You are on page 1of 10

published at: ETC7 – 7th Conference on Turbomachinery, Athens, March 2007, Paper No.

110

BRUSH SEAL POROSITY MODELING –


APPLICABILITY AND LIMITATIONS

M. Neef – F. Hepermann – N. Sürken – J. Schettel


Steam Turbine Technology, Siemens Power Generation, Mülheim/Germany,
matthias.neef@siemens.com

ABSTRACT
The present paper discusses three related approaches to efficient brush seal CFD modeling
by using a porous medium approach. The corresponding sets of resistance coefficients are
reviewed and resulting characteristics essential to steam turbine applications are evaluated.
The important radial clearance between bristle pack and rotor is used to calibrate the
numerical model according to experimental data. The influence of a changing clearance
height is studied with respect to the percentage of flow leaving the bristle pack.
As already shown, the porous medium approach proves to be suitable for brush seal
modeling without the necessity to resolve the flow around individual bristles. While previous
works have focused on leakage and the simulation of pressure drops in the porous region to
provide a basis to calculate contact forces, the current porous model is employed to study the
influence of the brush on swirl across the blade shroud of a steam turbine.
It is shown by comparison to standard swirl breakers that the brush effectively reduces
swirl and thus contributes to rotor-dynamic stability.

NOMENCLATURE
a, A, A viscous resistance factor or matrix v flow velocity
Acyl bristle or cylinder cross section V total volume
bbrush axial width of bristle pack Vfluid fluid volume fraction
b, B, B inertial resistance factor or matrix Vsolid solid volume fraction
C porosity factor used in Table 1 Vcyl bristle or cylinder volume
d bristle or cylinder diameter z, r, θ axial, radial, circumferential coord.
~
d equivalent sphere diameter α viscous resistance coefficient
dh hydraulic diameter β inertial resistance coefficient
D porosity factor used in Table 1 ∆p pressure difference
L length in direction of flow ε porosity
Lpipe, Lbristle pipe or bristle length ϕ bristle lay angle
n bristle packing density λL pipe friction coefficient
N number of bristles or cylinders µ dynamic viscosity
Re Reynolds number ρ density
Subscripts
n, z normal to bristles s in bristle direction

INTRODUCTION
Brush seals are widely used in the turbomachinery industry today to obtain leakage reductions
and improved efficiency (e.g. Neef et al., 2006). The theoretical modeling of brush seal behavior,
however, is still lagging behind. A governing approach to predict brush seal performance is
currently not available. The contemporary open literature on CFD modeling of brush seals may be
distinguished by the approach to simulate the bristle pack in the flow path. One method is to treat
the bristle pack as a porous medium which simulates the pressure drop across the bristles and
further effects of the brush on the fluid flow. A second method is to calculate the complete fluid
flow field around the individually packed bristles.

Modeling the brush as porous medium


Two dimensional approaches have to rely purely on calibration with experimental data for each
individual application and operating condition. Transferability is therefore limited. Dogu (2005), for
instance, who has recently discussed a large number of publications available on brush seal
modeling, calibrates the porosity factors in a 2D-CFD-model to achieve consistent results with
earlier publications. By nature of the approach, the model cannot represent the anisotropy of the
bristle pack. Different porosity factors are suggested for plate supported upper part and free
standing lower part of the brush in order to better match experimental results.
Chew et al. (1995, 1997) consequentially propose a 3-D, non-uniform description of the
porosity of the bristle pack and then use the pressure distribution within the porous region to derive
pressure forces which should be experienced by the bristles. Integration of these forces leads to an
estimate of blow down forces, bristle bending and consequently the influence on the rotor
underneath the bristle pack. In some calculations a significant clearance between the bristle pack
and the rotor is considered, which is then treated to be completely closed by the blow down during
operation. Extension of the model and further comparison with experimental results is presented by
Turner et al. (1997) and Chen et al. (1998). Bearing in mind the sensible simplifications in this
approach, the method is very useful and highlights the applicability of brush seal porosity modeling
for predicting brush seal performance. Along the same line, Pröstler (2005) makes use of a 3D-
model to present an optimized brush design which influences the blow-down effect.

Modeling the brush region as group of individual bristles


Two dimensional flow simulations with resolution of individual bristles, i.e. bristle cross
sections, have been published. But since they cannot capture the complicated flow through the
anisotropic bristle pack in all directions, they are primarily of academic value. Studies by Braun et
al. (1990) result in good agreement with findings from experimental measurements and flow field
visualizations, but are unable to incorporate details of the fluid-flow in a bristle wise direction. A
radial pressure distribution, which is responsible for the blow-down effect (Ferguson, 1988), cannot
be captured.
Three dimensional CFD simulations of packs of bristles theoretically offer the opportunity to
include all physical effects such as the flow around individual bristles. However, they require either
large model sizes and computational power or are reduced to only a few rows of individual bristles
which do not represent the complete seal. Full resolution of individual bristles becomes interesting
when coupled with mechanical models to allow fluid structure interaction, see for example Lelli et
al. (2005). In an iterative process, the effects of the fluid on the mechanical behavior of single
bristles are calculated. This model also considers the blow-down effect and leads to reasonable
agreement with experimental results.

POROSITY MODELING OF BRUSH SEALS

General equations
Early investigations of flows through porous media have been performed by Darcy (1856) and
Carman (1937). The influence of friction in such flow is adequately reflected by equation (eq. 1),
which takes into account the friction of the cylinder surface and the dissipation of the flow energy
due to obstacles in the flow path:
∆p
= A ⋅ µ ⋅ v + B ⋅ ρ ⋅ v2 (eq. 1)
L
The pressure drop ∆p occurs across the length L of the packed bed of cylinders in the direction
of flow. The coefficients A and B in the viscous and the inertial resistance terms, respectively, are
defined by Ergun (1952) as follows:
α ⋅ (1 − ε )
~ 2
A= ~ (eq. 2)
ε3 ⋅ d 2
~
β ⋅ (1 − ε )
B= ~ (eq. 3)
ε3 ⋅ d
~
Variable d is the equivalent sphere diameter that is related to the cylinder or bristle diameter d by
~ 6 × cylinder volume
d= = 1.5 d (eq. 4)
cylinder surface
neglecting the cross-sectional area for long bristles. ε represents the porosity of the medium, which
is defined by the quotient of the fluid volume Vfluid divided by the total volume V enclosing the
porous region including the solid bristle volume Vsolid:
V V
ε = fluid = 1 − solid (eq. 5)
V V
Based on measurements, Ergun suggests values for the empirical resistance coefficients ~ α and
~ ~ ~
β as α =150 and β = 1.75 for spherical particles. For bristles or packed cylinders, making use of
~
the equivalent sphere diameter in eq. 4 the coefficients are α = ~ α / d² ⋅ d² = 66.7 and
~ ~
β = β / d ⋅ d = 1.17 . For a packed bed of cylinders, others have derived the constants α = 80 and β =
1.16, see for example Pröstler (2005) and Chew and Hogg. (1997).

Application to brush seals


The porosity of a bristle pack can be calculated from the geometrical parameters as the packing
density n (bristles per circumferential seal length), the lay angle of the bristles ϕ (see Figure 1), the
bristle diameter d and the axial width of the brush bbrush:
π ⋅ d2 ⋅ n
ε = 1− (eq. 6)
4 ⋅ b brush ⋅ cosφ
The axial width of the brush varies with the pressure difference ∆p over the bristle pack, i.e. the
porosity itself is a function of pressure drop, especially for low pressure drops. According to
Pröstler (2005), the porosity is constant for pressure differentials above 1.5-2 bars. In his
investigation, the minimum packing width - which can be achieved in reality only at higher pressure
differentials - exceeds the theoretical value for the closest packing of bristles. This leads to an
increase in actual porosity of about 28% compared to the porosity as defined in eq. 6. The increase
is taken into account for the porosity values in the current investigation.
as, bs - „bristlewise loss“
az, bz
„transverse loss“
an, bn bristle pack cross section
d/cosϕ
ϕ
r d
© Siemens
Lbristle bbrush
z
bbrush
θ
Figure 1: Schematic sketch of the bristle pack
When calculating the fluid flow through the bristles, it is important to regard the anisotropic
structure of the brush as illustrated in Figure 1. The flow resistance in the direction of the bristles
(bristlewise loss) is much smaller than orthogonal to the bristles (transverse loss) since the flow is
channeled in the bristle direction. This can be taken into account in numerical modeling, when the
flow is treated in a three-dimensional domain. The resistance coefficients then have different values
for different directions.
For this purpose the general approach in (eq. 1) can be written in the form of vectors and tensors
in the following equation - see also Chew et al. (1995). It thus represents an extra force on the fluid
flow exerted by the bristle pack which can be included in the conservation equations:
⎡a n 0 0 ⎤ ⎡b n 0 0 ⎤
Fr = A ⋅ µ ⋅ v + B ⋅ ρ ⋅ v ⋅ v with A = ⎢ 0 a s 0 ⎥ and B = ⎢⎢ 0 b s 0 ⎥⎥ (eq. 7)
⎢ ⎥
⎢⎣ 0 0 a z ⎥⎦ ⎢⎣ 0 0 b z ⎥⎦
Here, the vector Fr is the resistance force on the fluid per unit volume. In this equation the
matrices A and B contain the coefficients for viscous resistance (friction) and for inertial
resistance (dissipation). The coefficients are most conveniently defined in a coordinate system
aligned to the bristles, i.e. in directions n, s and z (see Figure 1). In this case, only the diagonal
values of A and B , an, as, az as well as bn, bs, and bz need to be defined. By transformation, along
with the bristle angle and their orientation in the fluid domain, these can be converted into the
global coordinate system r, θ, z - see again Chew et al. (1995). However, most commercial CFD
codes feature a formulation in an arbitrary, user-defined coordinate system such as the bristle-fixed
frame of reference, which allows to input the coefficients directly.
The coefficients an and bn (orthogonal to the bristles) are expressed analogously to the terms A
and B defined by Ergun (eq. 2 and eq. 3) as
α ⋅ (1 − ε ) β ⋅ (1 − ε )
2
an = az = and bn = bz = . (eq. 8)
ε ⋅d
3 2
ε3 ⋅ d
For α and β the coefficients of Ergun or Pröstler can be used as given above.
Apart from Ergun’s approach to treat a region of spheres or cylinders as a homogeneous porous
region, the anisotropic approach now defines the coefficient for flow in bristle direction as a
separate value. Here, two formulations suggested in recent publications shall be compared.
Based on comparison between experimental values and numerical results, Chew et al. (1995)
assume a ratio of 1:60 for the friction loss in the bristle direction relative to the loss in the
transverse direction. Along with the viscous friction coefficient α = 80 this leads to:
80 ⋅ (1 − ε ) (1 − ε)
2 2
1 1
as = ⋅an = ⋅az = = 1.33 ⋅ 3 2 and bs = 0 (eq. 9)
60 60 60 ⋅ ε 3 ⋅ d 2 ε ⋅d
The dissipative loss is neglected by setting bs to zero – in the bristle direction, there will be no
wakes behind obstacles if perfectly aligned bristles are assumed.
Rather than assuming an empirical ratio between bristlewise and transverse friction losses, the
fluid flow in the bristle direction can also be regarded as a flow in a pipe to obtain a value for as
(Pröstler, 2005). The pressure drop of laminar flow in a pipe with the hydraulic diameter dh and the
length Lpipe is then defined by:
∆p 1 ρ 2
− = λL ⋅ ⋅ ⋅v (eq. 10)
L pipe dh 2
In this equation, λL represents the friction coefficient within the pipe. The hydraulic diameter of
such a densely packed bed of cylinders (bristles) is defined by the quotient of the free volume for
the fluid Vfluid and the adjacent cylinder cross sections Acyl as follows:
Vfluid 4 ⋅ Vfluid V
dh = = with N = solid (eq. 11)
N ⋅ A cyl N ⋅ π ⋅ d 2
Vcyl
N gives the number of cylinders and is defined by the quotient of the blocked volume divided by
the volume of a single cylinder. Along with the definition of the porosity (eq. 5) the hydraulic
diameter of the free fluid volume becomes:
d⋅ε
dh = (eq. 12)
(1 − ε )
By making use of Reynold’s number and the friction coefficient for laminar flow in pipes:
ρ ⋅ v ⋅ dh 64
Re = and λ L = (eq. 13)
µ Re
the pressure drop is calculated as:


∆p (1 - ε )2
= 32 ⋅ 2 2 ⋅ µ ⋅ v (eq. 14)
L bristle ε d
For the flow inside the bristle pack of a brush, the pipe length is represented by the bristle length
Lbristle. Comparison with equation (eq. 1) leads to the definition of the resistance coefficient in
bristle direction as resulting in:

a s = 32 ⋅
(1 - ε )2 (eq. 15)
ε 2d 2
Although this factor looks similar to the ones given above, it should be noted that the
dependency on the porosity ε has changed.
Friction /viscosity Dissipation /inertial The three different approaches for
resistance coefficient resistance coefficient porosity modeling of a brush seal
a = 66.7 C bn = 1.17 D are listed in Table 1 with respect to
Ergun n the bristle diameter d. It shows
as = an bs = bn
(1952) Ergun’s empirical constants without
az = an bz = bn
considering the anisotropy in
an = 80 C bn = 1.16 D
Chew permeability in contrast to Chew’s
a = 1/60 an = 1.33 C bs = 0
(1995) s and Pröstler’s method of modeling
az = an bz = bn
brush anisotropy. The latter both
a = 80 C bn = 1.16 D
Pröstler n refer to Kay & Neddermann (1974)
as = 32 ε C bs = 0
(2005) for the resistance coefficients in the
az = an bz = bn
transverse direction. Assuming a
1 ⎛1− ε ⎞
2
1 ⎛1− ε ⎞
with C = 3 ⋅ ⎜ ⎟ with D = 3 ⋅ ⎜ ⎟ typical brush seal porosity of 15%,
ε ⎝ d ⎠ ε ⎝ d ⎠ it can be seen that Ergun’s approach
~ will lead to the lowest leakage rates
with d = d /1.5 (see eq. 4)
Table 1: Comparison of the different resistance coefficients followed by Pröstler’s and finally
Chew’s approach. Their relative
and quantitative implications at defined operating conditions, though, need to be assessed
separately. This shall be pursued in the following section.

APPLICATION OF POROSITY MODELING TO A BRUSH SEAL


The above approaches are compared in a numerical model of a single stage brush seal using the
commercial code CFX 5. A 5-degree slice of the circumference is modeled in this 3-dimensional
approach. The modeled geometry with a brush and two succeeding labyrinth fins represents a
configuration which has been tested in an air test rig at the University of Munich (Urlichs, 2005) at
pressure differentials of between 0 and 10 bar yielding mass flow rates and pressures for
comparison with the numerical data. Given the double-flow nature of the rig and experience from
repeatability measurements, an accuracy of the leakage results of about ±7% can be expected.
The brush used in this experiment consists of bristles inclined at 45° with a diameter of 0.07
mm and a packing density of 200 bristles per millimeter circumferential seal length. These
parameters are used to calculate the porosity of the brush seal according to the approach described
in the section above. Since the brush is assembled with an initial radial clearance, this gives an
additional parameter which influences the leakage flow of the seal. Accordingly, the experimental
values are used to calibrate the numerical model of the brush by adjusting the modeled free gap
height underneath the bristle pack (Figure 2).
First of all, the most sophisticated porosity model suggested by Pröstler (2005) is employed. As
datum point for all succeeding model analyses, the leakage rate at 3.5 bar pressure drop is calibrated
to match the experimental result at that specific operating condition. The computationally necessary
free radial gap width is then found to be 0.092 mm. The calculation results used for calibration of
the model have been checked for reasonable grid-independency and are converged to a level of at
least 3 x 10-6 for averaged residuals for velocity and mass flow. During the experiments, detailed
optical measurements of the gap were taken with the rig being in operation (Urlichs, 2005). These
measurements show that, at maximum pressure differentials, the initial cold clearance of the seal is
closed down to 0.02 mm by the blow-down effect. Therefore, the simulated gap might be
considered to be significantly too large. However, it seems sensible to the authors not to vary other
parameters of the model in order to achieve the measured mass flow rate at said measured
clearance. Those parameters generally exhibit a relatively high uncertainty as for instance the actual
package width of the porous region.
0.12
0,12
CFD-Results Ergun In the current configuration, the
h = 0.092 mm
numerical model already achieves
0.10
0,10 CFD-Results Pröstler the desired accuracy and validity
CFD-Results Chew for a wide range of pressure
0,08
0.08 differentials. The anisotropic
measured values formulation of the porous region
mass flow rate [kg/s]

0,06
accounts for the bristle lay angle,
0.06
and the geometric boundary
conditions account for the existing
0.04
0,04
gap underneath the bristle pack
despite its exact physical value.
0.02
0,02 © Siemens Calculations have then been
∆p = 3.5bar extended to the application of the
0,00
0 other resistance coefficients
0
0,0 2,5 5
5,0 inlet pressure
7,5 [bar] 10
10,0 according to Ergun and Chew,
leaving all other boundary
Figure 2: Comparison of calculated and measured leakage flows
through the brush seal conditions including the free gap
size untouched. Results for the low
pressure differential range are not discussed here due to the stronger dependency of the porosity on
pressure drop which is judged to exceed the scope of this investigation.
As already expected by comparing the similar nature and values given in Table 1 for the three
models, the results do not vary too much, with highest mass flows for Chew’s approach due to
lowest resistance coefficients. It may be noted that the numerical models predict slopes of the curve
in good agreement with sealing theory – linear dependence on pressure with the curve intersecting
the origin of the coordinate system. For conventional fin seals, this is true for pressure drops above
the critical pressure ratio where sonic speed is reached in the gap. The experimental values do not
show this trend, since hysteresis effects in the pressure differential dependent blow down of the
brush seal have a strong influence on the leakage. Moreover, the slope of the measured leakage rate
related to the pressure differential is lower, which points to a noticeable dependency of the radial
clearance on pressure differential not taken into account in the current investigation.
Now, in order to assess the influence of the radial gap on the performance of the seal in more
detail, the flow through the brush seal is divided in two fractions, one flow underneath the bristle
pack and the flow which leaves the seal through the bristles (see Figure 3). For the calibration point
of the model (see Figure 2, long dashed line), the fractional mass flows leaving the bristles relative
to the total mass flow are compared in Figure 4 (left, dashed line). Due to the smaller bristlewise
resistance coefficient, Chew’s porosity values lead to a bristle mass flow of about 30% related to the
total flow for the modeled gap of 0.092 mm. The
isotropic approach of Ergun with the highest resistance
back coefficients leads to about 18% of the full mass flow
front plate through the bristle pack. For smaller values for the free
plate
gap underneath the bristle pack, the fraction of flow
through the bristle pack increases while the total
bristle pack bristle flow quantity of the mass flow through the seal decreases
gap height h gap leakage (Figure 4, right). When the gap is completely closed,
the differences in modeling the permeability become
most apparent. Less resistance in the bristle direction
rotating surface leads to higher leakages for the approaches of Pröstler
Figure 3: Brush seal schematic drawing –
and Chew.
bristle and gap mass flow
Since experimental evidence suggests smaller radial
gaps during operation (see above) than the one used for calibration in this calculation, actual
resistance factors might be lower than the ones given by the three discussed porosity models.
The dependence of mass flow on radial gap height (Figure 4) can also be used to introduce a more
accurate leakage model which takes into account the closure of the free gap for higher pressure
drops. This was already discussed for Figure 2, where experimental values show lowest flow rates
for higher pressures, where the blow down tends to close remaining gaps between bristles and rotor.
100 0,05
0.05
∆p = 3.5 bar Ergun ∆p = 3.5 bar
80 0,04
0.04
Pröstler
total mass flow [kg/s]

h = 0.092mm
bristle flow [%]

60 0,03
0.03
Chew
40 0,02
0.02
h = 0.092mm
20 0,01
0.01
© Siemens © Siemens
0 0
0 0.02 0,04
0,02 0.04 0.06 0,08
0,06 0.08 0.1
0,1 gap height h [mm] 0 0.02
0,02 0.04
0,04 0.06
0,06 0.08
0,08 0.1
0,1

Figure 4: Percentage of flow leaving the bristle pack (left) and total mass flow through the seal (right) as a
function of free gap height (performed at 3.5 bar pressure drop)

CFD-INVESTIGATION OF A SHROUD SEAL


While most CFD analyses of brush seals have focused on representing detailed features of the
seals such as back plate pressure distributions, an aim of the work presented here is to predict
overall flow features of a close-to-application seal configuration with a brush. Having said this,
many details of the bristle pack flow such as blow-down effects etc. do not necessarily have to be
resolved. Yet, flow turning characteristics to be tackled by permeability models are of key interest.
In the previous chapters it has been shown that the brush model proposed by Pröstler (2005)
captures the leakage flow quite well. Therefore, this model shall now be used to investigate a brush
in the blade path of a turbine and the influence of this brush on leakage flow swirl.

Model of the seal and components


A porosity brush model was set up for the leakage path of a shroud seal of the fourth stage in an
intermediate pressure steam turbine. A sketch of the model is shown in Figure 5. For computational
efficiency the tip of the turbine blade foil was also modeled as an anisotropic porous medium,
which produces the same pressure drop and influences flow direction and velocity in a way similar
to the actual turbine blade.
seal brush
casing leakage flow

blade
shroud outlet
main flow
r © Siemens
inlet blade foil
θ z
Figure 5: Computational blade shroud seal model

To compare the influence of the brush, the effect of a standard swirl breaker (see Fig. 6) is also
examined. The swirl (or circumferential velocity component) within the seal influences both,
turbine efficiency and rotordynamic stability. Efficiency is mainly affected when the leakage flow
mixes with the main flow. Differing circumferential velocities will lead to dissipation and
unfavorable changes in incidence to succeeding blade rows. The rotordynamic stability of the
turbine is influenced because of the swirl in the leakage flow. This swirl is a driver of the ‘steam
whirl effect’ that contributes to the excitation of rotor vibrations.

swirl breaker

seal fins z
θ © Siemens

© Siemens (short version, swirl breaker B in Fig. 9)

Figure 6: Labyrinth seal segment with standard swirl breaker

Swirl Analysis
Figure 7 draws a comparison between a labyrinth seal with regular fins (gap = 0.8 mm), a
labyrinth seal including a thick fin with a small gap representing the geometric dimensions of a
blocked bristle pack (gap = 0.1 mm), and a labyrinth seal including a brush seal modeled as a
porous medium (gap = 0.1 mm). The plot shows the average swirl component at distinct axial
locations corresponding to the geometric representation of the seal at the top of the graph.
Due to the porosity modeling, the swirl velocity in the third configuration is reduced
considerably from 180 m/s to a minimum of about 30 m/s. Due to wall friction at the rotor surface,
the average circumferential velocity subsequently increases to a value of 108 m/s at the outlet of the
seal. This is close to half the shroud surface speed of approx. 280 m/s. The two brushless labyrinth
seals are shown for comparison. The thick fin with a radial gap corresponding to the brush seal
demonstrates that the swirl reduction of the brush seal results from the porous bristle pack rather
than the small clearance as might be expected. The influence of the thick fin with a small clearance
on swirl is larger than that of the regular labyrinth, but much smaller than that of the brush seal.
Corresponding provisional measurements of the swirl component of the fluid behind a brush seal,
which shall not be discussed here in more detail, provide evidence that the level of swirl reduction
is reasonably captured by the numerical model.
250
without brush b=0.3mm
gap=0.8mm
circumferential velocity [m/s]

200

150
without brush b=2mm
gap=0.1mm
100

50 © Siemens brush

0
10 20 30 40 50 60
distance to seal entrance [mm]
Figure 7: Influence of the brush on the swirl

Figure 8 shows the function of different swirl breakers within the leakage path of the seal in
comparison as well as in conjunction with a brush seal. The circumferential velocity component is
most significantly reduced in the seal with a swirl breaker only, and by omitting a subsequent brush.
The long version A of the swirl breaker is slightly more effective since the ratio of the radial swirl
fin extension relative to the remaining gap is better than in the case of the short swirl breaker B. A
combination of swirl breaker and brush does not result in a difference for the swirl velocity level at
the outlet of the seal. This is an interesting result, since it has been previously suggested that swirl
breakers may be used in combination with brush seals in order to improve rotor stability. Here, it is
shown that the brush itself reduces the swirl and that the addition of a swirl breaker upstream of a
brush seal has no additional influence even for the chosen free gap under the bristle pack of 0.1
mm. While the swirl breaker only provides partial flow guidance to reduce axial velocity, the
porous bristle pack governs the seal flow by its high flow resistance and the corresponding total
pressure drop. Additionally, the brush pack acts like a flow straightener. Further investigation has
shown that increasing the gap beneath the bristle pack lessens the swirl breaking effect, but it
remains noticeable even for large radial clearances of up to 0.4 mm.
250
without brush
200
circumferential velocity [m/s]

150
brush

100
swirl breaker A

50

swirl breaker B
0 © Siemens

-50 swirl breaker A and brush


10 20 30 40 50 60
distance to seal entrance [mm]
Figure 8: Comparison of different swirl breakers with brush and regular fin seal
When extended to a complete 360° circular seal with eccentric rotor position, the model
discussed can be effectively used to assess rotor-dynamic coefficients with CFD as already known
for labyrinth seals (Schettel et al., 2005).

CONCLUSIONS
The present paper discusses different approaches to brush seal CFD modeling and shows how
experimental data can be used to calibrate a numerical model for a wide range of operating
parameters. Despite valuable results regarding leakage prediction of brushes, the authors clearly
show the limitations of the porous medium approach. This approach is not suitable to capture
effects related to movement of the bristles such as flutter, blow down or blow over (at high pressure
differentials). This challenges the leakage results in the low pressure differential range. The
approach is also limited when wear or the durability of brushes are studied (see Neef et al., 2006).
Although the treatment with the porous medium approach cannot represent the flexible nature of
the brush, it is suitable to capture general features such as leakage or the capability of the brush to
break inlet swirl of seal flows. This is especially important when the influence of a brush on rotor-
dynamic stability is to be evaluated. Since the anisotropic permeability models can be employed to
account for predominant flow paths in the bristle pack, the sensitivity to the swirl breaking
capability of the brush can be investigated. Finally, it is shown by comparison to standard swirl
breakers that a brush can effectively reduce swirl and can thus contribute to rotor-dynamic stability.

ACKNOWLEGEMENTS
The authors would like to thank Klaus Kwanka († 2006) and Martin Deckner from the Lehrstuhl
für Energiesysteme at the TU Munich for providing the experimental data used for calibration. The
work presented in this paper was supported by the Bundesministerium für Wirtschaft und Arbeit in
Germany under the support code 0327714D as part of the COOREF-T project. Responsibility for
the content of the publication rests with the authors.

REFERENCES
Braun, M., Hendricks, R.C.; Canacci, V.A.: Flow visualisation in a simulated brush seal. ASME-Paper 90-GT-217,
1990
Carman, P.: Fluid Flow through Granular Beds. Trans. Inst. Chem. Engineers Vol. 15, pp. 150-166, London, 1937
Chen, L; Wood, P.; Jones, T.; Chew, J.: An iterative CFD and mechanical brush seal model and comparison with
experimental results. ASME-Paper 98-GT-372, 1998
Chew, J.; Lapworth, B.; Millener, P: Mathematical modeling of brush seals. International Journal of Heat and Fluid
Flow, 1995, Vol. 16, pp. 493-500
Chew, J.; Hogg, S.: Porosity modeling of brush seals. Journal of Tribology. 119, Oct. 1997, pp. 769-77
Darcy, H: Les Fontaines publiques de la ville de Dijon. Paris, 1856 - in: Carman (1937)
Dogu, Y.: Investigation of brush seal flow characteristics using bulk porous medium approach. Journal of Engineering
for Gas Turbines and Power. January 2005, Vol. 127, pp. 136-144
Ergun, S.: Fluid Flow through Packed Columns. Chemical Eng. Progress Vol. 48, pp. 89-94, 1952
Ferguson, J.G.: Brushes as High Performance Gas Turbine Seals ASME-Paper 88-GT-182, 1988
Kay, J.M.; Neddermann, R.M.: An Introduction to Fluid Mechanics and Heat Transfer, Camb. Univ. Press, 3rd ed., 1974
Lelli, D.; Chew, J.; Cooper, P.: Combined 3D fluid dynamics and mechanical modeling of brush seals. ASME-Paper
GT2005-68973, 2005
Neef, M; Sürken, N.; Sulda, E.; Walkenhorst, J.: Design Features and Performance Details of Brush Seals for Turbine
Applications. ASME-Paper GT2006-90404, 2006
Pröstler, S.: Modellierung und numerische Berechnung von Wellenabdichtungen in Bürstenbauart. PhD Thesis
University of Bochum, 2005
Schettel, J.; Deckner, M.; Kwanka, K; Lüneburg, B; Nordmann, R.: Rotordynamic coefficients of labyseals for turbines
– comparing CFD results with experimental data on a comb-grooved labyrinth. ASME-Paper GT2005-68732, 2005
Turner, M.; Chew, J.; Long, C.: Experimental investigation and mathematical modeling of clearance brush seals.
ASME-Paper 97-GT-282, 1997
Urlichs, K.: Ermittlung der Eigenschaften adaptiver Dichtungen für hohen Druck und hohe Temperatur für die
Anwendung in Dampfturbinen, Abschlußbericht Vorhaben 414/00, Bayr. Forschungsstiftung, München, 2005

You might also like