Sie sind auf Seite 1von 13

R EYNOLDS N UMBER BASED B LADE

T IP VORTEX M ODEL
Manikandan Ramasamy∗ J. Gordon Leishman†

Alfred Gessow Rotorcraft Center


Department of Aerospace Engineering
Glenn L. Martin Institute of Technology
University of Maryland
College Park, Maryland 20742

Abstract Nb Number of blades


rc Core radius of the vortex, m
A mathematical model has been developed to estimate the r Radial distance, m
temporal growth properties of helicopter blade tip vortices r Non-dimensional radial distance, = r/rc
at any vortex Reynolds number. One unique feature of R Radius of the blade, m
the model is that it takes into account rotational stratifi- Rev Vortex Reynolds number, = Γv /ν
cation (Richardson’s number) effects on the distribution Ri Richardson number
of turbulent viscosity inside the tip vortices. This model t Time, s
is combined with another model for the effects of fila- T Rotor thrust, N
ment stretching in predicting the temporal evolution of V1 Peak swirl velocity, ms−1
the vortex. A turbulent growth model solves exactly for V1new Peak swirl velocity in new model, ms−1
the tangential (swirl) velocity starting from the Navier– Vr Radial velocity of the tip vortex, ms−1
Stokes equations using an assumed variation in eddy vis- Vz Axial velocity of the tip vortex, ms−1
cosity across the vortex core. This variation is a function Vθ Swirl velocity of the tip vortex, ms−1
of the local Richardson’s number, and the final solution αL Lamb’s constant, = 1.25643
becomes dependent on vortex Reynolds number. A more αI Iversen’s constant, = 0.01854
parsimonious functional approximation is given to repre- αnew New empirical constant, = 0.0655
sent the induced velocity distribution in the tip vortices for γ Reduced circulation, = rVθ , m2 s−1
practical applications. It is shown that the temporal core γv Reduced circulation at large distances, m2 s−1
growth rate predicted by the new model increases with an γ Non-dimensional circulation, = γ/γv
increase in vortex Reynolds number, which is consistent Γ Circulation, = 2πrVθ , m2 s−1
with experimental observations. The predictions from the Γv Circulation of the vortex at large distances, m2 s−1
model were validated, wherever possible, with tip vortex Γ1 Circulation at the core radius, m2 s−1
measurements from both model- and full-scale rotors. δ Ratio of apparent to actual viscosity
ζ Wake age, deg.
η Similarity variable, = r2 /4γvt
Nomenclature η1 Similarity variable at the core radius,= rc2 /4γvt
ηa Empirical constant
a, b Empirical constants
η Scaled similarity variable, = η/αnew 2
c Blade chord, m
κ Newly developed function, = αnew 2 VIF
C0 Constant
µ Dynamic viscosity, kgm−1 s−1
CT Coefficient of thrust, = T /ρAΩ2 R2
ν Kinematic viscosity, = µ/ρ, m2 s−1
g Core circulation function
νt Eddy viscosity
l Prandtl’s mixing length, m
νT Total kinematic viscosity, = ν + νt
∗ Research Associate. mani@glue.umd.edu ρ Density, kg/m3
† Minta Martin Professor. leishman@eng.umd.edu σ Shear stress, N/m2
Presented at the 61st Annual Forum and Technology Display of σe Effective rotor solidity, = Nb c/πR
the American Helicopter Society International, Grapevine, TX, ψ Azimuthal position, deg.
June 1–3, 2005. 2005
c by M. Ramasamy & J. G. Leishman. Ω Rotational speed of the rotor, rad/s
Published by the AHS International with permission.
Introduction wind tunnel models, and rotating-wing micro air vehicles
(MAVs) means that several orders of magnitude of differ-
Understanding the temporal development of helicopter ence in the tip vortex Reynolds numbers is involved. Until
rotor blade tip vortices has been the subject of intensive the Reynolds number issues are understood and modelled,
research for several decades. The motivation is clear, in airload predictions will remain unreliable.
that a more complete understanding of the structure of the The overall turbulence present inside the tip vortex has
tip vortices is essential for accurately predicting the un- been previously hypothesized to change with the geomet-
steady airloads on helicopter blades. It has been hypothe- ric scale of the rotor (Ref. 14). Consequently, this af-
sized by many investigators that the evolution of rotor tip fects the core growth, peak swirl velocity, and induced
vortices, such as the core growth and induced velocity dis- velocity distribution of the tip vortex. Bearing in mind
tribution, are directly related to the details of the turbulent that most vortex models are developed empirically from
flow structures present inside the tip vortex core (Refs. 1– measurements made on sub-scale laboratory size rotors,
5). A better understanding of these details is critical for this for that matter raises many questions about the appli-
helicopters because the tip vortices generated by one blade cability of these models to full-scale rotors or to MAVs.
can interact with following blades, resulting in a problem Certainly, full-scale rotor tip vortex measurements have
known as blade-vortex interaction (BVI). This BVI prob- been made in the past, but are very few in number and
lem results in high unsteady airloads, and is a source of cannot be made under the same controlled conditions that
significant noise and vibration levels on helicopter rotors. are possible in the laboratory or the wind tunnel. There
Because of the continued emphasis on reducing helicopter are no vortex measurements that have been made at MAV
noise and rotor vibration, higher-fidelity tip vortex mod- scale. Boatwright (Ref. 16) made measurements in the
els need to be developed to more accurately predict BVI. wake of a hovering, full-scale OH-23B rotor using hot-
Eventually, a deeper understanding of how vortices de- wire anemometry, while Cook (Ref. 17) made hot-wire
velop and the factors that influence their evolution should measurements using a full-scale S-58 rotor on a hover
help analysts to devise better strategies to alleviate the ad- tower. The paucity of full-scale rotor measurements has
verse effects associated with vortex induced airloads. This its roots not only from the substantial financial invest-
goal, however, is a longer way off. ments but also from the numerous practical difficulties in-
Most existing vortex models assume either a com- volved in making vortex flow measurements at this scale,
pletely laminar or turbulent interior flow. For example, including the required spatial and temporal fidelity.
the classic Lamb–Oseen model (Refs. 6, 7) assumes a As previously alluded to, the inherent difficulties in
completely laminar interior flow, while Squire (Ref. 1) making full-scale measurements in an already compli-
and Iversen (Ref. 2) assume completely turbulent flow for cated rotor flow field has caused rotor analysts to perform
their models. Even though there are measurements to sup- tip vortex experiments mostly on sub-scale model rotors
port both laminar and turbulent vortex flow assumptions (e.g., Refs. 11, 12, 18–20). Vortex models that have been
(Refs. 8, 9), new improvements in measurement instru- developed based on these sub-scale measurements have
mentation have allowed for better clarity into the details of been used to try to explain the persistence of rotor tip vor-
the flows inside rotor tip vortices. Flow visualization stud- tices to the relatively old wakes ages that are observed in
ies (Refs. 10, 11) and high-resolution flow measurements full-scale rotor tests, but with limited success. This failure
performed on model scale helicopter rotors (Refs. 10, 12, can be attributed, in part, to the neglect of Reynolds num-
13) have confirmed an original hypothesis made by Tung ber scaling issues while developing a tip vortex model.
et al. (Ref. 14) that the tip vortex can be classified into This can be explained using Fig. 1, which includes
three distinct flow regions: an inner laminar region free of Cook’s full-scale rotor tip vortex measurements (Ref. 17),
all turbulence, a transitional region with eddies of small and measurements by Martin et al. (Ref. 10), Ma-
scale, and an outer turbulent region with larger eddies. halingam et al. (Ref. 18), Ramasamy & Leishman
The extent of the three regions, however, depends on sev- (Ref. 11), McAlister (Ref. 12), and the 40% full-scale
eral factors, including the vortex Reynolds number. HART II test measurements (Ref. 21). For a helicopter
Semi-empirical vortex models have been developed in rotor, the tip vortex Reynolds number is given approxi-
the past that have recognized such a multi-region vortex mately by the result
structure, i.e., Tung model (Ref. 14) and Hoffman &  
2ΩRc CT
Joubert model (Ref. 15). However, all the aforementioned Rev = (1)
ν σ
models have been limited in their application to helicopter
rotor problems because they are not general enough to so that matching tip speed ΩR and blade loading coeffi-
be applied for any vortex Reynolds number, i.e., they do cient CT /σ leaves only a linear dependence on geometric
not account for scaling issues. Today, the size scales in- rotor scaling. From the results in Fig. 1, it is apparent that
volved between full-scale helicopter rotors, laboratory or the vortex Reynolds number of the model-scale experi-
v
Vortex Reynolds number, Re
10
8
Full scale Core Growth Theory
7
10
An important aspect of predicting vortex evolution is pre-
Model scale
10
6 dicting the temporal growth rate of the vortex core. It is
convenient to quantify the development of the vortex in
5 Ramasamy & Leishman, 2004
10 terms of its core size because the peak swirl velocities are
Martin & Leishman, 2003
4 McAlister, 2003 obtained at the core boundary. It is widely accepted that
10
Micro-air Cook, 1972 the growth of the tip vortex core depends on the nature
Mahalingam et al., 1998
1000 vehicles of the flow inside the tip vortices, i.e., whether it is lam-
HART II test
inar or turbulent and to what extent. For example, a tur-
100 bulent flow state increases mixing and so the transfer of
0 0.2 0.4 0.6 0.8 1 1.2
momentum across the layers of the vortex. This causes the
Ratio rotor radius (Model scale/Full scale)
core to grow as its vorticity spreads radially away from the
Figure 1: Comparison of the vortex Reynolds number core axis. In a laminar flow, momentum transfer is possi-
for sub-scale and full-scale rotor measurements. ble only by molecular diffusion. A schematic explaining
ments is lower by orders of magnitude when compared the effect of diffusion on the growth properties of the tip
with the full-scale tests, mainly because of geometric scal- vortices is shown in Fig. 2. The vortex core size increases
ing issues. In the case of the HART II tests, which are with increasing in time and the core vorticity decreases
40% of full-scale, the measured circulation of the tip vor- (but total circulation is conserved).
tices seems relatively lower than the expected value (the Lamb (Ref. 6) and Oseen (Ref. 7) assumed the flow
Rev values are closer to 105 than to 106 , for reasons that inside the vortex to be completely laminar and derived
are not yet clear). an exact solution to the one-dimensional, incompressible
Another important, but most often neglected, issue is Navier–Stokes equations. The core growth with time pre-
the effect of vortex filament strain on the growth prop- dicted by the Lamb–Oseen model (Refs. 6, 7) is given
erties of tip vortices. Helicopter rotor vortices develop by

in a highly three-dimensional and nonuniform velocity rc = 4αL νt (2)
field, which causes the vortex filaments to undergo ei-
ther a stretching or contraction process as they convect where αL is Lamb’s constant (αL = 1.25643). This result,
in the flow. Stretching of the vortex filaments with posi- however, suggests a growth rate that is substantially lower
tive velocity gradients can intensify the core vorticity and than found with experimental measurements – see Fig. 3.
increase swirl velocities; contracting the vortices in a neg- Also, the Lamb–Oseen model approaches a singularity at
ative strain field produces the opposite effect. Therefore, time t = 0 with infinite kinetic energy, which is not phys-
neglecting this strain process can alter significantly BVI ically realistic.
noise and rotor vibration predictions. The effects of vor- Squire (Ref. 1) and Bhagwat & Leishman (Ref. 23)
tex filament strain on the development of rotor tip vortices modified the Lamb–Oseen model by including an eddy
have been shown significant within the context of free- viscosity for turbulence that was present inside the tip vor-
vortex rotor wake predictions (Ref. 22). Γv
A blade tip vortex model combining the effects of dif- Swirl velocity
fusion and strain in predicting vortex evolution was pro-
posed by Ananthan et al. (Ref. 22), and was validated
in an experiment by Ramasamy & Leishman (Ref. 11).
A Reynolds number dependent transitional vortex model
that takes into account the effects of rotational stratifica- ω Filament undergoes
tion effects (or Richardson number effects) on the tur- viscous diffusion Γv
bulence present inside the vortex was also developed by
Ramasamy et al. (Ref. 13), and was validated using avail-
able measurements from various sources. The present
work involves combining and extending these two vor-
tex models to develop a comprehensive vortex model that ω
takes into account both the filament straining issues and
the rotational stratification (Richardson number) effects Swirl velocity
on the effective turbulent viscosity within the vortex in-
terior. The final model is dependent on vortex Reynolds Figure 2: Schematic explaining the physics of a vortex
number. filament undergoing diffusion.
tex. The modified core growth is given by McAlister, 2003
 Martin et al., 2001
Lamb–Oseen model
rc = r02 + 4ανδt (3) 0.25 Squire model, δ = 2
Squire model, δ = 8

Non-dimensional core radius, rc/ c


Squire model, δ = 16
where r0 is the initial core radius that removes the sin- 0.2 Ramasamy & Leishman, 2004
δ = 16
gularity at t = 0, and δ is the ratio of apparent to actual Cook, 1972

viscosity, i.e., δ=8


0.15
ν + νt νt
δ= = 1+ (4)
ν ν 0.1
δ=2
where νt is the effective turbulent value of viscosity. The δ = 1 (Lamb-Oseen)
0.05
values of r0 and δ (or νt ) must be obtained from mea-
surements. Squire assumed that the eddy viscosity, which
0
results from turbulence, is a function of the kinematic vis- 0 180 360 540 720 900
cosity, however, with a different magnitude. Because the Wake age, ζ (deg)
principal permanent characteristic of a tip vortex is its cir-
culation, Squire assumed that the eddy viscosity was pro- Figure 3: Vortex core growth predicted by Squire’s
portional to the total vortex circulation, i.e., model (In this case the results that ζ0 ≈ 30◦ ).
γ   
v Γv
δ = 1+a = 1 + a1 (5)
ν ν
where a and a1 are empirical constants and the ratio Γv /ν
is the vortex Reynolds number. 1
For very low vortex Reynolds numbers, the value of δ
approaches 1, so it reduces to the laminar Lamb–Oseen 2
model. Higher values of δ correspond to an increased
level of turbulence inside the vortex. This would result 3
in increased vortex core growth rate, as shown in Fig. 3.
It should, however, be noted that δ is not a function of r,
which means that the turbulence present inside the tip vor-
tex is implied to be independent of radial location. Even
though, the Squire vortex model differs from the laminar Region 1: Fully laminar
Region 2: Transitional
Lamb–Oseen model by including (on average) the effects Region 3: Fully turbulent
of turbulence on the core growth properties, the swirl ve-
locity distribution predicted by both models are the same. Figure 4: A representative flow visualization image
This result is also independent of vortex Reynolds num- of a tip vortex emanating from a rotor blade showing
ber or other scaling. Yet this is not consistent with mea- three distinct flow regions: (1) a fully laminar region,
surements and this deficiency with the modeling was one (2) a transitionally turbulent region, and (3) a fully tur-
motivation for the present work. bulent region.
Estimating the value of a1 (and, hence, δ) from Eq. 5
largely depends upon the way in which the eddy viscos- rotor noise. This point has been addressed previously by
ity is assumed to vary radially across the entire tip vor- Ramasamy & Leishman (Ref. 13).
tex region from its core axis into the outer far-field region. There are also observations from both velocity field
For example, a1 approaches zero for laminar flow assump- measurements (Refs. 10, 13–15) and flow visualization
tions, such as in Lamb–Oseen model. While Squire as- (Refs. 10, 13) suggesting that flow structures in the vor-
sumed uniform eddy viscosity, Iversen (Ref. 24) hypoth- tex are not consistent with either the fully laminar or fully
esized that the eddy viscosity varies linearly with the ra- turbulent assumption. It is now known that a rotor tip vor-
dial distance. This different assumption affects the core tex is made of three main regions: an inner laminar re-
growth. Bhagwat & Leishman (Ref. 23) suggested that gion that rotates like a solid body, a transitional flow re-
the average value of a1 lies within the fairly broad range gion, and a more fully turbulent outer region. One such
from 0.0004 to 0.00005 based on a summary of all avail- example is shown in Fig. 4. This means the eddy viscos-
able measurements. However, determining a more exact ity is nearly zero near the vortex core axis and approaches
value for a1 is required for applications that must pre- a value equivalent to the fully turbulent region far away
dict accurately rotor airloads, helicopter vibrations, and from the center of the tip vortex. Between these regions is
Γv
a transitional region with eddies of different length scales.
Ramasamy & Leishman (Ref. 13) developed a transi- Swirl velocity
tional vortex model that is a function of vortex Reynolds
number based on the premise of this multi-region vortex
concept. The predictions of the model were later validated ω
l
with measurements, with good agreement. S Γv

Filament is strained
or "stretched"
Methodology
A vortex Reynolds number dependent core growth model l + ∆l
for rotor tip vortices can be developed that takes into ac-
S
count the effects of vortex filament strain and flow rotation ω
effects on turbulence (or the eddy viscosity) present inside Swirl velocity
the tip vortices. This is obtained by combining two ex-
perimentally validated individual vortex models (a strain Figure 5: Schematic explaining the physics of a vortex
model and a transitional flow model). filament undergoing positive filament strain.

measured in a known strain field (Ref. 11). The measure-


Vortex Strain Model ments were found to correlate well with the core growth
predicted by the strain model.
A vortex strain model was developed by Ananthan et al.
(Ref. 22), which quantifies the effects of vortex filament
stretching on the core properties of rotor tip vortices. Con- Transitional Vortex Model
servation of fluid mass and momentum were used in de-
riving this model. A filament undergoing pure diffusion A transitional rotor tip vortex model was developed by
results in increased core size per unit length (for the same Ramasamy & Leishman (Ref. 13) and takes into account
filament circulation) with increase in time, as shown pre- the swirling flow (rotation) effects on the turbulent struc-
viously in Fig. 3. However, the strain model suggests that ture of tip vortices. This model was developed using
a tip vortex filament must result in a reduced core size and an eddy viscosity function in such a way that the func-
an increased core vorticity when it is subjected to a pos- tion smoothly and continuously models the eddy viscosity
itive stretching in a flow with positive velocity gradients, variation across the vortex from its inner laminar region to
as shown by the schematic in Fig. 5. Conservation of mass the outer turbulent flow region.
and momentum leads to an expression for the vortex core An intermittency function was developed based on a
growth, which is given by Richardson number concept, which also brings in the ef-
 fects of swirling flow (rotation) on the turbulence present
 ζ inside the vortex boundaries. The Richardson number is
rc = r0 2 + 4ανδ (1 + ε)−1 dζ (6) defined as the ratio of turbulence produced or consumed
ζ0
as a result of centrifugal forces to the turbulence produced
by shear. Bradshaw (Ref. 25) derived an expression for
where r0 is the initial core radius, ζ is the (wake) age of
the Richardson number based on an analogy between ro-
the filament, and ε(ζ) is the instantaneous filament strain
tational and stratified flows. This expression, which was
as given by
later modified by Holzapfel (Ref. 26), is given by
∆l
ε= (7)    2
l 2Vθ ∂(Vθ r) ∂ (Vθ /r)
Ri = r (8)
Applying zero strain rate will reduce the strain model r2 ∂r ∂r
(Eq. 6) into a diffusion based Squire-like core growth
model, as given previously in Eq. 3. and involves the velocity gradients in the vortex flow.
To validate this model, measurements were made in the Cotel & Breidenthal (Ref. 27) and Cotel (Ref. 28)
wake of a hovering rotor in the presence of a ground plane. suggested that the tip vortex will not develop or sustain
This resulted in very high velocity gradients being pro- any turbulence until the local gradient Richardson num-
duced to strain the vortices. Flow visualization of the tip ber, Ri, falls below a critical value (or stratification thresh-
vortex developments allowed the strain field to be mea- old). Based on experiments, Cotel et al. (Ref. 29) deter-
sured. These high velocity gradients strained the vortex mined that the critical value of Richardson number to be
1/4
filaments allowing the core properties of tip vortices to be Ri = Rev . This would mean that at any radial location
10 6 1
3

Vortex intermittency function, VIF


10 5 Lamb–Oseen model
Iversen's model
0.8
Richardson number, Ri

4
10 Measurements
Intermittency function
Stratification line
1000 0.6
100 2

10 0.4
1/4
Stratification line, Ri = Re
V
Core radius
1
0.2
0.1
1
0.01 0
0 0.5 1 1.5 2 2.5 1 2 3
Non-dimensional distance from core center, r / r Non-dimensional distance from core center, r / rc
c

Figure 6: Plot of Richardson number with radial coor- Figure 7: Eddy viscosity intermittency function across
dinate for a vortex flow. the vortex: (1) laminar flow region, (2) transitional
flow region, and (3) fully turbulent region.
the vortex will not be able to develop or sustain any turbu-
lence if the local value of Ri stays above the stratification based on Eq. 9. Far away from the vortex core axis, the
1/4
threshold of Rev . value of VIF approaches one, resulting in the value of
Figure 6 shows the variation of Ri for the previously eddy viscosity equivalent to the eddy viscosity for a com-
mentioned Squire and Iversen models along with the mea- pletely turbulent flow.
surements made by Ramasamy & Leishman (Ref. 11). It This expression for eddy viscosity was then incorpo-
is evident that there exists a multi-region vortex structure rated into the momentum equation governing the develop-
with laminar flow until a particular distance from the cen- ment of an axisymmetric vortex flow, i.e.,
ter of the vortex where the Richardson number is always     
above the threshold value. This is followed by a transition ∂γ ∂ ∂ γ ∂ γ
=r νT r + 2ν T (11)
flow region and then an outer turbulent region on moving ∂t ∂r ∂r r2 ∂r r2
far away from the vortex core axis. This concept is clearly
consistent with flow visualization (Fig. 4). This results in a similarity solution for the circulation dis-
Using this Richardson number concept, Ramasamy & tribution that is a function of vortex Reynolds number, i.e.,
Leishman (Ref. 13) developed a generalized eddy viscos-   2
∂γ ν 1 ∂ γ 2|X|X ∂
ity function to represent the variation of eddy viscosity − = + 4VIF · |X|
2
+ VIF
across the tip vortex. The expression for eddy viscosity, ∂η γ0 α2i ∂η2 η ∂η
(12)
which was derived using an analogy based on boundary
layer theory, is given mathematically by
where  
1 ∂γ
   X= η −γ
∂ Γ γ0 ∂η
νt = VIF αnew 2 r2 (9)
∂r r2 The sequence of steps involved in deriving the solution
for this equation, along with the various assumptions, is
where αnew is an empirical constant (found empirically). given in Ref. 13. It can be understood from Eq. 12 that
VIF is called the vortex intermittency function, which is the model reduces to the laminar Lamb–Oseen model or
defined by to a completely turbulent model for values of the VIF ap-


 proaching 0 and 1, respectively. Also, for very low vor-
1 η tex Reynolds numbers, the model approaches the laminar
VIF = 1 + erf b − ηa (10)
2 η1 Lamb–Oseen model for any value of the VIF. The three
empirical constants involved in the model were derived
where b and η a are empirical constants, η is the similarity using vortex flow measurements from various available
variable and η/η1 is equivalent to the ratio (r/rc )2 . sources.
The variation of the vortex intermittency function (VIF) Solving Eq. 12 numerically using a Runge-Kutta
with respect to the radial location of tip vortices is shown scheme showed that the circulation and induced veloc-
in Fig. 7. It can be observed that near the center of the ity distribution of tip vortices predicted by the transitional
vortex the VIF approaches zero. This results in zero eddy model correlated extremely well with experimental mea-
viscosity, as assumed by the laminar Lamb–Oseen model surements. Examples are shown in Figs. 8 and 9 in terms
Lamb–Oseen model
Iversen's model
1.2 Transitional model
Ramasamy & Leishman, 2004
1 3
v

2
Non-dimensional circulation, Γ / Γ

Martin & Leishman, 2000


McAlister, 2003
1 McAlister, 1996
Cook, 1972
0.8 Cliffone & Orloff, 1975
Jacob et al., 1996
Mahalingam & Komerath, 1998
0.6 Bhagwat & Leishman, 1998
Govindaraju & Saffman, 1971
Jacob et al., 1995
Measurements Kraft, 1955
0.4
Lamb–Oseen model McCormick, Tangler & Sherrib, 1963
Iversen's model 4 Rose & Dee, 1963
10 Corsiglia et al., 1973
0.2

Effective viscosity coefficient, δ


Transitional model
Baker et al., 1974
Tung model Dosanjh et al., 1964
1000
0
0 2 4 6 8 10 12
Non-dimensional distance from core center, r / r 100
c
Figure 8: Predicted ratio of circulation to circulation
10
at large distance, Rev = 48, 000, (1) laminar region, (2)
transitional region and, (3) turbulent region. 1
1.2
Non-dimensional swirl velocity, Vθ / V1

Model scale Full scale


0.1
1 Measurements 1000 10
4
10
5
10
6 7
10
Transitional model Vortex Reynolds number, Re
v
0.8 Lamb–Oseen model
Iversen's model Figure 10: Variation of δ with vortex Reynolds num-
0.6 ber based on new core growth model.

0.4
where αnew is a “new” empirical constant estimated based
0.2
1 2 3
on measurements from various available sources that were
0
listed in Ref. 13, and V1 is the peak swirl velocity pre-
0 1 2 3 4 5 6 7 8 dicted by the transitional model.
Non-dimensional distance from core center, r / r
c
The ratio δ predicted by the transitional vortex model is
Figure 9: Predicted swirl velocity distribution pre-
plotted against vortex Reynolds number in Fig. 10. It can
dicted by the transitional vortex model for Rev =
be observed that the Lamb–Oseen model predicts a con-
48, 000, (1) laminar region, (2) transitional region and,
stant core growth independent of Reynolds number (be-
(3) turbulent region.
cause of the inherent laminar flow assumption), while the
of the circulation distribution and the swirl velocity pro- Iversen model and the transitional model predicts an in-
files, respectively. The agreement of the new model with creased growth rate for an increase in vortex Reynolds
the measurements is clearly better than for either a model number. It should, however, be noted that the Iversen core
developed on laminar flow assumptions or a model based growth model shows a much higher core growth rate com-
on fully turbulent flow assumptions. pared with measurements or with the transitional vortex
model. This is because the Iversen model assumes that the
New Vortex Model eddy viscosity inside the vortex is fully turbulent from the
vortex core axis to the outer potential region. The value
The development of the new vortex model, which is ob- of δ predicted by the transitional model correlates well
tained by combining the strain and transitional models with measurements both in sub-scale experiments (where
previously discussed, will be complete by: (i) formulat- the vortex Reynolds number is lower) as well as full-scale
ing an expression for δ in Eq. 6 using the temporal growth tests (that have vortex Reynolds numbers at least an order
predictions from the transitional vortex model, and (ii) de- of magnitude higher).
riving a more convenient algebraic expression to represent
the swirl velocity distribution. Comparing the value of δ predicted by the transitional
The expression for δ based on the transitional vortex model with the value of δ based on the uniform eddy vis-
model is given by cosity model proposed by Squire (as given in Eq. 5) en-
ables the determination of a unique value for the constant
 
Rev α2new Γv Lamb 2 a1 , which is 6.5 × 10−5 – see Fig. 11. With this unique
δ= (13) value for a1 , substituting the expression for δ from Eq. 5
2παL V1new
5
10 1.2

1
Non-dimensional swirl velocity, V θ / V
Effective viscosity coefficent, δ
4
10 New model 1
-5
a = 6x10
1
1000 Lamb–Oseen model 0.8 All results
overlap
a1 here
100 0.6

10 0.4 Transitional Model


Lamb–Oseen model
Curve fit
1 0.2

0.1 0
4 6 8
1 100 10 10 10 0 0.5 1 1.5 2 2.5 3 3.5 4
Vortex Reynolds number, Re Non-dimensional distance from core center, r / r
v c
Figure 11: Comparison of δ between the transitional Figure 12: Numerical prediction of swirl velocity ver-
vortex model and Squire’s model to determine the con- sus expeonential approximation for Rev = 1 × 102 .
stant a1 . 1.2

1
Non-dimensional swirl velocity, V θ / V
into Eq. 6 results in the core growth model 1
 Transitional model
 ζ 0.8 and
curve fit
rc = r0 2 + 4αν(1 + a1 Rev ) (1 + ε)−1 dζ (14)
ζ0 0.6

Because this model is a function of vortex Reynolds num- 0.4 Transitional model
Lamb–Oseen model
ber, the core growth properties for a tip vortex that devel-
Curve fit
ops in time (i.e., with age) through any strain field ε(ζ) at 0.2

any vortex Reynolds number, Rev , can now be determined.


0
0 0.5 1 1.5 2 2.5 3 3.5 4
Non-dimensional distance from core center, r / r
Swirl Velocity Distribution c

Figure 13: Numerical prediction of swirl velocity ver-


The induced velocity profile as a function of the Reynolds sus expeonential approximation for Rev = 48 × 103 .
number can be obtained from the circulation distribution
are determined by fitting this expression to the numeri-
predicted by the vortex model, as given by Eq. 12. The
cally predicted velocity profiles. Because the swirl veloc-
difficulty in using this model arises from the relatively
ity profile changes with vortex Reynolds number, different
inconvenient mathematical form of the final result and
values of an , and bn are obtained as Rev varies. The curve-
the need for its numerical solution. The computational
fit was made in such a way that
cost involved in repeatedly solving a differential equa-
tion for the velocity profile is significant within the con- 3
text of routine velocity field evaluations in comprehensive ∑ an = 1 (16)
rotor codes. Therefore, a more approximate solution was n=1
sought. so as to satisfy the boundary condition that the swirl ve-
Algebraic vortex models have gained increasing popu- locity is zero at the vortex core axis. Care was also taken
larity over the past few years because of their extremely to make sure that the predicted swirl velocity based on the
low computational cost (such as for inclusion within free- curve fit asymptotes to zero at large values of r.
vortex wake models) and good fidelity when compared to The value of the constants that are obtained for var-
measurements, at least at a single Reynolds number, e.g., ious vortex Reynolds number are given in Table 1. It
the Vatistas vortex model (Refs. 30, 31). To have a rel- can be observed that at low vortex Reynolds numbers
atively parsimonious mathematical function to represent a1 = 1 and an = 0 for n = 1 and b1 = αL = 1.25643
the velocity profile, an expression of the form and bn = 0 for n = 1. This confirms that the expression

for estimating the swirl velocity distribution reduces to the
Γv 3
laminar Lamb–Oseen model for low values of Rev .
Vθ = 1 − ∑ an exp(−bn r) (15)
2πr n=1 Figures 12, 13 and 14 show the results for the swirl
velocity that were obtained at three different vortex
has been used to approximate the results given by the vor- Reynolds numbers. Using this parsimonious mathemat-
tex model in Eq. 12. Here, an and bn are constants that ical form in representing the induced velocity field of the
Rev a1 b1 a2 b2 b3
1 1.0000 1.256 0.0000 0.0000 0.0000
100 1.0000 1.2515 0.0000 0.0000 0.0000
1000 1.0000 1.2328 0.0000 0.0000 0.0000
10, 000 0.8247 1.2073 0.1753 0.0263 0.0000
2.5 × 104 0.5933 1.3480 0.2678 0.01870 0.2070
4.8 × 104 0.4602 1.3660 0.3800 0.01380 0.1674
7.5 × 104 0.3574 1.3995 0.4840 0.01300 0.1636
1 × 105 0.3021 1.4219 0.5448 0.0122 0.1624
2.5 × 105 0.1838 1.4563 0.6854 0.0083 0.1412
5.0 × 105 0.1386 1.4285 0.7432 0.0058 0.1144
7.5 × 105 0.1011 1.4462 0.7995 0.0048 0.1078
1 × 106 0.0792 1.4716 0.8352 0.0042 0.1077

Table 1: Values of the coefficients used in the curve fit to the numerical predictions of swirl velocity as a function of
vortex Reynolds number.
th 4
1.2 1/7 scale model, Re = 8X10
1

v
Non-dimensional swirl velocity, V θ / V

th 5
1/15 scale model, Re = 1.7X10
v
1 Lamb-Oseen model
5
S-58 Full-scale, Re = 8 X 10
v
0.8
Ramasamy & Leishman, 2004
McAlister, 1996
0.6 McAlister, 2003
Cook, 1972
0.4 Mahalingam et al., 1998
0.4
Non-dimensional core radius, rc / c

Transitional model
Lamb–Oseen model
Curve fit
0.2
0.3
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Non-dimensional distance from core center, r / r 0.2
c
Figure 14: Numerical prediction of swirl velocity ver-
sus expeonential approximation for Rev = 1 × 106 . 0.1

tip vortices is not only computationally inexpensive, but


also takes into account the required change in the form of 0
0 180 360 540 720
Vθ with Rev , which will be necessary for accurate aeroa-
Wake age, ζ (deg)
coustic predictions. Values of the coefficients for interme-
diate Reynolds numbers can be found through interpola- Figure 15: Core growth predicted by the new vortex
tion. model at different vortex Reynolds numbers.

Core Growth It can be observed that for full-scale flight conditions


the vortex core size grows up to 25% of the blade chord
The core growth predicted by the new model for a range within 360◦ of wake age but only up to 35% by two rotor
of vortex Reynolds numbers is shown in Fig. 15. This in- revolutions. Because the new vortex model has its basis
cludes vortex Reynolds numbers that correspond to both in the Iversen model, a logarithmic growth rate is found at
full- and model-scale rotors. Vortex measurements from older wake ages. Again, it is seen that the model reduces
various available sources at different vortex Reynolds to the laminar Lamb–Oseen model at very low values of
number are shown. It can be observed that the core size vortex Reynolds number (Rev < 1000). The new model
predicted by the new model correlates well with the mea- slightly overpredicts the tip vortex core growth measured
surements that are made at different vortex Reynolds num- by Cook (Ref. 17) using the full-scale Sikorsky S-58
bers. An increase in the vortex Reynolds number corre- rotor. This difference can, at least in part, be attributed to
sponds to an increase in the amount of turbulence present the measurement uncertainties involved in making a full-
inside the tip vortices and this, therefore, increases the scale measurement in the open air environment; for exam-
growth rate of the vortex cores. ple, the problem of vortex wandering must clearly be an
issue as it is in the laboratory (Refs. 32, 33). The measure- The importance and sensititivity of vortex filament
ments of Boatright (Ref. 16) were noted to exhibit sub- strain on the growth properties of tip vortices can be il-
stantial scatter, most likely because of this problem and lustrated with reference to the viscous development of a
are not shown here. The measurements of Mahalingam et rectilinear vortex at a particular vortex Reynolds num-
al. (Ref. 18) suggest a somewhat higher core growth rate ber. The effects on the filaments were examined as a
at young wake ages than that might be expected based on function of prescribed strain rate defined by Vε = dε/dt.
the vortex Reynolds number. This could be because the Figure 16 shows the growth rate predicted by the model
measurements were made using a rotor operating in for- given by Eq. 14 for a rectilinear vortex filament operat-
ward flight; as previously alluded to in this case the vortex ing at Rev = 5 × 105 . This is a representative value of
filaments experience velocity gradients in the rotor wake, most model-scale rotor measurements. This figure also in-
which can strain or squeeze filaments and can affect the cludes the core growth predicted by the model for a vortex
measured growth properties of the tip vortices. The im- filament undergoing a constant filament strain rate, both
portance of vortex filament strain and its interdependence positive and negative. It can be observed that an appli-
on diffusion (and vortex Reynolds number) is explained in cation of positive filament strain rate results in a reduced
the next section. core growth rate. Generally, these results show that the
effects of stretching counter the effects of diffusion with
Combining Diffusion & Stretching Model a notable reduction in the core size. On the other hand, a
negative filament strain results in an increased core growth
The growth of the vortex core, as shown in Eq. 14, de- rate, showing that the effects of strain enhances the effects
pends not only on the vortex Reynolds number but also on of diffusion. This would mean that under the action of
the local velocity gradient experienced by the vortex fila- certain velocity gradients the vortex would exhibit an in-
ments as they age in the flow. These local velocity gradi- creased core growth rate even if the vortex Reynolds num-
ents can stretch or squeeze the vortex filaments, changing ber is small.
the vorticity (but not the net circulation) and so producing The core growth predicted by the model at different vor-
a different induced velocity field. This is especially impor- tex Reynolds numbers when the vortex filament undergoes
tant in the vortex ring state when the helicopter rotor oper- a constant positive strain rate is shown in Fig. 17. The core
ates in forward flight, or in ground effect, where velocity growth predicted by the model in the absence of strain for
gradients are typically higher. Ananthan et al. (Ref. 22) the same type of vortex Reynolds numbers is shown in
used a free vortex-wake method to show that the magni- Fig. 18. It can be observed that the core growth changes
tude of the velocity gradients when the rotor operates in both with the sign and magnitude of the strain on the vor-
forward flight are large enough to have significant effects tex filament and also with vortex Reynolds number. Even
on growth properties of tip vortices. This suggests that if the magnitude of diffusion of the vortex is increased
any vortex measurements that are made in forward flight through an increase in turbulence in the core (or an in-
must be corrected for the effects of strain if the intent is to crease in vortex Reynolds number), the vortex filament
isolate the competing mechanisms of stretching and diffu-
sion from each other. The ability to do this would seem to Re = 100
v
be a prerequisite to validate any kind of blade tip vortex Re = 1 X 10
4
v
model. 5
Re = 1 X 10
v
0.1 0.06
Non-dimensional core radius, r c/c

6
Non-dimensional core radius, r c/c

Re = 1 X 10
v

Rectilinear vortex, Vε = 0 0.05


0.08
Positive strain, V = 0.25
ε

Negative strain, V = -0.25 0.04


ε
0.06
0.03
0.04
0.02

0.02 0.01

0 0
0 360 720 0 360 720
Wake age, ζ (deg.) Wake age, ζ (deg.)
Figure 16: Variation of core growth predicted by the Figure 17: Variation of core growth predicted by the
new model for different values of uniform strain rates new model for a uniform strain rate vε = 0.25 at dif-
at Rev = 5 × 105 . ferent vortex Reynolds numbers.
Re = 100 ported to be significant. An increased or decreased value
v

4
of positive filament strain corresponds to reduced or in-
Re = 1 X 10
v
5
creased core size, respectively, for the same wake age.
0.06 Re = 1 X 10
v
This change in core size can significantly alter the wake
Non-dimensional core radius, r c/c

6
Re = 1 X 10
0.05
v geometry and the induced velocities inside the rotor wake.
The implications are that this will also affect predictions
0.04 of BVI airloads on the rotor and also BVI induced noise.

0.03

0.02
Conclusions
0.01 A vortex model in terms of vortex Reynolds number was
successfully developed and validated using available rotor
0 tip vortex measurements. The model was developed by
0 360 720
combining two experimentally validated individual vor-
Wake age, ζ (deg.)
tex models that take into account the effects of filament
Figure 18: Variation of core growth predicted by the stretching and rotational flow (Richardson number) effects
new model for a rectilinear vortex, vε = 0.0, at different on the turbulence present inside the tip vortices in predict-
vortex Reynolds numbers. ing their temporal evolution. An useful exponential series
approximation was also developed to represent the swirl
might show reduced core growth and increased vorticity
velocity distribution predicted by the transitional vortex
(with higher peak swirl velocity values) if the vortex fila-
model. The following conclusions have been derived from
ment undergoes high positive strain. Overall, these results
the work:
show the highly interdependent nature of filament strain
and diffusion (or vortex Reynolds number) on the growth
1. Vortex filament strain and diffusion efffects are inter-
of tip vortices.
dependent processes. Tip vortices will show an in-
For rotors operating near ground, the magnitude of ve-
creased growth rate even when the vortex Reynolds
locity gradients will be of the same order of magnitude
number is small if the filament undergoes signifi-
as that of forward flight. Strain values measured using
cant (negative) strain. Similarly, the tip vortices will
a model scale rotor operating in the presence of a solid
exhibit reduced growth rate despite being at higher
boundary (Ref. 13) were applied to the core growth model
Reynolds numbers if they experience large positive
given by Eq. 14, the results of which are shown in Fig. 19.
strain.
It can be observed that the model correlates well with the
measurements for a value of δ that represents the core 2. The core growth rate predicted by the model for a
growth without the effects of vortex filament strain. Also, rectilinear vortex filament increases with an increase
the core growth predicted by the new model deviates from in vortex Reynolds number (for Rev > 1, 000). In-
the simple diffusion based Squire core growth model at creasing the vortex Reynolds number increases the
older wake ages where the vortex filament strains were re- turbulence present inside the tip vortices and, hence,
0.25 gives an increased core growth rate. At lower Rev ,
Lamb-Oseen model however, the core growth rate reduces to the laminar
Non-dimensional core radius, rc/ c

Squire model, δ = 8
0.2 Martin & Leishman, 2001 Lamb–Oseen model by being independent of vortex
Ramasamy & Leishman, 2003
New model with measured strain rate Reynolds number but with a finite core radius at time
New model with reduced strain rate
0.15 New model with increased strain rate δ=8 t = 0.

3. An exponential series approximation emulated the


0.1
properties of the swirl velocity distribution predicted
δ = 1 (laminar)
by the transitional vortex model at a much lower
0.05
computational cost. This provides a useful approach
for incorporating vortex Reynolds number effects
0
0 180 360 540 720 900
into the vortex model for use in a variety of aeroa-
Wake age, ζ (deg.) coustic applications.
Figure 19: Core growth predicted by the new model 4. The effect of strain on growth properties of a tip vor-
for measured values of strain rates. Rev = 48, 000. tex at a given vortex Reynolds number was found
to favor, balance, or counter the effects of diffusion
based on the nature and magnitude of the stretch- 11 Ramasamy, M., and Leishman, J. G., “Interdependence

ing. Vortex measurements that are made in velocity of Diffusion and Straining of Helicopter Blade Tip Vor-
gradients must be corrected for the effects of vortex tices,” Journal of Aircraft, Vol. 41, No. 5, Sept.–Oct.
filament strain if the results are to be interpreted in 2004, pp. 1014–1024.
correct manner for the development of better vortex
12 McAlister, K., “Rotor Wake Development During
models.
the First Rotor Revolution,” American Helicopter Soci-
ety 59th Annual National Forum, Phoenix, AZ, May 6–8
Acknowledgments 2003.
13 Ramasamy, M., and Leishman, J. G., “A Generalized
This research was supported, in part, by the National Ro-
Model For Transitional Blade Tip Vortices,” American
torcraft Technology Center under Grant NCC 2944.
Helicopter Society 60th Annual National Forum, Balti-
more, MD, June 7–11, 2004.
References 14 Tung,C., Pucci, S. L., Caradonna, F. X., and Morse,
1 Squire,
H. A., “The Structure of Trailing Vortices Generated
H. B., “The Growth of a Vortex in Turbulent by Model Helicopter Rotor Blades,” NASA TM 81316,
Flow,” Aeronautical Quarterly, Vol. 16, August 1965, 1981.
pp. 302–306.
15 Hoffman, E. R., and Joubert, P. N., “Turbulent Line
2 Iversen,
J. D., “Correlation of Turbulent Trailing Vor- Vortices,” Journal of Fluid Mechanics, Vol. 16, 1963,
tex Decay Data,” Journal of Aircraft, Vol. 13, No. 5, May pp. 395–411.
1976, pp. 338–342.
16 Boatwright, D. W., “Three-Dimensional Measure-
3 Govindraju, S. P., and Saffman, P. G., “Flow in a Tur- ments of the Velocity in the Near Flow Field of a Full-
bulent Trailing Vortex,” Physics of Fluids, Vol. 14, No. 10, Scale Hovering Rotor,” Mississippi State University Re-
October 1971, pp. 2074–2080. port EIRS-ASE-74-4, 1974.
4 Zeman, O., “The Persistence of Trailing Vortices: A 17 Cook, C. V.,“The Structure of the Rotor Blade Tip Vor-
Modeling Study,” Physics of Fluids, Vol. 7, No. 1, January tex,” Paper 3, Aerodynamics of Rotary Wings, AGARD
1995, pp. 135–143. CP-111, Sept. 13–15, 1972.
5 Baldwin, B. S., Chigier, N. A., and Sheaffer, Y. S., 18 Mahalingam, R., and Komerath, N. M., “Measure-
“Decay of Far-Flowfield in Trailing Vortices,” AIAA ments of the Near Wake of a Rotor in Forward Flight,”
2nd Atmospheric Flight Mechanics Conference, Palo Alto, AIAA Paper 98-0692, 36th Aerospace Sciences Meeting
California, Sept. 11–13, 1972, 72–89. & Exhibit, Reno, NV, January 12–15, 1998.
6 Lamb, H., Hydrodynamics, 6th ed., Cambridge Uni- 19 McAlister, K. W., “Measurements in the Near Wake
versity Press, Cambridge, 1932. of a Hovering Rotor,” AIAA Paper 96-1958, Proceedings
of 27th AIAA Fluid Dynamic Conference, New Orleans,
7 Oseen, C. W., “Uber Wirbelbewegung in einer reiben- LA, June 18–20, 1996.
den Flüssigkeit,” Arkiv för Matematik, Astronomi och
20 Bhagwat, M. J., and Leishman, J. G., “Correlation
Fysik., Vol. 7, No. 14, 1911, pp. 14–21.
of Helicopter Tip Vortex Measurements,” AIAA Journal,
8 Ciffone,D. L., and Orloff, K. L., “Far-Field Wake- Vol. 38, No. 2, Feb. 2000, pp. 301–308.
Vortex Characteristics of Wings,” Journal of Aircraft,
21 Yu.,Y. H., and Tung, C., “The HART-II Test: Rotor
Vol. 12, No. 5, May 1975, pp. 464–470.
Wakes and Aeroacoustics with Higher Harmonic Pitch
9 Singh, P. I., and Uberoi, M. S.,
“Experiments on Vortex Control (HHC) Inputs,” American Helicopter Society 58th
Stability of Aircraft Wake Turbulence,” The Physics of Annual National Forum, Montréal, Canada, June 11–13,
Fluids, Vol. 19, No. 12, Dec. 1976, pp. 1858–1863. 2002.
10 Martin,P. B., Pugliese, G., and Leishman, J. G., “High 22 Ananthan, S., and Leishman, J. G., “Role of Fila-
Resolution Trailing Vortex Measurements in the Wake of ment Strain in the Free-Vortex Modeling of Rotor Wakes,”
a Hovering Rotor,” Journal of the American Helicopter Journal of the American Helicopter Society, Vol. 49,
Society, Vol. 48, No. 1, Jan. 2003, pp. 39–52. No. 2, April 2004, pp. 176–191.
23 Bhagwat, M. J., and Leishman, J. G., “Generalized

Viscous Vortex Core Models for Application to Free-


Vortex Wake and Aeroacoustic Calculations,” Proceed-
ings of the 58th Annual Forum of the American Heli-
copter Society International, Montréal Cananda, June 11–
13, 2002.
24 Iversen, J. D., and Corsiglia, V. R., “Hot-Wire,
Laser-Anemometer, and Force Measurements of Interact-
ing Trailing Vortices,” Journal of Aircraft, Vol. 16, No. 7,
July 1979, pp. 448–454.
25 Bradshaw,P., “The Analogy Between Streamline Cur-
vature and Buyoyancy in Turbulent Shear Flows,” Journal
of Fluid Mechanics, Vol. 36, 1969, pp. 177–191.
26 Holzapfel,A. Hofbauer, T., Gerz, T., and Schumann,
U., “Aircraft Wake Vortex Evolution and Decay in Ide-
alized and Real Environtments: Methodologies, Benefits
and Limitations,” Proceedings of the Euromech Collo-
quium, 2001.
27 Cotel, A. J., and Breidenthal, R. E.,
“Turbulence Inside
a Vortex,” Physics of Fluids, Vol. 11, No. 10, Oct. 1999,
pp. 3026–3029.
28 Cotel,
A. J., “Turbulence Inside a Vortex: Take Two,”
Physics of Fluids, Vol. 14, No. 8, Aug. 2002, pp. 2933–
2934.
29 Cotel,A. J., Gjestvang, J. A., RamKhelawan, N. N.,
and Breidenthal, R. E., “Laboratory Experiments of a Jet
Impinging on a Stratified Surface,” Experiments in Fluids,
Vol. 23, No. 2, June 1997, pp. 155–160.
30 Vatistas,
G. H., Kozel, V., and Mih, W. C., “A Simpler
Model for Concentrated Vortices,” Experiments in Fluids,
Vol. 11, 1991, pp. 73–76.
31 Vatistas,
G. H., “New Model for Intense Self-Similar
Vortices,” Journal of Propulsion and Power, Vol. 14,
No. 4, April 1998, pp. 462–469.
32 Leishman,J. G., “On the Aperiodicity of Helicopter
Rotor Wakes,” Experiments in Fluids, Vol. 25, 1998,
pp. 352–361.
33 Devenport, W. J., Rife, M. C., Liapis, S. I., and Follin,
G. J., “The Structure and Development of a Wing-Tip Vor-
tex,” Journal of Fluid Mechanics, Vol. 312, 1996, pp. 67–
106.

Das könnte Ihnen auch gefallen