Skip to main content
Log in

Experimental analysis and modeling of the losses in the tip leakage flow of an isolated, non-rotating blade setup

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

Improving pressure rise capabilities of axial compressors requires an in-depth understanding of the losses produced in the tip leakage region. Here, a generic setup that magnifies the tip region of an isolated, non-rotating blade is used with the objectives of describing the main flow components and evaluating the related sources of loss. The flow at the tip is structured by the jet flow out of the gap which, under the effect of the main stream, rolls-up into a tip-leakage vortex. The current setup is characterized by the tip gap height and the thickness of the incoming boundary layer at the casing, here a flat plate, for a given incidence of the blade. Measurements are performed using LDV and a multi-port pressure probe. Variations in the tip-leakage flow are found to be mainly driven by gap height. A small, intermediate and large gap regimes are more specifically found, with threshold around 4% and 8% of gap to chord ratio for the present setting. The incoming boundary layer thickness is shown to provoke a notable effect on the vortex lateral position and total pressure losses. The local entropy creation rate is computed from LDV data and used to identify the sources of loss in the flow. A decomposition into wake and vortex losses is further proposed, allowing to relate the contributions of the various flow components to the overall losses. An empirical model of the formation of the tip vortex is developed to account for the increased losses as a function of gap height. The model provides a useful mean for the practical approximation of the gap sensitivity of pressure losses.

Graphic abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24

Similar content being viewed by others

Abbreviations

\(\alpha \) :

Angle of attack

c :

Blade chord

h :

Gap height

\(\delta _{0.99}\) :

\(0.99U_{{\max }}\) thickness

\(\delta ^*\) :

Displacement thickness

\(\theta ^*\) :

Momentum thickness

\(H^*=\delta ^*/\theta ^*\) :

Shape factor

\({\hbox {Re}}=U_\infty c /\nu \) :

Chord Reynolds number

\({\hbox {Re}}_{\theta ^*}=U_\infty \theta ^* /\nu \) :

\(\theta ^*\) Reynolds number

M :

Mach number

\(\Gamma \) :

Tip-leakage vortex circulation

\(\Gamma ^{\prime \prime }\) :

Total secondary vorticity

\((y_v,z_v)\) :

Vortex center position

\(C_\mathrm{{Pt}}=(P_{\mathrm{{t}}\infty }-P_\mathrm{{t}})/(\frac{1}{2}\rho U_\infty ^2)\) :

Total pressure loss coeff.

\(C_\mathrm{{wake}}=1-u_x^2/U_\infty ^2\) :

Wake loss coeff.

\(C_\mathrm{{vortex}}=C_\mathrm{{Pt}}-C_\mathrm{{wake}}\) :

Vortex loss coeff.

\(K_\mathrm{{P}}=(P_\mathrm{{s}}-P_{\mathrm{{s}}\infty })/(\frac{1}{2}\rho U_\infty ^2)\) :

Pressure coeff.

\(C_\mathrm{{L}}\) :

Airfoil lift coeff.

\(\chi _{\mathrm{{D}}}\) :

Discharge coeff.

\(U_\infty \) :

Upstream velocity

\(U_\mathrm{{j}}\) :

Gap exit jet velocity

P :

Pressure

T :

Temperature

s :

Entropy per unit of mass

\(\rho \) :

Density

\(\mu \) :

Dynamic viscosity

\(\nu \) :

Kinematic viscosity

\(u_i\) :

Mean velocity

\(u_i'\) :

Fluctuating velocity

\(S_{ij}=\frac{1}{2}\big (\frac{\partial u_i}{\partial x_j}+\frac{\partial u_j}{\partial x_i}\big )\) :

Strain rate tensor

\(\tau _{ij}=-\langle u_i'u_j' \rangle\) :

Reynolds stress tensor

\(k=\langle u_i'u_i' \rangle\) :

Turbulent kinetic energy

\(\bullet ^\mathrm{{m}}\) :

Mass-flow average

\(\bullet ^\mathrm{{s}}\) :

Surface average

\(\bullet _\infty \) :

Upstream quantity

\(\bullet _\mathrm{{t}}\) :

Total quantity

\(\bullet _\mathrm{{s}}\) :

Static quantity

\(\langle \bullet \rangle\) :

Ensemble average

\(\bullet _+\) :

Pressure side quantity

\(\bullet _-\) :

Suction side quantity

\(\nabla \bullet \) :

Nabla operator

LDV:

Laser Doppler velocimetry

References

  • Barker SJ, Crow SC (1977) The motion of two-dimensional vortex pairs in a ground effect. J Fluid Mech 82(4):659–671

    Article  Google Scholar 

  • Bloxham M, Bons J, Hollis R (2008) Horseshoe vortex control with leading edge endwall boundary layer removal. In: 4th Flow control conference, p 4319

  • Borello D, Hanjalic K, Rispoli F (2007) Computation of tip-leakage flow in a linear compressor cascade with a second-moment turbulence closure. Int J Heat Fluid Flow 28(4):587–601

    Article  Google Scholar 

  • Brandt H, Fottner L, Saathoff H, Stark U (2002) Effects of the inlet flow conditions on the tip clearance flow of an isolated compressor rotor. In: ASME Turbo Expo 2002: power for land, sea, and air. American Society of Mechanical Engineers, pp 1123–1132

  • Chassaing P (2010) Mécanique des Fluides : Éléments d’un premier parcours, 3rd edn. Cépaduès, Toulouse

    Google Scholar 

  • Decaix J, Balarac G, Dreyer M, Farhat M, Münch C (2015) Rans and les computations of the tip-leakage vortex for different gap widths. J Turbul 16:309–341. https://doi.org/10.1080/14685248.2014.984068

    Article  MATH  Google Scholar 

  • Denton J (1993) Loss mechanisms in turbomachines. In: ASME 1993 International gas turbine and aeroengine congress and exposition. American Society of Mechanical Engineers, pp V002T14A001–V002T14A001

  • Devenport WJ, Simpson RL (1990) Time-dependent and time-averaged turbulence structure near the nose of a wing-body junction. J Fluid Mech 210:23–55

    Article  Google Scholar 

  • Doukelis A, Mathioudakis K, Papailiou K (1998a) The effect of tip clearance gap size and wall rotation on the performance of a high-speed annular compressor cascade. In: ASME 1998 International gas turbine and aeroengine congress and exhibition. American Society of Mechanical Engineers

  • Doukelis A, Mathioudakis K, Papailiou K (1998b) Investigation of the 3-d flow structure in a high-speed annular compressor cascade for tip clearance effects. In: ASME 1998 international gas turbine and aeroengine congress and exhibition. American Society of Mechanical Engineers, pp V001T01A012–V001T01A012

  • Fearn R, Weston RP (1974) Vorticity associated with a jet in a cross flow. AIAA J 12(12):1666–1671

    Article  Google Scholar 

  • Flachsbart O (1931) Spaltverluste an Tragflügeln. ZAMM J Appl Math Mech 11(6):411–414

    Article  Google Scholar 

  • Hasan M, Casarella M, Rood E (1986) An experimental study of the flow and wall pressure field around a wing-body junction. J Vib Acoust Stress Reliab Des 108(3):308–314

    Article  Google Scholar 

  • Heyes F, Hodson H, Dailey G (1992) The effect of blade tip geometry on the tip leakage flow in axial turbine cascades. J Turbomach 114(3):643–651

    Article  Google Scholar 

  • Hu H, Yang Z (2008) An experimental study of the laminar flow separation on a low-Reynolds-number airfoil. J Fluids Eng 130(5):051101

    Article  Google Scholar 

  • Inoue M, Kuroumaru M (1989) Structure of tip clearance flow in an isolated axial compressor rotor. J Turbomach 111(3):250–256

    Article  Google Scholar 

  • Jacob M, Grilliat J, Camussi R, Caputi Gennaro G (2010) Aeroacoustic investigation of a single airfoil tip leakage flow. Int J Aeroacoust 9:253–272

    Article  Google Scholar 

  • Jacquin L (1994) Phenomenological description and simplified modelling of the vortex wake issuing from a jet in a cross flow. La Recherche Aérospatiale 2:117–133

    Google Scholar 

  • Julien HL, Kays W, Moffat R (1969) The turbulent boundary layer on a porous plate: an experimental study of the effects of a favorable pressure gradient. Technical Report HMT-4, Thermosciences Division, Department of Mechanical Engineering, Stanford University

  • Kameier F, Neise W (1997) Experimental study of tip clearance losses and noise in axial turbomachines and their reduction. J Turbomach 119(3):460–471

    Article  Google Scholar 

  • Kang S, Hirsch C (1993) Experimental study on the three-dimensional flow within a compressor cascade with tip clearance: part Iâvelocity and pressure fields. J Turbomach 115(3):435–443

    Article  Google Scholar 

  • Kang S, Hirsch C (1996) Numerical simulation of three-dimensional viscous flow in a linear compressor cascade with tip clearance. J Turbomach 118(3):492–502

    Article  Google Scholar 

  • Karagozian AR (2014) The jet in crossflow. Phys Fluids 26(10):101303

    Article  Google Scholar 

  • Lakshminarayana B, Horlock JH (1963) Tip-clearance flow and losses for an isolated compressor blade. Technical Report ARC-R/M-3316, Aeronautical Research Council London (England)

  • Menter FR (1994) Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J 32(8):1598–1605

    Article  Google Scholar 

  • Millikan CB (2018) Aerodynamics of the airplane. Courier Dover Publications, Mineola

    Google Scholar 

  • Moore J, Tilton J (1988) Tip leakage flow in a linear turbine cascade. J Turbomach 110(1):18–26

    Article  Google Scholar 

  • Muthanna C, Devenport W (2004) Wake of a compressor cascade with tip gap, part 1: mean flow and turbulence structure. AIAA J 42(11):2320–2331

    Article  Google Scholar 

  • Pope SB (2011) Turbulent flows. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Praisner T, Smith C (2006) The dynamics of the horseshoe vortex and associated endwall heat transfer. Part I: Temporal behavior. J Turbomach 128(4):747–754

    Article  Google Scholar 

  • Prandtl L (1919) Tragflügeltheorie. II. Mitteilung. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse, pp 107–137

  • Rains DA (1954) Tip clearance flow in axial compressors and pumps. PhD thesis, California Institute of Technology

  • Saathoff H (2001) Rotor-Spaltströmungen in Axialverdichtern. PhD thesis, Technische Universität Braunschweig

  • Saathoff H, Stark U (2000) Tip clearance flow induced endwall boundary layer separation in a single-stage axial-flow low-speed compressor, 2000-GT, p 0501

  • Smith G, Cumpsty NA (1982) Flow phenomena in compressor casing treatment. Technical Report, CUED/A-Turbo/TR 112, Whittle Laboratory, Department of Engineering, University of Cambridge

  • Storer J, Cumpsty N (1991) Tip leakage flow in axial compressors. J Turbomach 113(2):252–259

    Article  Google Scholar 

  • Storer J, Cumpsty N (1994) An approximate analysis and prediction method for tip clearance loss in axial compressors. J Turbomach 116(4):648–656

    Article  Google Scholar 

  • Vo HD, Tan CS, Greitzer EM (2008) Criteria for spike initiated rotating stall. J Turbomach 130(1):011023

    Article  Google Scholar 

  • Wu JZ, Ma HY, Zhou MD (2007) Vorticity and vortex dynamics. Springer, Berlin

    Google Scholar 

  • Yakhot A, Anor T, Liu H, Nikitin N (2006) Direct numerical simulation of turbulent flow around a wall-mounted cube: spatio-temporal evolution of large-scale vortices. J Fluid Mech 566:1–9

    Article  Google Scholar 

Download references

Acknowledgements

The present work is supported by the ANR project NumERICCS (ANR-15-CE06-0009), led by Pr Georges Gerolymos. The 2D RANS calculation has been performed in the framework of the elsA three-party agreement between AIRBUS, SAFRAN and ONERA which are co-owners of this software. The authors would like to thank Pr Jacques Borée and Pr Georges Gerolymos who have shared their expertise on theoretical aspects regarding the entropy production in the flow.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to B. Deveaux.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1: Entropy creation rate for the mean flow

The first objective of this section is to clarify the definition of the entropy production rate for the mean flow. The second objective is to demonstrate that this entropy production rate can be estimated using velocity measurements from LDV. This work generalizes the analysis carried out by Denton (1993) on the relation between entropy increment and viscous forces. The following assumptions are considered :

  1. 1.

    The fluid is a perfect gas: \(P=\rho R T\)

  2. 2.

    The flow is in a steady state and incompressible: \({\text {div}}(\mathbf {u})=0\) and \(M\ll 1\) (for the present case \(M=0 .1\))

  3. 3.

    Variations of total pressure and temperature are small, i.e., \(|P_\mathrm{{t}}-P_{\mathrm{{t}}\infty }|\ll P_{\mathrm{{t}}\infty }\) and \(|T_\mathrm{{t}}-T_{\mathrm{{t}}\infty }|\ll T_{\mathrm{{t}}\infty }\) (in the present experiment \(|P_\mathrm{{t}}-P_{\mathrm{{t}}\infty }|< 10^{-2}\times P_{\mathrm{{t}}\infty }\))

  4. 4.

    Heat transfers are neglected.

Starting with Gibbs–Duhem relation, one can relate entropy variation to total enthalpy and pressure for the instantaneous field following

$$\begin{aligned} T_\mathrm{{t}} {\hbox {d}}s = {\hbox {d}}h_\mathrm{{t}} - R T_\mathrm{{t}}\frac{{\hbox {d}}P_\mathrm{{t}}}{P_\mathrm{{t}}}. \end{aligned}$$
(28)

Assuming that \(|P_\mathrm{{t}}-P_{\mathrm{{t}}\infty }|\ll P_{\mathrm{{t}}\infty }\) and \(|T_\mathrm{{t}}-T_{\mathrm{{t}}\infty }|\ll T_{\mathrm{{t}}\infty }\) we have

$$\begin{aligned} T_{\mathrm{{t}}\infty } {\hbox {d}}s = {\hbox {d}}h_\mathrm{{t}} - \underbrace{\frac{R T_{\mathrm{{t}}\infty }}{P_{\mathrm{{t}}\infty }}}_{1/\rho _{\mathrm{{t}}\infty }}{\hbox {d}}P_\mathrm{{t}}. \end{aligned}$$
(29)

Insofar as the Mach number is small (\(M=0.1\)), we assume that \(\rho \approx \rho _{\mathrm{{t}}\infty }\), which leads to

$$\begin{aligned} T_{\mathrm{{t}}\infty }\frac{{\hbox {d}} s}{\mathrm{{d}} t} = \frac{{\hbox {d}} h_t}{\mathrm{{d}} t} - \frac{1}{\rho } \frac{{\hbox {d}} P_\mathrm{{t}}}{\mathrm{{d}} t} \quad {\text {where}} \quad \frac{{\hbox {d}} \bullet }{\mathrm{{d}} t}=\dot{\bullet }=\frac{\partial \bullet }{\partial t} + u_i \frac{\partial \bullet }{\partial x_i}. \end{aligned}$$
(30)

Then, the Reynolds decomposition is applied to Eq. (30), considering the assumptions of a steady and incompressible flow, which leads to

$$\begin{aligned} T_{\mathrm{{t}}\infty }\,\dot{s}_\mathrm{{m}} =\langle u_i \rangle \frac{\partial }{\partial x_i} \langle h_t \rangle - \frac{1}{\rho } \langle u_i \rangle \frac{\partial }{\partial x_i} \langle P_\mathrm{{t}} \rangle , \end{aligned}$$
(31)

with \(\langle \bullet \rangle\) the Reynolds average and with \(\dot{s}_\mathrm{{m}}\) defined as

$$\begin{aligned} \begin{aligned} \dot{s}_\mathrm{{m}} =& \langle u_i \rangle \frac{\partial }{\partial x_i} \langle s \rangle +\frac{\partial }{\partial x_i} \langle u_i's' \rangle \\&- \frac{1}{T_{\mathrm{{t}}\infty }}\frac{\partial }{\partial x_i} \langle u_i'e' \rangle . \end{aligned} \end{aligned}$$
(32)

In (33), \(e'\) represents the fluctuation of internal energy per unit of mass. This quantity \(\dot{s}_\mathrm{{m}}\) can be understood as the equivalent rate of entropy production for the mean flow, which is not the same as the mean entropy production rate \(\dot{s}\). From now on, the symbol \(<\bullet>\) is dropped for mean flow quantities in order to reduce the amount of notation. Equation (31) thus becomes

$$\begin{aligned} T_{\mathrm{{t}}\infty }\,\dot{s}_\mathrm{{m}} = u_i\frac{\partial h_t}{\partial x_i} - \frac{1}{\rho }u_i\frac{\partial P_\mathrm{{t}}}{\partial x_i}. \end{aligned}$$
(33)

To replace the right-hand side of Eq. (33), let us consider the RANS energy equation

$$\begin{aligned} \begin{aligned} u_j \frac{\partial h_t}{\partial x_j}=&\left( 2 \nu S_{ij} + \tau _{ij}\right) \frac{\partial u_i}{\partial x_j} + u_i\frac{\partial }{\partial x_j}\left( 2 \nu S_{ij} + \tau _{ij}\right) \\&- \frac{1}{\rho }\frac{\partial q_j}{\partial x_j}, \end{aligned} \end{aligned}$$
(34)

and the RANS mean flow kinetic energy equation

$$\begin{aligned} \frac{1}{\rho } u_i\frac{\partial P_\mathrm{{t}}}{\partial x_i} = u_i\frac{\partial }{\partial x_j}\left( 2\nu S_{ij} + \tau _{ij} \right) , \end{aligned}$$
(35)

with \(S_{ij}\) the strain rate tensor, \(\tau _{ij}=-<u_i'u_j'>\) the Reynolds stress tensor and \(q_j\) the heat flux. Inserting relations (34) and (35) in Eq. (33) leads to

$$\begin{aligned} T_{\mathrm{{t}}\infty }\dot{s}_\mathrm{{m}} = \left( 2 \nu S_{ij} + \tau _{ij}\right) \frac{\partial u_i}{\partial x_j} - \frac{1}{\rho }\frac{\partial q_j}{\partial x_j}, \end{aligned}$$
(36)

which is similar to the relation obtained by Chassaing (2010). This relation can also be written as follows

$$\begin{aligned} T_{\mathrm{{t}}\infty }\dot{s}_\mathrm{{m}} = 2 \nu S_{ij}S_{ij} + P_\mathrm{{k}} - \frac{1}{\rho }\frac{\partial q_j}{\partial x_j}, \end{aligned}$$
(37)

with

$$\begin{aligned} P_\mathrm{{k}}=\tau _{ij}\frac{\partial u_i}{\partial x_j}, \end{aligned}$$
(38)

the production of turbulent kinetic energy. Equation (37) shows that the equivalent mean flow entropy production rate \(\dot{s}_\mathrm{{m}}\) corresponds to the sum of viscous dissipation (\(2\nu S_{ij} S_{ij}\)), production of turbulent kinetic energy (\(P_\mathrm{{k}}\)) and heat power (\({\text {div}}(\mathbf {q})\)). Upon assuming that the heat power is negligible, Eq. (37) yields

$$\begin{aligned} \dot{s}_\mathrm{{m}} = \frac{1}{T_{\mathrm{{t}}\infty }}\left( 2 \nu S_{ij}S_{ij} + P_\mathrm{{k}}\right) . \end{aligned}$$
(39)

This relation only depends on mean flow velocity gradients and on the Reynolds stress tensor; therefore, it is possible to compute \(\dot{s}_\mathrm{{m}}\) from LDV measurements.

Appendix 2: Averaged total pressure losses and entropy creation rate

The objective of this section is to demonstrate the equivalence between mass-flow average total pressure losses \(C_\mathrm{{Pt}}^\mathrm{{m}}\) and the volume integrated rate of entropy production \(\rho \dot{s}_\mathrm{{m}}\) in a streamtube. The start is relation (33), which, taking into account that the flow is incompressible, leads to

$$\begin{aligned} T_{\mathrm{{t}}\infty } \rho \dot{s}_\mathrm{{m}}= {\text {div}}(\mathbf {u}. \rho h_t) - {\text {div}}(\mathbf {u}.P_\mathrm{{t}}). \end{aligned}$$
(40)

This relation is integrated over the streamtube \(\Sigma \) composed of an inlet boundary \(S_\mathrm{{in}}\), an outlet boundary \(S_{0}\) and a lateral boundary \(S_{lat}\). The inlet boundary is taken upstream of the blade where flow and thermodynamic conditions are homogeneous. The lateral boundary \(S_{lat}\) is characterized by \(\mathbf {u}.\mathbf {n}=0\), where \(\mathbf {n}\) is the local normal vector. Then, the divergence theorem applied to the integral of (40) yields

$$\begin{aligned} \begin{aligned} T_{\mathrm{{t}}\infty } \iiint _\Sigma \rho \dot{s}_\mathrm{{m}} \, {\hbox {d}}V&= \iint _{_{S_\mathrm{{in}}}} P_{\mathrm{{t}}\infty }U_\infty (\mathbf {e}_x.\mathbf {n}) {\hbox {d}}S\\&\quad - \iint _{_{S_{0}}} P_\mathrm{{t}} (\mathbf {u}.\mathbf {n}) {\hbox {d}}S + \Delta H_t , \end{aligned} \end{aligned}$$
(41)

with

$$\begin{aligned} \Delta H_t=\iint _{_{S_{0}}} \rho h_{t} (\mathbf {u}.\mathbf {n}) \,{\hbox {d}}S -\iint _{_{S_\mathrm{{in}}}} \rho h_{\mathrm{{t}}\infty }U_\infty (\mathbf {e}_x.\mathbf {n}) \,{\hbox {d}}S. \end{aligned}$$
(42)

In the present study, we consider a fixed setup, i.e., no mechanical power is exchanged. Therefore, assuming no heat transfer at the boundaries of the streamtube \(\Sigma \), we have \(\Delta H_t=0\). Moreover, mass-flow being conserved between \(S_\mathrm{{in}}\) and \(S_{0}\), Eq. (41) becomes

$$\begin{aligned} T_{\mathrm{{t}}\infty } \iiint _\Sigma \rho \dot{s}_\mathrm{{m}} \, {\hbox {d}}V = \iint _{_{S_{0}}} (P_{\mathrm{{t}}\infty }-P_\mathrm{{t}})(\mathbf {u}.\mathbf {n}) {\hbox {d}}S . \end{aligned}$$
(43)

The mass-flow averaged total pressure losses \(C_\mathrm{{Pt}}^\mathrm{{m}}\) is then introduced in Eq. (43), with

$$\begin{aligned} C_\mathrm{{Pt}}^\mathrm{{m}}= \frac{1}{\dot{m}}\iint _{S_{_{0}}} C_{P_\mathrm{{t}}}\,\rho \, (\mathbf {u}.\mathbf {n}) \,{\hbox {d}}S, \end{aligned}$$
(44)

where \(\dot{m}\) represents the mass flow through the streamtube \(\Sigma \), and with

$$\begin{aligned} C_\mathrm{{Pt}}=\frac{P_{\mathrm{{t}}\infty }-P_\mathrm{{t}}}{\frac{1}{2}\rho U_\infty ^2}, \end{aligned}$$
(45)

leading to

$$\begin{aligned} \iiint _\Sigma \rho \dot{s}_\mathrm{{m}} \, {\hbox {d}}V = \frac{1}{2} \frac{U_\infty ^2}{T_{\mathrm{{t}}\infty }} \,\dot{m} \,C_\mathrm{{Pt}}^\mathrm{{m}}. \end{aligned}$$
(46)

The average entropy increase per unit of mass \(\mathcal {L}\), defined as

$$\begin{aligned} \mathcal {L}=\frac{1}{\dot{m}}\iiint _{\Sigma } \rho \dot{s}_\mathrm{{m}}\,{\hbox {d}}V \quad , \end{aligned}$$
(47)

is introduced in relation (46), which gives

$$\begin{aligned} \mathcal {L}= \frac{1}{2}\frac{U_\infty ^2}{T_{\mathrm{{t}}\infty }} C_\mathrm{{Pt}}^\mathrm{{m}}. \end{aligned}$$
(48)

Relation (48) states the equivalence between the mass-flow averaged total pressure losses and the average entropy increase per unit of mass.

Appendix 3: Steady axisymmetric vortex analysis

This section establishes a relation between the vortex loss coefficient \(C_\mathrm{{vortex}}\), introduced in Eq. (13), and the circulation of the tip-leakage vortex. In a cylindrical reference frame \((x,r,\theta )\), \(C_\mathrm{{vortex}}\) is defined by

$$\begin{aligned} C_\mathrm{{vortex}}=\frac{P_{\mathrm{{s}}\infty }-P_\mathrm{{s}}}{\frac{1}{2}\rho U_\infty ^2}-\frac{u_\theta ^2}{U_\infty ^2} \end{aligned}$$
(49)

The radial velocity \(u_r\) is taken equal to zero. An axisymmetric steady inviscid flow is in radial equilibrium, that is

$$\begin{aligned} \frac{\partial P_\mathrm{{s}}}{\partial r} = \rho \frac{u_\theta ^2}{r} \end{aligned}$$
(50)

Upon introducing this relation in (49) leads to

$$\begin{aligned} C_\mathrm{{vortex}}= \frac{2}{U_\infty ^2} \int _r^\infty \frac{u_\theta ^2}{\xi } {\hbox {d}}\xi -\frac{u_\theta ^2}{U_\infty ^2} \quad . \end{aligned}$$
(51)

Let us now consider the Rankine vortex, defined as

$$\begin{aligned} u_\theta (r) = {\left\{ \begin{array}{ll} \frac{\Gamma }{2\pi a}\left( \frac{r}{a}\right) , \quad r<a \\ \\ \frac{\Gamma }{2\pi r},\quad r\ge a \end{array}\right. }, \end{aligned}$$
(52)

where a is the vortex core radius and \(\Gamma \) is the vortex circulation. Based on this model, the coefficient \(C_\mathrm{{vortex}}\) becomes

$$\begin{aligned} C_\mathrm{{vortex}}= {\left\{ \begin{array}{ll}\frac{2}{U_\infty ^2} \left( \frac{\Gamma }{2\pi a}\right) ^2\left[ 1-\left( \frac{r}{a}\right) ^2\right] , \quad r<a \\ ~ \\ 0,\quad r\ge a\end{array}\right. }. \end{aligned}$$
(53)

which shows that the \(C_\mathrm{{vortex}}\) coefficient is maximum at the vortex center and is equal to zero outside the vortex core.

The surface average of \(C_\mathrm{{vortex}}\) over a disk of radius R is defined as

$$\begin{aligned} C_\mathrm{{vortex}}^\mathrm{{s}}= \frac{1}{\pi R^2} \int _0^R C_\mathrm{{vortex}}(r)\cdot 2\pi r dr \end{aligned}$$
(54)

If the integration area includes the whole vortex core (\(R>a\)), one has

$$\begin{aligned} C_\mathrm{{vortex}}^\mathrm{{s}}= \frac{A}{S_0}\frac{\Gamma ^2}{U_\infty ^2}, \; S_0=\pi R^2, \; A=\frac{1}{4\pi },\; R > a \end{aligned}$$
(55)

Equation (55) shows that the surface average of \(C_\mathrm{{vortex}}\) is proportional to the vortex circulation squared, divided by the integration area. This relation does not depend on the vortex core radius a. If \(S_0\) and \(\Gamma \) are normalized by the velocity \(U_\infty \) and a reference length, relation (55) eventually becomes

$$\begin{aligned} C_\mathrm{{vortex}}^\mathrm{{s}}= \frac{A}{S_0}\Gamma ^{2}. \end{aligned}$$
(56)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Deveaux, B., Fournis, C., Brion, V. et al. Experimental analysis and modeling of the losses in the tip leakage flow of an isolated, non-rotating blade setup. Exp Fluids 61, 126 (2020). https://doi.org/10.1007/s00348-020-02957-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-020-02957-z

Navigation