meridional distance on the Z-R plane

meridional velocities are independent. The algorithm resulted in a form that is correlated to the stage efficiency and head coefficient. Tuzson [3] estimated the effect of inlet recirculation on the power consumption and head rise based on dimensional analysis. Aungier [4] estimated the recirculation work based on Liebleins equivalent diffusion ratio. His final correlation is a function of blade loading and the outlet velocity triangle. Jansen [5] tried to model the effect of recirculation on the head rise and power using a combination of effective blade number and assuming a large recirculation circuit from inlet to exit. One of the difficulties in developing a recirculation model is the lack of accurate recirculation data. Recirculation power loss data is typically derived from torque or power measurement. It is thus critical to separate other mechanisms, such as Euler work, disk friction and leakage, from the overall power consumption in order to get reliable recirculation data. For example, simple calculations of slip factor or deviation on one hand and flow recirculation on the other are intertwined so that error in one leads to similar errors in the other. The traditional slip factor models are known to perform poorly at reduced flow rate, especially near shut-off condition, which will lead to large errors in derived recirculation power loss. Fortunately, recent progress in the slip and deviation model research (Japikse and Pelton et al. [6], Qiu, Mallikarachchi and Anderson [7]) has made this process more reliable. This has set a solid foundation for the current recirculation modeling work.

meridional momentum to pass through the impeller passage, it is forced back into the upstream of the impeller. This back flow causes the well-known phenomenon called pre-rotation since the flow upstream of the impeller now possesses a swirl velocity. With the pressure gradient in the radial direction, this pre-rotated flow will move down to the lower radius location and then get readmitted into the impeller passage, thus completing the inlet circulation path shown in Figure 1. Although many problems are created or amplified because of flow recirculation, such as flow instability, noise or even cavitation in pump applications [8], one of the beneficial effects of the recirculation flow pattern is that the large incidence of the active flow can be mitigated. The recirculation flow effectively provides area blockage for the active flow, thus increasing the meridional velocity and improving the flow incidence at the lower radii. Evidence that this flow redistribution is indeed occurring can be found in the stall recovery typically present in axial and mixed flow impellers (Figure 2). After the initial stall, the formation of the recirculation path improves the incidence of the active flow, and creates a wholly new flow state that brings the aerodynamic loss in check, allowing the head curve to rise continuously towards the shut-off condition.

FIGURE 1. INLET AND EXIT FLOW RECIRCULATION MODES

FIGURE 2. STALL RECOVERY IN AXIAL AND MIXED FLOW IMPELLERS Based on the observations above, a simple inlet recirculation model can be formulated. It is assumed that at low flow rates, a recirculation path will be formed and provide blockage to the active flow. The blockage size is such that the active flow in the unblocked section has an acceptable incidence. Following Japikses [9] Two Elements in Series (TEIS) impeller model, the inlet portion of the flow can be treated as an artificial diffuser, the area ratio of which can be defined as

RECIRCULATION MODEL
a) Inlet recirculation Two main factors contribute to the formation of the inlet recirculation flow pattern, inlet stall and radial pressure gradient. At low flow rates, large incidence angles cause the flow to separate at the blade leading edge. The stalled flow is then acted on by the impeller and obtains significant swirl velocity, which in turn creates a pressure gradient in the radial direction with the pressure at the tip higher than that of the hub. When this highly swirling flow does not have enough

AR =

cos( 1b ) cos( 1 )

(1)

The larger the area ratio becomes, the more diffusion occurs at the inlet. As the incidence becomes larger, the area ratio eventually will reach a critical value, ARcrit, and the local

Copyright 2008 by Concepts ETI, Inc.

pressure recovery will exceed a critical value, Cp,crit, at which point the diffuser will stall. The current model stipulates that once this critical point is reached, inlet recirculation occurs and provides blockage for part of the flow passage. After this point, for further reduced flow levels, the active flow may maintain much of the same flow angle corresponding to the flow point with an area ratio of ARcrit, even when the bulk flow rate decreases to the shut-off condition. This, of course, is made possible only by the enlarging of the recirculation zone, that is the blockage. Assuming that the blockage due to the recirculation zone is B1, from the continuity equation we have,

is assumed to be negligible. In other words, only impellers with substantially radial leading edges have significant inlet recirculation. Based on this observation, we can assume a simple profile for meridional velocity at the inlet (Figure 3). Between Rh and Rc, it is the active flow and has a uniform velocity of Cm1. Between Rc and Rt is the recirculation zone, where the velocity decreases linearly and reverses at radius R0. This assumption of the flow profile is supported by the inlet flow measurements at reduced flow rate, such as in the works of del Valle [11], Tanka [12], Kammer and Rautenberg [13], Shiavello and Sen [14] et al. Rc can be related to the blockage as follows:

Cm1 =

1 A1 (1 B1 )

& m

(2)

Rc 2 = ct = 1 B1 (1 ht ) Rt

(8)

With some simple manipulation, we can relate the critical area ratio ARcrit to the blockage. W 1 = U 1 C 1 (3)
2

With Rc known, the back flow velocity at the tip (Cmt), can be calculated by taking the net flow rate in the recirculation zone as zero.

cos(1 ) = 1 / 1 + (W 1 / C m1 )

(4) (5)

Rt

Rc

C mt C m1 (r Rc ) * 2rdr = 0 C m1 + Rt Rc

(9)

& )2 ARcrit = cos(1b ) * 1 + ( 1 A1 (1 B1 )(U1 C 1 ) / m

From equation (9), we can obtain Cm1.


C mt = 2 ct + 1 C m1 ct + 2

(10)

From equation (5), we can solve for the blockage required so that the active flow will not exceed the critical area ratio.

B1 = 1

& m 1 A1 (U1 C 1 )

ARcrit cos( ) 1 1b

(6)

When B1 from equation (6) is negative, no recirculation is implied and the flow has an acceptable incidence. When B1 is positive, it represents the area blockage needed so that the active flow has a flow angle corresponding to the critical area ratio, ARcrit. We are using a preliminary value of 1.5 for the critical area ratio based on our validation study, which corresponds to a Cp of 0.56. This critical pressure recovery coefficient is in agreement with the stall value observed by Japikse [10]. In practice, a small adjustment to this value may be needed, depending on the details of the inlet design, so as to better represent the initiation of the flow recirculation. Next, we will relate the blockage to the power loss in the recirculation zone. The presence of the pressure gradient is significant in helping the formation of the inlet circulation path. Without the pressure gradient, the separation zone may remain localized inside the impeller passage. The pressure gradient at the leading edge hub-to-tip direction can be calculated from the radial equilibrium equation

FIGURE 3. INLET MERIDIONAL VELOCITY PROFILE With Cmt and thus the velocity profile in the recirculation zone available, power consumption can be estimated. Assuming that the readmitting flow between Rc and R0 has little swirl, while the backflow between R0 and Rc has acquired a swirl that is proportional to the impeller rotational speed, the power loss in the recirculation zone may be estimated as follows:

C dp dp 2 cos 1 = cos 1 K 1 2 r cos 1 = db dr r


2

& = r 2 2 (C m )2rdr Precirc = U 1C 1 dm


R0 R0

Rt

Rt

(11)

(7)

where C = K1r and K1 is a constant. In equation (7), when inlet inclination angle 1 is zero, as is the case for the axial inlet, there is a positive pressure gradient from tip to hub. For a radial inlet in which the inlet inclination angle is 90, the pressure gradient in the hub-to-tip direction is zero or small. In the latter case, the stalled flow mostly stays inside the impeller passage due to the centrifugal force, and the inlet recirculation

where 1 = C 1 / U1 is the ratio between the acquired swirl and rotational speed, which based on our calibration study, is set at 0.5 in our model. This factor may need to be adjusted for different inlet design to better fit the level of recirculation power loss. Integrating equation (11), we have our final form of inlet recirculation power loss,

1 2 5 Precirc = 1 2 k ( Rt5 R0 ) + (Cm1 kRc )( Rt4 R04 ) (12) 2 5

Copyright 2008 by Concepts ETI, Inc.

where

W2 p = W 2

k=

C mt C m1 Rt Rc

W 2

(15)

kR C m1 R0 = c k
A final word should be said on the effect of inlet recirculation on head rise. Although the pre-rotated flow will certainly add a pressure rise to the flow upstream according to the current model, the pre-rotated flow is limited in the recirculation zone and does not affect the active flow. However, recirculation does implicitly help head rise by alleviating the active flow incidence. b) Outlet recirculation The mechanism for exit flow recirculation is even more complicated. Downstream components such as vaned diffusers and volutes have been known to influence exit recirculation [1,15,16]. However, one dominant factor affecting the outlet recirculation that this study identifies is the blade-toblade loading near the impeller outlet. The blade-to-blade loading, originating from the Coriolis force and the flow turning, creates a velocity difference between the pressure surface and suction surface. The velocity difference near the exit can be estimated by the following equation from Johnson [17] and Cumpsty [18].
W = Ws Wp = s * {2 sin cos W ( d dm cos sin db )} (13)

The current outlet recirculation model stipulates that the critical point is reached when W2p becomes zero. Further reduction of the flow velocity will result in the flow redistribution, making the flow separate from the suction surface, thus creating a blockage to ease the loading and to ensure that the pressure surface velocity is above zero. From equation (15) we can obtain the critical velocity that brings the velocity on the pressure side to zero:

W2 cr = K 2

W 2

(16)

where a constant K2 is added. This constant is defaulted to 1.0 in our study, and may need to be adjusted to account for other factors such as the downstream components that affect the outlet recirculation. Substituting the critical flow velocity in equation (16) with equation (14), we can obtain the blockage needed to maintain W2p at a positive level.

B2 = 1

& 2m K 2 2 A2 cos( 2 )W

(17)

When B2 from equation (17) is negative, no recirculation exists at the outlet. When B2 is positive, a recirculation zone maintains a positive velocity on the pressure surface. It should also be obvious in equation (17) that the larger the loading (W), the larger the size of the recirculation zone. Besides the effect of maintaining a desired velocity level, the exit blockage also reduces the loading near the impeller exit. Because of this blockage, the flow sees a smaller gap between the pressure surface and suction surface, which reduces the loading according to equation (12). Furthermore, according to Qiu, Mallikarachchi and Anderson [7], the reduction of loading will also lower the slip velocity, thus increasing the slip factor at a very low flow rate. Jansen [5] models this phenomena using an effective blade number. The larger the blockage, the more blades the flow effectively sees, and thus the higher slip factor.

dm

FIGURE 4A. OUTLET RECIRCULATION FLOW Note that the above equation does not apply to the exit plane, where the velocity difference approaches zero (much as the Kutta condition). Instead, this loading equation should be applied to the exit throat, or the end of the covered passage as marked by the dash-line PS in Figure 4A. This is the last plane where the flow is still guided by both pressure and the suction surfaces of the blade. Assuming there is a blockage B2 at the impeller exit, the average relative velocity can be calculated as
& m W2 = 2 A2 cos 2 (1 B2 )

(14)

FIGURE 4B. OUTLET RELATIVE VELOCITY PROFILE WITH RECIRCULATION The parasitic power consumption due to the outlet recirculation flow can now be estimated by assuming a velocity profile at the impeller exit, as shown in Figure 4B. The region from the pressure surface (PS) to the start of the recirculation zone

Assuming a linear variation of velocity from the pressure to the suction surface along the line PS, the pressure surface velocity can be calculated along with equation (13) and (14):

Copyright 2008 by Concepts ETI, Inc.

(RC) is the active flow zone; while the region from RC to the suction surface (SS) is the recirculation zone. Again it is assumed that the tangential velocity acquired through the recirculation is proportional to the impeller rotational speed. Therefore the power lost in the outlet recirculation zone can be estimated as follows:

Precirc out =

2
2

present. At the outlet, the loading is close to zero according to equation (13) and therefore there is no recirculation. Figure 5D is the efficiency plot. Without the recirculation model, the predicted efficiency is much higher than the efficiency at the low flow rate calculated from the test data. With the recirculation model, a good match appears across the flow range. To better understand the flow field, CFD runs are made to check the assumptions made in the model. Figure 5E shows 3 the result of a CFD simulation at the flow rate of 0.001 m /s. The figure shows the meridional velocity contour at the impeller leading edge and trailing edge, viewed from the axial direction. At the leading edge, reverse flow is clearly visible on the suction surface near the tip. At the exit, flow is relatively clean without visible reverse flow. Both observations are consistent with the proposed model.

2 B2 A2 2W2 cos 2U 2

(18)

where 2 = C 2 / U 2 is the ratio between the acquired swirl and impeller rotational speed in the recirculation zone. Based on our validation study, this value is set to a default of 0.7. For axial or mixed flow impellers, the radial pressure gradient can play an important role in determining the eventual location of the reverse flow in the meridional plane. Because of the radial pressure gradient, the flow with low momentum is convected from the tip to the hub, creating a strong pressure resistance at the hub. Eventually, a recirculation zone is formed near the hub, which can be typically observed in an axial impeller. In radial impellers, the axial to radial turning of the flow passage produces a secondary flow that tends to drive the low momentum flow to the shroud. Depending on the strength of the secondary flow, the low momentum zone may end up near the shroud (Johnson and Moore [19]). Regardless of the final location of the recirculation flow, the current outlet recirculation model should remain valid. The largest contributor to the relocated low momentum flow is from the blade-to-blade near the impeller exit, which is the root cause of the flow separation and redistribution. The other contributors include tip clearance and end-wall boundary layer, which may also become significant in certain designs.

CASE STUDY AND VALIDATION


FIGURE 5A. AN AXIAL INDUCER PUMP To validate the new recirculation model, three cases are discussed in the following. The first case is a 3-blade axial inducer pump (Figure 5a). The blade has a helix shape with the blade angle 11.4 measured from the tangential direction. The test data is taken from Carpenter [20], where detailed measurements of power and head rise are reported. Assuming that there is no recirculation loss and that all of the hydraulic power consumed is being used to provide the angular momentum of the flow, the exit flow angle, and thus the slip factor, can be derived from measured torque or power using the Euler turbomachinery equation. Figure 5B shows the slip factor derived from the test data. At the low flow rate the derived slip factor is larger than 1.0, which means that the flow is leading the impeller. This is of course unrealistic. This exercise confirms the existence of another source of power consumption: the inlet recirculation. Figure 5C shows the comparison between the test data and the model prediction. When no recirculation loss model is used (see the long dash line in Figure 5C), the power shows a significant drop with the decrease of the flow rate, which is in contrast to the rising trend of the test data. Using the current model, initial recirculating flow is predicted around 45% of the design mass flow rate, and the power predicted matches quite well with the test data at the low flow rate (solid line, Figure 5C). Also shown on Figure 5C are the inlet and outlet power loss due to recirculation. In this case, only inlet recirculation is
1.8 1.6 1.4 derived slip factor 1.2 1 0.8 0.6 0.4 0.2 0 0 0.001 0.002 0.003 flow rate (cm s)

Figure 5b. Derived slip factor from power

Copyright 2008 by Concepts ETI, Inc.

power(W)

180 160 140 120 100 80 60 40 20 0 -20 0

test data model w /o recirculation model w . recirculation inlet recirc. pow er loss outlet recirc. pow er loss

Figure 6A shows another impeller with a radial Inlet. Figure 6B and 6C shows a comparison of power and efficiency. In this case, inlet recirculation is zero, because there is little pressure gradient from tip to hub. On the other hand, outlet recirculation rises steadily with reduced flow rate. Note the large difference in efficiency with and without the recirculation model. Figures 6D and 6E show the CFD results at the flow rate of 0.0005 m3/s. Figure 6D is the velocity vector on the meridional plane. This figure indicates that the flow is quite clean at the leading edge of the impeller, confirming our assumption that the inlet recirculation effect can be ignored for an impeller with a radial inlet. At the outlet is a recirculation zone near the shroud. On the blade-to-blade plane near the half-span (Figure 6E), a large recirculation zone is visible close to the suction surface, which is consistent with the velocity profile assumed in the current model (Figure 4B).
0.003

0.001

0.002

flow rate (cm s)

FIGURE 5C. RECIRCULATION AND POWER

Figure 7A-E shows another impeller with a mixed flow inlet. Both inlet and outlet recirculation are present in this case. The CFD results shown are at the flow rate of 0.1 m3/s. Overall, the model prediction of power and efficiency compares favorably with the test data.

0.7 0.6 efficiency (T-T) 0.5 0.4 0.3 0.2 0.1 0 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 flow rate (cms) test data model w/o recirculation model w. recirculation

FIGURE 6A. A RADIAL PUMP

FIGURE 5D. RECIRCULATION EFFECT ON EFFICIENCY

70 60 50 power(W) 40 30 20 10 0 0

test data model w /o recirculation model w . recirculation inlet recirc. pow er loss outlet recirc. pow er loss

0.0005

0.001

0.0015

flow rate (cm s)

FIGURE 5E. MERIDIONAL VELOCITY CONTOUR AT IMPELLER INLET AND EXIT

FIGURE 6B. RECIRCULATION LOSS AND POWER

Copyright 2008 by Concepts ETI, Inc.

0.8 0.7 0.6 efficiency (T-T) 0.5 0.4 0.3 0.2 0.1 0 0 0.0005 0.001 0.0015 flow rate (cm s) test data model w /o recirculation model w . recirculation

FIGURE 7A. A PUMP WITH MIXED FLOW INLET FIGURE 6C. RECIRCULATION EFFECT ON EFFICIENCY

90 80 70 power(kW) 60 50 40 30 20 10 0 0 0.1

test data model w /o recirculation model w . recirculation inlet recirc. pow er loss outlet recirc. pow er loss

0.2

0.3

0.4

0.5

FIGURE 6D. VELOCITY VECTOR ON MERIDIONAL PLANE

flow rate (cm s)

FIGURE 7B. RECIRCULATION LOSS AND POWER CONSUMPTION

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2

efficiency (T-T)

test data model w /o recirculation model w . recirculation

0.3

0.4

0.5

flow rate (cm s)

FIGURE 6E. VELOCITY VECTOR ON BLADE-TO-BLADE PLANE

FIGURE 7C. RECIRCULATION EFFECT ON EFFICIENCY

Copyright 2008 by Concepts ETI, Inc.

ACKNOWLEDGEMENTS
The authors want to thank John Keeling for help running the CFD cases and Scott Penny for his help in editing this paper.

REFERENCES
[1] Stepanoff, A. J. 1967. Centrifugal and Axial Flow Pumps, John Wiley & Sons, New York. [2] Neumann, B. 1991. The Interaction between Geometry and Performance of Centrifugal Pump, Mechanical Engineering Publication, London. [3] Tuzson, J. J. 1983. Inlet Recirculation in Centrifugal Pumps, Symposium on Performance Characteristics of Hydraulic Turbines and Pumps, ASME. [4] Aungier, R. H. 2000. Centrifugal Compressors, ASME Press, New York. [5] Jansen W. & Sunderland P. W. 1990. Off-design Performance Prediction of Centrifugal Pumps, Fluid Machinery Components, Fed-Vol.101 [6] Japikse D. Pelton R., Maynes D. and Oliphant K. N., 2005. Turbomachinery Performance Models, IMECE 2005-79414. [7] Qiu X., Mallikarachichi C., Anderson M. 2007. A New Slip Factor Model for Axial and Radial Impellers, ASME Paper No. GT2007-27064. [8] Fraser, W. H. 1981. Recirculation in Centrifugal Pumps, Symposium on Materials of Construction of Fluid Machinery and Their Relationship to Design and Performance, ASME. [9] Japikse D., Marscher W. D. & Furst R. B. 2006. Centrifugal Pump Design and Performance, Concepts NREC, White River Junction, Vermont [10] Japikse D. 1984. A Critical Evaluation of Stall Concepts for Centrifugal Compressors and Pumps Studies in Component Performance, Part 7. Stability, Stall and Surge in Compressors and Pumps. ASME Winter Annual Meeting, New Orleans, LA. [11] Del Valle J., Braisted D. M. and Brennen C. E. 1992. The Effects of Inlet Flow Modification on Cavitating Inducer Performance, J. of Turbomachinery, Vol. 114, No. 2, ASME. [12] Tanaka, T. An Experimental Study of Backflow Phenomena in a High Specific Speed Propeller Pump, ASME 80-FE-, 1980. [13] Kmmer N. and Rautenberg M., 1986. A Distinction Between Different Types of Stall in a Centrifugal Compressor Stage, ASME Paper No. 82-GT-82. [14] Shiavello, B. & Sen, M. 1980. On the Prediction of Reverse Flow Onset at the Centrifugal Pump Inlet, Symposium on Performance Prediction of Centrifugal Pumps and Compressors, ASME. [15] Brennen, C. E. 1994. Hydrodynamics of Pumps, Concepts ETI, Norwich, VT. [16] Yedidiah, S. 1996. Centrifugal Pump Users Guidebook, Chapman & Hall. [17] Johnson J. P., 1986. Radial Flow Turbomachinery, Lecture in series Fluid Dynamics of Turbomachinery. ASME Turbomachinery Institute, Ames, Iowa. [18] Cumpsty N. A., 1989. Compressor Aerodynamics, Addison Wesley Longman, England. [19] Johnson M. W. and Moore J. 1983. Secondary Flow Mixing Losses in a Centrifugal Impeller, Trans. ASME Journal of Engineering for Power 100:553-60. [20] Carpenter S. H. 1957. Performance of Cavitating Axial Inducers with Varying Tip Clearance and Solidity, Thesis, California Institute of Technology, Pasadena, CA.

FIGURE 7D. VELOCITY VECTOR ON MERIDIONAL PLANE

FIGURE 7E. VELOCITY VECTOR ON BLADE-TO-BLADE PLANE

CONCLUSIONS
A meanline recirculation model is proposed. This model offers a practical method for estimating the additional power consumption as well as the effect on the head rise due to the flow recirculation. At the inlet, a recirculation zone is derived from the assumption that the inlet pressure recovery of the active flow remains constant after it passes a critical stall point. This is made possible by an enlarging recirculation zone at the reduced flow rate. The inlet recirculation not only consumes additional power, its blockage effect also reduces the active flow incidence and results in the continued rise of head curve to the shut-off condition in axial impellers. At the outlet, the recirculation zone is primarily the result of blade loading and subsequent flow separation on the blade suction surface. The size of the exit recirculation zone is estimated by requiring a positive velocity on the pressure surface. The exit recirculation flow also has the effect of reducing the exit blade loading and increasing the slip factor, which also contributes to the rise of the head curve at the reduced flow rate. The validation study suggests that the current model may be applied to axial, mixed flow, or radial impellers. In axial impellers, inlet recirculation dominates, while in pure radial impellers, recirculation occurs mostly at the impeller exit.

Copyright 2008 by Concepts ETI, Inc.

S-ar putea să vă placă și