Vous êtes sur la page 1sur 7

374 J. AIRCRAFT VOL. 28, NO.

Numerical Solutions of Forward-Flight Rotor Flow


Using an Upwind Method
C. L. Chen*
Rockwell International Science Center, Thousand Oaks, California 91360
and
W. J. McCroskeyt and S. ObayashiJ
NASA Ames Research Center, Moffett Field, California 94035

A finite-volume upwind algorithm for solving the three-dimensional Euler equations with a moving grid has
been developed for computing helicopter forward-flight rotor flows. The computed pressure distributions and
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

shock positions of high-speed rotor flow are compared with various experimental data as well as with other
numerical results, and the agreement is encouraging. A comparison of quasisteady solutions with unsteady
solutions reveals that when a shock occurs in the flowfield, the assumption of quasisteady flow may fail due to
the time lag of the shock motion. Similarly, three-dimensional effects cannot be neglected. Sufficient subitera-
tions for each time step are required to avoid numerical lag effects in using the present method. The redistribu-
tion of the residual due to the coordinate transformation is discussed. For high-order Monotone Upstream-
Centered Conservation Law (MUSCL)-type schemes, a coordinate-independent solution can be obtained by
interpolating primitive variables.

Introduction In the process of validating the approach, three-dimensional


and unsteady time-lag shock motion effects for for ward-flight
A NUMBER of unsteady Euler/Navier-Stokes solvers have
been developed for helicopter rotor flow, including cen-
tral-difference methods and partial flux-splitting methods.1'8
rotor flow are examined. Both of the nonlifting rectangular
untwisted rotors studied have NACA 0012 profiles with con-
The latter consists of upwind differencing in the mainstream stant chord.
direction and central differencing in the other directions. Up-
wind algorithms in all of the directions, based on Roe's Governing Equations
scheme, have been widely and successfully used for various The basic equations under consideration are the unsteady
complex fluid flow computations.9'12 Such schemes not only Euler equations with general grid motions. The computational
capture shocks well but also predict accurately boundary-layer domain consists of structure-wise hexahedral cells. Each cell
quantities in viscous calculations.13 Therefore, it seems useful surface can be identified in the £, 77, or f direction, respec-
to extend the scheme to unsteady rotor flows with a moving- tively. The general form of a conservation law applied to an
grid system. Unfortunately, for implicit schemes, the true flux arbitrary grid cell is
Jacobian of Roe's scheme is too complicated and the direct
solver is inefficient for practical use. Thus, an implicit itera-
tive procedure, using Jameson and Toon's simplified flux QdV- QdV+ (1)
Jacobian, is adopted in the present approach. The LU-SGS 5(0
method12'14 is used along with quasi-Newton iterations to en- where
sure time-accurate solutions. The finite-volume cell-averaged
discretization procedure is unchanged from previous work.7'8 P pu
The primary objectives of this paper are to demonstrate the
ability of a finite-volume upwind algorithm based on an up-
pu puu + pnx
wind-biased flux-difference-splitting approach to calculate ro- pv pvu + pny
tor flows and to improve upon the efficiency of previous pw pwu + pnz
work.7'8 This raises some numerical issues regarding coordi- e eu + pun
nate dependency between the inertial and wing-fixed frame.
Thus, the residual redistribution for each cell due to the coor-
dinate transformation is examined. where V(t) is the cell volume, ndS(t) is a vector element of
surface area with outward unit normal vector n(nx,ny9nz)9 Q
is a vector of conservative variables per unit volume, p and u,
v, w are the density and the fluid velocity components in the
x, y, z directions, respectively, and e is the total internal
Presented as Paper 89-1846 at the AIAA 20th Fluid Dynamics, energy per unit volume. In addition, Fis the flux of Q per unit
Plasma Dynamics, and Lasers Conference, Buffalo, NY, June 12-14, area per unit time, p is the pressure, un - n - u is the normal
1989; received Jan. 20, 1990; revision received July 9, 1990; accepted velocity component of the fluid, vn - n - v where v is the grid
for publication Aug. 26, 1990. Copyright © 1990 by the American velocity, and u = n • (u - v) = un - vn is the normal relative
Institute of Aeronautics and Astronautics, Inc. No copyright is as- velocity component. The term vn is not a function of time and
serted in the United States under Title 17, U.S. Code. The U.S. is fixed at each cell for the grid with constant rigid-body
Government has a royalty-free license to exercise all rights under the rotation.
copyright claimed herein for Governmental purposes. All other rights
are reserved by the copyright owner. The preceding Euler equations are written using the abso-
*Member Technical Staff. Member AIAA. lute variables in the inertial frame. In our calculation of for-
tSenior Staff Scientist. Associate Fellow AIAA. ward-flight rotor flow, this formulation was based mainly on
JNRC Research Associate. Member AIAA. two reasons: 1) the far-field boundary conditions are easily
JUNE 1991 FORWARD-FLIGHT ROTOR FLOW 375

implemented, and 2) the formulation is in conservation form. where


However, this formulation can be recast in terms of absolute
variables written in a rotational frame.15'16 In other words, the | | 5 \(Ap-Ap/c2)
absolute fluid velocity, which is the velocity relative to an
inertial frame, can be represented as (u*, v*, w*) in the | 1 S + c | (Ap + pcAu)/2c2
rotational frame. Similarly, the surface unit vector n can be
represented as (n*, n*, n*) and calculated only once without | | u -c | (Ap-pcAu)/2c2
the rotational matrix being involved.8 The equations remain in
exactly the same form as in the original formulation, except 0.2 + 0:3
for an additional source term R, which is introduced into the
Ci5 = ~ Ct3)
right side of the equation as
a 6 = \S | \S \(pAu-nxpAu)
Q*dV Q*dV F* dSdt=R (2) a7 = \S\ | u \(pAv-nypAu)

as = | S | | u | (pAw-nzpAu)
where
where Ap = pR — pL , and so on, H is the total enthalpy, c is the
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

pu 0 sound speed, and superscript ~ denotes Roe-averaged quanti-


ties. The area magnitude of the cell interface S is | S \ ,
pu* pu*u+pn* Qpv* n = S/\ S | , and u = n • (u-v) = un-vn. The so-called time
pv* pv*ii+pn* R =V -Qpu* metrics for the grid motion are included in the evaluations of
pw* pw*ii+pn* 0 A, F (QR,S), and F (QL,S). This is considered to be more
straightforward than the extension of Van Leer flux splitting
e eu +pun 0 to a moving-grid system.17
The scheme yields first-order accuracy if QL = Qj and
and Q is the angular velocity in the z direction relative to the QR = QJ+I- Higher order schemes are constructed from a
inertial frame. The next section shows how easily the present one-parameter family of interpolations of the primitive vari-
algorithm can be adapted to this formulation. The advantages ables q = (p, u, v, w, p)T:
of the latter formulation are shown for the case of a hovering
rotor with constant rotating speed and constant ascent (or - K) V, (5a)
descent) speed. The far-field boundary conditions are still very
easy to implement for this specific case, and the flow may be (5b)
treated as independent of time (if unsteadiness in the wake can
be ignored). The source term in the preceding discussion is where AC = (1/3) is used in the following for the third-order
nonzero only in the momentum equations. Thus, for the full- computations, and K = - 1 for the second-order fully upwind
potential equation, where only the continuity equation is con- method. The symbols V and A are backward and forward
sidered, it is unnecessary to discuss the absolute variables difference operators, respectively, which are limited to ensure
written in either the inertial or the rotational frame. the monotone interpolation18 by
3AV + e
5- = (6)
Numerical Algorithm 2(V - A)2 + 3AV
Upwind Method with e < 1.0J51 - 6 for K = (1/3). The reasons for choosing
A finite-volume upwind method developed especially for primitive variables for interpolation will be explained later.
helicopter rotor flow is described here. The convective terms
are differenced with the upwind-biased flux-difference split- LU-SGS Method
ting of Roe for the explicit side, while Jameson and Yoon's14 The Jameson and Yoon LU-SGS implicit operator is used in
LU-SGS method is used for the implicit operator. In addition, the present study. Let A, B, C be Jacobian matrices in the £,
quasi-Newton iterations are used at each time step to maintain it], and f directions, respectively, and let the indices j, k, and /
time accuracy. correspond to these directions. The LU-SGS implicit operator
A numerical flux of the inviscid terms, due to Roe's up- consists of L, D, U, as follows:
winding in the locally one-dimensional form, can be written
for the £ (index j) direction as L = VI + At(-A -B-

F(QL ,QR ,Sj+ ,/2) = >/2 [F(QR ,Sj+ 1/2) + F(QL (7a)
+ +
- A(QL,QR,Sj+y2)\(QR-QL)] (3) D = [vi + At(A -A- +B -B- +c
where A is the Roe-averaged Jacobian matrix and QL and QR (7b)
are the state variables to the left and right of the cell interface,
and

(7c)

\A\(QR-QL) = with A± = l/2(A± a^I) and a^ > (u + c) \ S \ , and similiarly for


B ± and C ± . Because of the simplified form of the Jacobian
(7-1)
•Oil
terms A*, B*9 and C*1, the matrix Z) is a sc^ar diagonal
matrix. Thus, this method requires only two (forward and
backward) sweeps with scalar inversions and leads to reduced
(4) factorization error.
376 CHEN, McCROSKEY, AND OBAYASHI J. AIRCRAFT

The whole algorithm applied to Eq. (1) may be symbolically be a fully converged solution in the inertial frame. Thus, the
described as first-order scheme is a coordinate-independent scheme.
For higher order schemes, using Eq. (5), flow variables QL
LDUAQP = - AtRHSP (8a) and QR can always be represented by Eo/Q/, with £«/ = 1 for
each flow variable interpolation. Three equations have to be
where RHSP is the residual of the unsteady Euler equations: solved at each cell for a one-dimensional problem. Here, the
QJ are the neighboring cell flow variables. Without any lim-
RHSp = P + [QnA V + V(QP - Qn)]/At (8t>) iter, the a/ are the same for all the equations. Then UL(UR) will
become UL(UR) - U& at each cell interface, regardless of
with numerical fluxes E, F, and G in the £, 77, and f directions, whether the interpolation is performed on the conservative or
respectively. At any time n , a few iterations on p may be primitive variables. Again, a coordinate-independent scheme
required. In general, p decreases as At is reduced. is easily obtained.
If absolute variables are written in the rotational frame as Once a slope limiter is introduced into each flow variable, a,
Q* in Eq. (2), in addition to a right side source term, a source may not be the same for all the equations. If conservative
term Jacpbian can be added in either the L or (/operator. This variables are used for the interpolation to obtain cell interface
source term Jacobian is state variables, a linear transformation of the velocity field
may not be obtained that guarantees the vanishing of the
residual. However, if primitive variables are used for the
0 0 0 0 0 interpolation/extrapolation, a linear velocity transformation
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

0 0 OF 0 0 relation is maintained. In addition, the interpolation relation


dR for p and p remains the same for both reference frames. This
0 -12F 0 0 0 (9)
dQ* guarantees that the scheme is coordinate-independent. Hence,
0 0 0 0 0 from this point of view, the interpolation of the primitive
0 0 0 0 0 variables would be preferable.

Numerical Results
Slightly more work is then required for 2 x 2 matrix inversions. In the process of validating this algorithm, the present al-
All of the boundary conditions are implemented explicitly, gorithm based on Roe's scheme is first applied to a transonic
as shown in Ref. 8. For inviscid flow, no flow normal to the flow past an airfoil, and the results are compared with those
wall boundary is allowed by imposing u -n = v -n on the solid obtained with the previous method. Subsequently, two for-
wall, where u, v are fluid velocity and grid velocity and n is ward-flight rotor flows were examined; the results are dis-
surface normal unit vector of wall. The wall pressure is ob- cussed and compared with experimental data. Both of the
tained by solving the normal momentum equation. nonlifting rectangular untwisted rotors studied have NACA
0012 profiles with constant chord.
Coordinate-Independent Scheme
The numerical scheme used here is a finite-volume scheme.7'8 NACA 0012 Airfoil
It captures freestream exactly. Another feature of the present A NACA 0012 airfoil at M* = 0.85, a = 1.0 deg was calcu-
scheme is the interpolations with primitive variables for high- lated using the present method, with K = (1/3). The C-type
order schemes, as in Refs. 11 and 12. The numerical solutions grid uses 181 x 31 points. The results are compared in Fig. 1
satisfy exactly the linear transformation of the velocity field with the solutions from a partial flux-splitting method and
between the inertial frame with moving grid and the wing- K = - i without limiter.7'8 The partial flux-splitting method
fixed frame. In other words, if a set of solutions p, u, v, w, p consists of upwind differencing in the mainstream direction
is obtained for freestream past an object in the wing-fixed and central differencing in the other directions. The shock
frame, it is equivalent to a set of solutions p, u-U^ v, w, p locations agree to within one grid point. The leading-edge
for the object moving in the negative x direction in the inertial suction peak is predicted slightly better with the present code.
frame. The relation mentioned is obvious for differential The result also agrees very well with Yee's calculation,19 which
equations, but it may not be exactly true for arbitrarily dis- is not included here. Roe's scheme with Keren's differential
cretized equations, as will be shown in the following discus- limiter can predict a one-point shock transition without strong
sion. A scheme that satisfies this relation will be called a
coordinate-independent scheme.
For simplicity, a one-dimensional flux-balance is considered
for each cell; however, the following discussions are still true -1.0
for multidimensional problems. Suppose the solutionp,u,pis
obtained by solving the equations in the wing-fixed coordi-
nate. The velocity w is then linearly transformed to u - U^,
and p and p remain the same as in the inertial frame. For a -0.5
first-order scheme, the interpolation results in QL = Qj and
QR = Qj+i> where Q may be either a primitive or a conserva-
tive vector. Therefore, at a cell interface, UL(UR) will become
UL(UR) - £/«>, and p and p will remain unchanged. Suppose the s
residuals in the wing-fixed coordinates are (ei, e2, e3) for the
mass, momentum, and energy equations, respectively. By
casting the transformed variables into the numerical-flux eval- 0.5 : ROE'S SCHEME
PARTIAL FLUX-SPLITTING METHOD
uation [Eq. (3)] and applying the conservation law at each cell,
the residuals are redistributed in the inertial coordinate system
as follows: 1.0
1) Residual ei remains the same as in the continuity equation.
2) Residual e2 is changed to e2 - U^EI. -0.1 0 0.10.20.30.40.50.60.70.80.91.01.1
3) Residual e3 is changed to e3 + (L£/2)ei - Ua>e2. X/C
If a fully converged solution (e/ = 0) is obtained for wing- Fig. 1 Surface pressure comparison with various right sides for a
fixed coordinates, the linearly transformed velocity would also NACA 0012 airfoil at Af*, = 0.85, a = 1.0.
JUNE 1991 FORWARD-FLIGHT ROTOR FLOW 377

postshock overshoots. However, for this transonic flow, no Two Forward-Flight Rotors
limiter was required for the partial flux-splitting scheme with The aspect ratio of the first rotor blade is 7.0 and the flight
the fully upwind Monotone Upstream-Centered Conservation conditions consist of tip Mach number MT = 0.8 and advance
Law (MUSCL) approach (K = - 1). Conversely, without a ratio \L = O.2.20 The grid size used was 81 x 25 x 35 points
limiter, numerical oscillations occurred with Roe's approach in an O-H-type topology. For this forward-flight rotor, the
using K = - I . initial solution was a hovering solution obtained by solving
Eq. (2). Then, the forward-flight speed was imposed at the
far field, and the time-accurate calculation of Eq. (1) was
started at azimuthal angle ^ = 0.0 deg. The solution from the
-1.0 first to the fifth revolution at ^ = 30 deg is shown in Fig. 2.
The numerical result shows a recursive surface pressure distri-
bution beginning at the second revolution, indicating the es-
tablishment of a periodic flow. The solutions for span wise
- 0.5 station r/R =0.88 and 0.90 at various azimuthal locations are
shown in Fig. 3 and compared with experimental data at
r/R = 0.89 (r/R = 1.0 at blade tip). The location of the shock
is predicted well. Shock locations may differ in one point with
different methods. The partial flux-splitting method occasion-
Q. 0.0 ally predicts shock locations a point farther toward the trailing
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

o
edge, as shown in Fig. 4. The leading-edge suction peak is
D 1st REVOLUTION slightly lower than that obtained from Roe's scheme with
O 2nd REVOLUTION limiter. These effects have also been observed in the airfoil
0.5
A 3rd REVOLUTION
calculation described earlier. Overall, from accuracy consider-
+ 4th REVOLUTION
X 5th REVOLUTION
O EXPERIMENTAL DATA

-1.0-
1.0
0.2 0.4 0.6 0.8 1.0
X/C
Fig. 2 Surface pressure at V - 30 deg with various revolutions for a
forward-flight rotor; MT - 0.8, fi = 0.2, aspect ratio = 7, nonlifting
untwisted untapered NACA 0012 blade.

-1:0

-0.5
0.88 ROE'S SCHEME
0.90 ROE'S SCHEME
0.0 0.88 PARTIAL FLUX-SPLITTING METHOD

0.5

> = 30° 120° 0.2 0.4 0.6 0.8 1.0


1.0
-1.0

Fig. 4 Surface pressure results with various right sides for a forward-
-0.5 flight rotor at ^ = 90 deg, MT = 0.8, /* = 0.2, aspect ratio = 7, non-
lifting untwisted untapered NACA 0012 blade.

Cp 0.0

0.5

= 60° 150°
1.0
-1.0

-0.5
O

0.0
-CALCULATIONS (r/R = 0.88)
CALCULATIONS (r/R = 0.90)
1 EXPERIMENTAL DATA (0.89)
0.5

90° = 180°

1.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
X/C
-2.0-1.5 -1.0 -0.5 0 0.5 1.0 1.5 2.0
Fig. 3 Surface pressure results at the advancing side for a forward-
X/C
flight two-blade rotor; MT = 0.8, p. = 0.2, aspect ratio = 7, nonlifting
untwisted untapered NACA 0012 blade. Fig. 5 Quasisteady pressure contours at blade plane for ¥ = 60 deg.
378 CHEN, McCROSKEY, AND OBAYASHI J. AIRCRAFT

o
o

-2.0-1.5 -1.0 -0.5 0 0.5 1.0 1.5 2.0 -2.0-1.5-1.0-0.5 0 0.5 1.0 1.5 2.0
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

X/C X/C
Fig. 6 Quasisteady pressure contours at blade plane for ¥ = 120 deg. Fig. 8 Unsteady pressure contours at blade plane for ¥ = 120 deg.

o
0.88, 4 SUBITERATIONS
0.90, 4 SUBITERATIONS
0.88, 2 SUBITERATIONS

60°
1.0
-2.0-1.5 -1.0 -0.5 0 0.5 1.0 1.5 2.0 0.2 0.4 0.6 0.8 1.0
X/C
Fig. 7 Unsteady pressure contours at blade plane for ¥ = 60 deg.
Fig. 9 Surface pressure at ¥ = 180 deg with various amounts of
subiteration for the rotor in Fig. 2.
ation, the advantage of the present scheme over the partial
flux-splitting method is not dramatic for the nonlifting invis- may not be sufficient to model rotor flows with unsteady
cid rotor flow calculation, but it avoids specification of ex- shock motions. The preceding observations of three-dimen-
plicit dissipation in the crossflow direction. sional and time-lag effects agree well with the full-potential
Three-dimensional radial-flow effects are examined by com- results in Ref. 21.
paring quasisteady solutions at ^ = 60 deg (outward radial However, when subcritical flow occurs at certain azimuthal
flow) and ^ = 120 deg (inward radial flow). The pressure locations (for example, ^ = 0.0 deg), the quasisteady solu-
contours at the plane of the rotor blade are shown in Figs. 5 tions are very close to the unsteady solutions. Therefore, it is
and 6. The quasisteady solution neglects the unsteady terms possible to start with a quasisteady solution as the initial
and solves the resulting equations using the instantaneous grid condition at ^ = 0.0 deg for an unsteady calculation as others
velocity associated with the specified azimuthal angle. This have stated.21'22
quasisteady flow cannot be investigated experimentally but is The present code with LU-SGS implicit operator enables
easily computed numerically. Comparing the two plots, cer- marching with two to three times larger time steps than the
tain differences in the pressure on the blade surface are ob- previous code.7'8 It requires about 30 jus per grid point per
served. The inward radial flow at ^ = 120 deg causes the iteration on a CRAY 2, whereas the previous code7'8 required
suction peak to decrease near the tip (supercritical region) and 75 jits. For the calculations performed here, two or three
increase near the root (subcritical region). In addition, the internal iteration per reasonable time step were usually suffi-
shock near the tip is moved slightly aft toward the trailing cient to obtain accurate solutions. (A reasonable time step is
edge. about one-third of a degree of rotation.) However, the num-
The differences in the surface pressure contours between the ber of internal iterations at each time step may need to be
quasisteady flow and the unsteady flow solutions are seen by increased when large time steps are used, otherwise the numer-
comparing Fig. 5 with Fig. 7 and Fig. 6 with Fig. 8. The ical error may play an important role in the solution. This
unsteady solution predicts a weaker shock at ^ = 60 deg than numerical error is due to the increase in linearization and
that predicted by the quasisteady solution, whereas the oppo- factorization error with increasing time step size and is re-
site occurs at ^ = 120 deg. The time-lag effects observed in duced during the iteration process. In Fig. 9, the solution at
these Euler calculations indicate that the quasisteady approach ^ = 180 deg with a time step size of half a degree is compared
JUNE 1991 FORWARD-FLIGHT ROTOR FLOW 379

-1.0 -LOT

-0.5 -0.5

0.0 0.0

0.5 0.5

j = 30° = 120° = 30° 120°


1.0 1.0
-1.0 -1.0

-0.5 -0.5

Cp 0.0 Cp 0.0

0.5
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

0.5

= 60° = 150° = 60° 150°


1.0 1.0
-1.0 -1.0
r/R = 0.88 r/R = 0.95

-0.5 -0.5

0.0 0.0

0.5 0.5

= 90° ip= 180° 90° 180°


1.0 1.0
0.0 0.2 0.4 0.6 0.8 1.00.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
X/C X/C
Fig. 10 Surface pressure results at the advancing side for a forward- Fig. 11 Surface pressure results at the advancing side for a forward-
flight two-blade rotor, MT = 0.7634, p = 0.25, spanwise location r/ flight two-blade rotor, MT = 0.7634, /* = 0.25, spanwise location r/
R = 0.88, aspect ratio = 7.125, nonlifting untwisted untapered NACA R = 0.95, aspect ratio = 7.125, nonlifting untwisted untapered NACA
0012 blade. 0012 blade.

for two and four subiterations at each time step. The solution solver. The residual redistribution due to the coordinate trans-
is converged after four iterations, but not after two. The formation is observed. The primitive variables interpolation is
comparison in Fig. 9 indicates that the present scheme requires preferred to obtain coordinate-independent solution.
enough subiterations to avoid numerical time lag. Overall, the
present code is twice as fast as the previous code7 for each run.
The speed-up is mainly due to the efficiency of the LU-SGS References
implicit operator. iSankar, N. L., Wake, B. E., and Lekoudis, S. G., "Solution of the
The flight conditions for the second rotor problem (aspect Unsteady Euler Equations for Fixed and Rotor Wing Configura-
ratio = 7.125) are MT = 0.763 and p = 0.25. The grid size used tions," Journal of Aircraft, Vol. 23, No. 4, 1986, pp. 283-289.
2
was 121x 25 x 35 points in an O-H-type grid. The numerical Chang, I. C., and Tung, C., "Euler Solution of the Transonic
solutions are compared in Figs. 10 and 11 with experimental Flow for a Helicopter Rotor," AIAA Paper 87-0523, Jan. 1987.
3
Agarwal, R. K., and Deese, J. E., "An Euler Solver for Calculat-
data at various spanwise locations and at various azimuthal ing the Flowfield of a Helicopter Rotor in Hover and Forward
angles. The shock locations are predicted reasonably well. Flight," AIAA Paper 87-1427, June 1987.
However, the predicted leading-edge suction peaks are not 4
Wake, B. E., Sankar, N. L., and Lekoudis, S. G., "Computation
as high as in the experimental data. This phenomenon was of Rotor Blade Flows Using the Euler Equations," Journal of Air-
not obvious in the previous test case. Some of the discrepancy craft, Vol. 23, No. 7, 1986, pp. 582-588.
5
may be due to tunnel blockage effects (the model occupies Stahl, H., "Application of a 3D Euler Code to Transonic Blade
about 8% of the test section), mentioned by Bridgeman et al. Tip Blow," Proceedings of the 12th European Rotorcraft Forum,
in Ref. 22. Paper No. 29, Garmish-Partenkirchen, Germany, 1986.
6
Kramer, E., Hertel, J., and Wagner, S., "Computation of Sub-
sonic and Transonic Helicopter Rotor Flow Using Euler Equations,"
Proceedings of the 13th European Rotorcraft Forum, Paper No. 2-14,
Concluding Remarks Aries, France, 1987.
7
A finite-volume upwind algorithm for solving the three-di- Chen, C. L., McCroskey, W. J., and Ying, S. X., "Euler Solution
mensional Euler equations with a moving grid has been devel- of Multiblade Rotor Flow," Vertica, Vol. 12, Oct. 1988, pp. 303-313.
8
oped and applied to compute helicopter rotor flows under Chen, C. L., and McCroskey, W. J., "Numerical Simulation of
forward-flight conditions. This code, which is twice as fast as Helicopter Multi-Bladed Rotor Flow," AIAA Paper 88-0046, Jan.
1988.
that in Ref. 7, is based on Roe's scheme with limiter, an 9
Chakraverthy, S. R., and Szema, K. Y., "An Euler Solver for
LU-SGS implicit operator, and quasi-Newton iterations. The Three-Dimensional Supersonic Flows with Subsonic Pockets," AIAA
numerical results agree well with experimental data as well as Paper 85-1703, July 1985.
with other numerical results. Three-dimensional radial-flow 10
Lawrence, S. L., Tannehill, J. C., and Chaussee, D. S., "An
effects, time-lag effects, and some numerical issues regarding Upwind Algorithm for the Parabolized Navier-Stokes Equations,"
the solution for rotor flows are examined using this Euler AIAA Journal, Vol. 27, No. 9, 1989, pp. 1175-1183.
380 CHEN, McCROSKEY, AND OBAYASHI J. AIRCRAFT

H
Vatsa, V. N., Thomas, J. L., and Wedan, B. W., "Navier-Stokes nates," AIAA Journal, Vol. 26, No. 1, 1988, pp. 113-115.
Computations of Prolate Spheroids at Angle of Attack," AIAA Pa- 18
Koren, B., "Upwind Schemes, Multigrid and Defect Correction
per SI-2621, Aug. 1987. for the Steady Navier-Stokes Equations," Proceedings of the llth
12
Obayashi, S., "Numerical Simulation of Underexpanded Plumes International Conference on Numerical Methods in Fluid Dynamics,
Using Upwind Methods," AIAA Paper 88-4360, Aug. 1988. June 1988, Springer-Verlag, Berlin, 1989, pp. 344-348.
13 19
Van Leer, B., Thomas, J. L., Roe, P. L., and Newsome, R. W., Yee, H., "Implicit TVD Schemes for Hyperbolic Conservation
"A Comparison of Numerical Flux Formulas for the Euler and Laws in Curvilinear Coordinates," AIAA Journal, Vol. 25, No. 2,
Navier-Stokes Equations," AIAA Paper 87-1104, June 1987. 1987, pp. 266-274.
14 20
Jameson, A., and Yoon, S., "Lower-Upper Implicit Schemes Caradonna, F. X., Laub, G. H., and Tung, C., "An Experimen-
with Multiple Grids for the Euler Equations," AIAA Journal, Vol. tal Investigation of the Parallel Blade-Vortex Interaction," NASA
25, No. 7, 1987, pp. 929-935. TM-86005, Nov. 1984.
15 21
Holmes, D. G., and Tong, S. S., "A Three-Dimensional Euler Chang, I. C., "Transonic Flow Analysis for Rotors, Part 2—
Solver for Turbomachinery Blade Rows," Journal of Engineering for Three-Dimensional, Unsteady, Full-Potential Calculations," NASA
Gas Turbines and Power, Vol. 107, No. 2, 1985, pp. 258-264. TP-2375, Jan. 1985.
16 22
Vinokur, M., "An Analysis of Finite-Difference and Finite-Vol- Bridgeman, J. O., Strawn, R. C., and Caradonna, F. X., "An
ume Formulations of Conservation Laws," NASA CR-177416, June Entropy and Viscosity Corrected Potential Method for Rotor Perfor-
1986. mance Prediction," 44th Annual Forum of the American Helicopter
Downloaded by OLD DOMINION UNIVERSITY on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.46037

17
Parpia, I. H., "Van Leer Flux Vector Splitting in Moving Coordi- Society, American Helicopter Society, Washington, DC, June 1988.

Recommended Reading from the AIAA


Progress in Astronautics and Aeronautics Series .

Dynamics of Flames and Reactive


Systems and Dynamics of Shock Waves,
Explosions, and Detonations
J. R. Bowen, N. Manson, A. K. Oppenheim, and R. I. Soloukhin, editors
The dynamics of explosions is concerned principally with the interrelationship between
the rate processes of energy deposition in a compressible medium and its' concurrent
nonsteady flow as it occurs typically in explosion phenomena. Dynamics of reactive
systems is a broader term referring to the processes of coupling between the
dynamics of fluid flow and molecular transformations in reactive media occurring in
any combustion system. Dynamics of Flames and Reactive Systems covers premixed
flames, diffusion flames, turbulent combustion, constant volume combustion, spray
combustion nonequilibrium flows, and combustion diagnostics. Dynamics of Shock
Waves, Explosions and Detonations covers detonations in gaseous mixtures, detona-
tions in two-phase systems, condensed explosives, explosions and interactions.

Dynamics of Flames and Dynamics of Shock Waves,


Reactive Systems Explosions and Detonations
1985 766 pp. illus., Hardback 1985 595 pp., illus. Hardback
ISBN 0-915928-92-2 ISBN 0-915928-91-4
AIAA Members $59.95 AIAA Members $54.95
Nonmembers $92.95 Nonmembers $86.95
Order Number V-95 Order Number V-94

TO ORDER: Write, Phone or FAX: American Institute of Aeronautics and Astronautics, c/o TASCO,
9 Jay Gould Ct., P.O. Box 753, Waldorf, MD 20604 Phone (301) 645-5643, Dept. 415 FAX (301) 843-0159
Sales Tax: CA residents, 7%; DC, 6%. Add $4.75 for shipping and handling of 1 to 4 books (Call for rates on higher quantities). Orders
under $50,00 must be prepaid. Foreign orders must be prepaid. Please allow 4 weeks for delivery. Prices are subject to change without
notice. Returns will be accepted within 15 days.

Vous aimerez peut-être aussi