This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
x + a f 2 ) + ...],
(20)
with the definition a = dcp/dX. On inserting these expansions into Eqs. (14) (where we have set K = 0 for the sake of simplicity) and collecting like powers of E, the following hierarchy of equations is obtained: 0(E-1):
- a U u o = 0,
(21.1)
- o U V 0 ,YY + (a 2 U + U Y Y ) a v 0 + 2 a 2 U UI = 0,
(21.2)
- a U u i - U Y v o = 0,
(21.3)
0(1):
- a U VI^YY + (a 2 U + U Y Y ) O V\ + 2 a 2 U u 2 = -VO,YYYY + V V^YYY +
+ U vo,XYY + (2a 2 + Vy) VO,YY - (<*2V + UXY) VO,Y - (« 2 U + UYY) VO,X + - ( a 4 + <X2VY + UXYY) VO - 2 a U X UI,Y - 2 a U X Y UI ,
(21.4)
- a U U2 - U Y VI = -U^YY + V UI?Y + Uui^x + ( a 2 + Ux) ui,
(21.5)
0(E):
etc.
11 However, a difficulty, common to many other boundary layer perturbation expansions, is encountered: the expansion given above is not uniformly valid in Y, because the highest Y-derivatives are lost at leading order and with them disappears the possibility of enforcing a certain number of boundary conditions. Exactly the same difficulty turns up, for instance, in the derivation of the Orr-Sommerfeld equation as applied to non-parallel problems. The possible solutions to this difficulty are twofold. One approach is to develop a multiple-deck description of the Y-structure of the problem, as, for instance, in Hall's 9 analysis of the large wavenumber limit in which the amplification turns back into damping far enough downstream. Alternatively, one can promote the largest Y-derivative terms to an earlier order in the hierarchy than their formal dependence on E would suggest, thus obtaining the modified equations - a U VO,YY + (a 2 U + Uyy) a VO + 2 a 2 U UI + E [VO,YYYY - V VO,YYY + - (2a 2 + Vy) VO,YY + (a 2 V + U X Y ) VO,Y + ( a 4 + CX2VY + UXYY) VO +
+ 2 a Ux UI,Y + 2 a UXY ui] = 0,
(22.1)
- a U u i - U y v o + E [UI,YY - V UI,Y - (ex2 + Ux) ui] = 0,
(22.2)
- a U V^YY + (oc2U + U Y Y ) a VI + 2 a 2 U U2 + E [V^YYYY - V V I Y Y Y + -(2a 2 + V Y ) VI?YY
+
(a 2 V
+ U X Y ) V I,Y + ( a 4 + « 2 V Y + UXYY) V I +
+ 2 a Ux U 2 ,Y + 2 a U X Y *2\ = U v0,xYY - (<*2U + UYY) VO,X, - a U U2 - U Y vi + E [U2,YY - V U2,Y - ( a 2 + Ux) U2] = U ui,x,
(22.3) (22.4)
The price to be paid for this modification is, just like in the Orr-Sommerfeld problem, that the single terms fo, fi, etc. are now themselves functions of E, SO that the final result of truncating the series will not be a polynomial in E; the reward is that a definite most amplified mode shape and amplification factor will be obtained independently of the initial transient from which the mode is generated. Of course, if the initial transient itself is to be studied other techniques must, and can, be employed; but this is a different subject (the so-called receptivity problem). In the just derived hierarchy, Eqs. (22.1-2) form a coupled system of homogeneous ordinary differential equations for the unknowns vo and ui which must be solved with a in the role of eigenvalue. The coefficients of these equations depend parametrically on X and so will the solution (i.e., the numerical solution will be different at successive Xstations). The X-derivatives of this leading-order solution then appear as known righthand-sides in Eqs. (22.3-4), which must be solved as ordinary differential equations in the
12 unknowns vi and U2, and so on goes the sequence. In singular perturbation problems of the multiple-scale type, care must be taken of secular behaviour. What this means in the present case is that Eqs. (22) are linear differential equations that possess at all orders a common lJi.s. homogeneous part. Once the leading-order homogeneous problem is solved by choosing a such that a nontrivial solution exists, the higher order inhomogeneous problems will generally noj admit a solution unless the known r.h.s. satisfies a compatibility condition. For instance, Eqs. (22.3-4), which have a l.h.s. operator of the same form as Eqs. (22.1-2), will not admit a solution unless the rJi.s. is orthogonal to the left eigenvector, or adjoint eigensolution, of that operator. On the other hand, the solution of the leading-order homogeneous problem is determined only up to an arbitrary multiplicative factor, which can generally be an arbitrary function of X. This degree of freedom is used to satisfy the compatibility condition for Eqs. (22.3-4), after which vi and U2 will contain an arbitrary function of X that can be used to satisfy the compatibility condition at the next higher order, and so on. In practice, suppose Eqs. (22.1-2) have been solved, thus providing for a certain abscissa X a set of normalized direct eigenfunctions Vd(X,Y), Ud(X,Y) and adjoint eigenfunction v a (X,Y), u a (X,Y), together with the eigenvalue a(X). The actual leadingorder solution will be expressed as vo = A(X)vd(X,Y), ui = A(X)ud(X,Y), where A(X) is a yet undetermined function. Substituting these into the compatibility condition for Eqs. (22.3-4) we obtain
J
» 00
0
{[Uv0,XYY " (a 2 U + UYY) VO,X] v a + U UJ ,x u a } dY =
= A X f " {[Uv d;Y Y - (<x2U + UYY) v d ] v a + U u d u a } dY + /% 0 0
+ AJ
{[Uvd,XYY - (a 2 U + UYY) v d ,x] v a + U u d ,x u a } dY =
= c i ( X ) A x + c 2 (X)A = 0,
(23)
whence A(X) = A ( 0 ) e x p ( - / o X g § § d x ) .
(24)
Without consideration of A(X) the total amplification factor of the instability would be N =
) is a nonlinear operator which typically is bilinear in the variable (p. Although the surfaces of homogeneity could be cylindrical or spherical surfaces, we focus on the planar case and adopt a Cartesian system of coordinates x,y,z with the ^-coordinate in the direction of inhomogeneity. Accordingly the operators of equations (1) may depend on z, but not on x,y or t. All considerations discussed in the following can be easily carried over to problems that are spherically or cylindrically symmetric. The stability of the basic state with respect to infinitesimal disturbances
(2)
can be assumed where r is the position vector and where the wave vector / has a vanishing z-component. The minimum value Rc for which the real part ar of the growthrate of the strongest growing solution of the linear problem vanishes is called the critical value and corresponds to a minimizing wave vector lc. We may choose the
19 ^/-coordinate in the direction of the vector lc which allows us to write the solution
the dimensionless deviation 0 of the temperature from the static profile of pure
26 conduction is needed for the description of convection. Since rigid boundaries with fixed temperatures are assumed at the top and bottom of the layer, the boundary conditions
= 0 = U = O a t z = ±(13) must be assumed and the expansion functions in representations (8) and (10) must be selected accordingly. Referring to the original work6 for details we describe here only some typical results. Knot convection represents the transition from convection in the form of thermal sheets, as in the case of roll like motions, to convection in the form of thermal plumes. Hot and cold fluid from the respective thermal boundary layer is collected into cylin drical columns of fluid that moves to the opposite boundaries. This process is evident in the left plots of figure 3. The collection of fluid through the finger like ejections from the thermal boundary layers as indicated in the upper right plot give rise to the typical knot like appearance of this type of convection in shadowgraph pictures 7 . The concentration of the buoyancy force into narrow columns leads to an acceleration of the fluid particles as they approach the other boundary and to a thinning of the opposite thermal boundary layer. This process permits a very efficient heat transfer as indicated in the lower right plot leaving a maximum area of the thermal boundary layer for the process of collection of fluid into the plumes. The stability analysis for steady knot convection is particularly simple in the limit of vanishing b and c?, since there are three symmetry properties available according to table 3. The stability equations thus separate into those for disturbances which share the respective symmetry with the steady solution and those which exhibit the opposite symmetry. Denoting by OSC, for instance, the class of disturbances with vanishing coefficients a/ mn for / + m -f n = even and with the properties a/mn = —a_/ mn , a/ mn = a>i-mn, we arrive at a total of 2 3 disturbance classes denoted by ECC, ESC, ECS, ESS, OCC, OSC, OCS, OSS
(14)
which can be analyzed separately. The letter E, 0,5, and C obviously stand for even, odd, sine and cosine like symmetry. Depending on the Prandtl number and the wavenumbers ax,ay of the steady knot solution, different classes of disturbances correspond to the highest real part crr of the growthrate cr. In the case P = 4, disturbances of the type ESC are the strongest grow ing ones. Their evolution to finite amplitude can be followed through time integration of the coefficients aimn(t) in the representation (10). Examples of such computations are shown in figures 4 through 6. Although there is already an increase in the heat
27
Fig. 4: Oscillatory knot convection with ax = 1.5, ay = 2.0 arising from the ESCtype instability in the case R = 36000, P = 4. Equidistant plots in time are shown in each column from top to bottom corresponding to a half period 0.020 of oscillation. Lines of constant vertical velocity in the planes z = 0 and z = —0.4 and isotherms in the same planes are shown (left to right). The x-coordinates run downward from 0 to 27r/ax.
28
Fig. 5: Isotherms (left) and lines of constant d
) given by (47a), provided / and a are replaced, respectively, by /* and new constant coefficients a*m. These coefficients satisfy (47b), provided / and am are replaced, respectively, by /* and a*m. The existence condition for the solution of the linear problem4 leads to the following result fl(0)(f ,77) = * (0) (Z,TJ) + &e + /z2£2 = 720
r
/
< 2/ + 1 ^ 1+ 462 /(/ + l)77j 2 +0(A ) + jU1e + Ai2e2>
(60)
where /?(0) is the Ravleigh number for the linear system, and it is assumed that /?(0)(/,7j) can differ from /T0)(l*,il) only by two terms of orders eande 2 designated by faemd/J^e2, respectively. It turns out that for non-self-adjoint case only the term fJ^ed^ * 0) need to be considered, while for self-adjoint case only the term /Mle2(fil = 0) is effective ( ^ * 0). The minimum Re as a function of / is achieved by 1 = 1 (77 > (77/17)*), 1 = 2 [(77/17)* > 77 > (77/68)*], = 3 [(77/68)* > 7] > (77/170)*],
(61)
and so on, by increasing values of / as r0 increases for a fixed C. It is seen from (61) that if 77 satisfies an inequality in (61), then a single value of / niinimizes Ri0\ However, if 77 is sufficiently close to an equality in (61), then the difference between the values of #(0) for two neighboring degrees /and/* can become sufficiently small (of order fixe + fx2e2 or less) and we obtain the singular case which is presented in this section. Let us first consider the non-self-adjoint case33. The solvability conditions at order 2 e yield
4*(,)a„ = AX Xa.^wZ-WW) n,p=0
138
+AlM2 j^aX^W^W^W^),
(62a)
n,p=o
A.R^al = 2\AJA, ^ajx^WJT^W^
(62b)
n,p=0
where angular bracket denotes average over the spherical surface and Aff.(l,2,3) denote some constants which depend only on / and /*. The following variational functional are then determined from Eq. (62)33: (R - rt^ = Nrf + A ^ 2 ,
(63a)
(R - R;)^ = IN^A,,
(63b)
Where Rt andRr designate /?(0)(/,7]), and /?(0)(f,7j), respectively, and the coefficients N^i = 1,2,3) contain information from Eqs. (62) which deals with the patterns of the bifurcating solutions. The fact that in the case of the functional (63) two wavelengths participate instead of one in the case of simple patterns permits larger values for the extremum. The effect of the combination of two wavelengths is most pronounced in the case / = 1, /* = 2 where the wavelengths differ by a factor of 2. The extremalising solution is axisymmetric in this case and is given by \Wt = (A/7/4)P2°, A2Wr = l ^ 0 , R* = V7/5,
(64)
where R* denotes the extremum value of the functional (63). This solution describes a strong rising motion at one pole and a broad descending motion spread over the major part of the spherical surface including the other pole and the equatorial region. For the case / = 2, /* = 3, it is found that a three-cell pattern extremalizes (63): Wi =i(7/3)*A°. AiWr =!(!)*A 3 cos30, R* =f > /f.
(65)
The four-cell pattern is obtained for / = 3, 1* = 4: A ^ . = ( f ) * % cos20, \Wt = ^ [ P 4 ° - (f fP44 cos40], /?*=V91/11.
(66)
This pattern exhibits the symmetry of a tetrahedron. A seven-cell pattern becomes realized when we consider the case / = 4, /* = 5: AlWl = V77/150P4°, A2Wr = V73/150 F55 cos50, R* = 2-V462/65.
(67)
In contrast to the realized patterns discussed above, the seven-cell pattern is no longer
139 symmetric with respect to each of the cells. This property will become the rule as the preferred patterns are sought for higher values of /. In the asymptotic case of large /, /* the preferred pattern will consist of a network of hexagonal cells with some pentagonal cells interdispersed33. Let us now consider the self-adjoint case 34 ' 35 ' 78 . The solvability conditions at order e3 yield the expressions for /? (2) which are lengthy34 and will not be given here. In general, several solutions of the form (59) satisfy the solvability conditions. To distinguish the physically realizable solution(s) among all the steady solutions which satisfy these conditions, the stability of these steady solutions with respect to arbitrary three-dimensional disturbances are investigated, and the stable solutions are determined for / = l(f = 2) 3 4 . It is of interest to note that the solvability conditions for the steady solution and for the stability equations have also been derived from an extremum principle34. A functional was discovered 34 whose stationary value(s) correspond to steady solutions of the nonlinear system and those solutions for which the stationary value is a maximum are stable. For the self-adjoint case only the cases / = i(f = 2) and / = 2(/* = 3) have been investigated so far. For 1 = 1, /* = 2, there are three distinct mixed solutions. Solution I is an unstable axisymmetric solution given by a\ = a*l = 1, ax = a* = a2 = 0.
(68)
Solution II (so called tennis ball solution33) is given by <x\ = a\ = 1, ax = a*0 = a* = 0.
(69a)
It is stable if ^ > 0.0017 faf A}.
(69b)
Solution El is given by a\ = a*\ =1, ccx = a*0 = a2 = 0
(70a)
and is stable if -O.OlOmTj 2 ^ 2 < ju2 < -0.00407XrfAt.
(70b)
For / = 2, /* = 3, there are five distinct mixed solutions35. No stability analysis for these solutions have yet been done, but it is found that the following non-axisymmetric solution carries the maximum amount of heat among all the other solutions and thus appears to be preferred31. <x\ = a\ = 1, a0 = ax = a\ = a\ = a\ = 0.
(71)
140
7.
Flow During Solidification
In this section we consider nonlinear stability analysis for convective flow of the melt during directional solidification for small segregation K «1. Systems with K < 0.01 have great practical interest since they become nearly pure substances upon solidification. We analyze the problem in the presence of either rotation or magnetohydrodynamic effects which were shown by Riahi47"49 to be able to stabilize the solution to the nonlinear stability system. Such solution is known to be unstable in the absence of such external We consider the problem of solutal convection in a horizontal layer of a binary alloy melt, rotating about a vertical axis with constant rotate rate £2 and permeated by magnetic field B, from which a semi-infinite slab of crystal is being grown. The magnetic field is assumed to be composed of a uniform external field \B2 y+ B3 z in plane of yoz and a perturbation field b*>. The governing hydromagnetic system orequations and boundary conditions for the flow velocity w, magnetic field and the solute concentration, under the usual Boussinesq approximation, are considered in a coordinate system moving at velocity V0 (crystal growth rate) with the solidification front at the vertical variable z = 0. These equations and the appropriate boundary conditions are given in Riahi^-**. They are lengthy and will not be repeated here. We shall make our analysis here specific and simpler by considering zero rotation case. In the above description of the problem, y and z are unit vectors along horizontal axis y and vertical axis z, respectively. Non-dimensional variables are used in which length, velocity, pressure, time, magnetic field and solute concentration are scaled, respectively, by d/V0, V0, pV02, d/V?, B3 and (1 - K)CJK. Here d is the diffusivity coefficient of solute, p is a constant reference density, K = CslCm> is the segregation coefficient, Cs is the solute concentration in the solid at interface z = 0, Cm is the solute concentration in the melt at z = 0, and C is the solute concentration in the bulk of the melt. It turns out that C = Cs in the present problem. Another assumption to be made here is that the adjoining medium at the upper boundary (z = oo) is a non-conductor medium, while the crystal is a high conductor. The non-dimensional parameters of the nonlinear stability system, which is the above described hydromagnetic equations and the boundary conditions, are the ratio B23=B2/B3,K, the solutal Rayleigh number R = pg{\-K)C00d2l(idKV*\ the Chandrasekhar number Q = ji!%d>faJmV*)9 theeiffusi^^ field to solute concentration and the Schmidt number Sc = jn/d. Here p is the fractional change in density due to a change in the solute concentration, gis acceleration due to gravity, ju is the kinematic viscosity, and ft is the magnetic permeability. The governing nonlinear stability system, in the limit of Sc = oo, can be converted into the following form: A2(V40-?C) = -QT
L
+ BB232:^-\E IE+MSE-VSEYA ;A A 22V V*2[ D D+ + S(SE'VSEYZ[ L V dy) ~\~ ~ ) ~J
(72a) (72a)
141
A2 fTV4 + V 2 D - | - V 2 l £ + V2fD + ^ | - l o j = 8ld<&V5E-8EV8®YzJ (V2 + D - %)C - exp(-z) A2
(72b) (72c)
O = D
(72d)
|4>|(oo, |£|(oo, c = 0 as z -» oo.
(72e)
Here D = d/dz, C is the deviation of solute concentration from the basic concentration C0, and <X> and E are the Poloidal components of u and b given by (w,&) = 5(0,£), S = V x V x z .
(73)
As in the non-magnetic field case45, the toroidal components of u and b are insignificant48 and are not included in (72)-(73). The simplification of the limit of Sc = <*> is assumed here since the results given in Hurle et al.39 for the regime of practical interest Sc »1 indicate that the limit of Sc = ©° is expected to be achieved so long as Sc > 10. First the linear stability analysis of (72) is considered by using the infinitesimal perturbation method: as was used by Riley and Davis45 for the case without magnetic field. We analyze the linear version of the stability system (72) by using normal modes in the horizontal variables x and y and in time t and by assuming that the critical wave number approaches zero as K —»039*40. We, thus, assume that JC, v and t dependence of the dependent variables is like expf at + iJC- r I, where r = (JC, V),
(74a) (74b) (74c)
and R0 = R(a = 0). The results (74) imply that o~ is real and independent of B23 and that it is impossible to have positive a for r < 1. Hence, all the solutions decay to zero for r < 1 as t increases. The non-trivial solutions, thus, correspond to cases where r > 1. It is also of interest to note that the condition of thermal diffusivity dt less than 7] is amply
142
satisfied for terrestrial materials80. Also dt » dso. Hence we shall assume the condition that x »1. For marginal stability (74a) yields RO = M{Q,T)
(75)
and the linear version of the evolution equation, to be introduced later in this section, leads us readily to the following expression for R: R/Ro = Ky*[Ky* + //£*) + 0(a 3 ),
(76)
where the expression for / is quite lengthy47 and will not be given here. Minimizing R given by (76) with respect to K% then yield Re = Ro{\ + 2K*) + 0(Ky<), ccc = (K/tf
(77)
We find that both Re and ac increase with Q, but they are independent of B^. Following Riley and Davis45 for the nonlinear stability system for small K9 we let K = e2KltR = R0(l + ey),
(78)
where 0 < e « l . The linear stability analysis discussed above suggests to introduce rescaled variables (jc , ,/) = eK(jc,v),z' = z,r / = e2r, (<*>',£') = £(,£), C' = C.
(79)
These rescaled variables and parameters are then used in (72) and the primes are dropped for simplifying the notation. Next, we seek a solution in terms of a series in powers of eYl:
(O.E,C) = X ^ ^ . ^ . C ^ ) .
(80)
Using (80) in the rescaled form of the nonlinear system, we find the following solutions in the orders e, em and e2: *i = % = %2 = £3/2 = 0, ( C ^ ) = [A(x,v,r), B(x,y,r)]exp(-z), (
(81)
where the expressions for the functions F2, H2, and G2 are given in Riahi48, and A andB are two amplitude functions. Integrating Eq. (72c), in the order e2, in the vertical direction from z = 0 t o z = «> and using (81) yield (75). In the limit of T » l , (75) yield
143
Q*2f. Ra0=2[l + Q^.
(82)
It is seen from (82) that R0 increases with Q but is independent of B23. For Q -> 0, the expression for R0 approaches the value 2 in agreement with the critical solute Rayleigh number computed in the limit K -* 0 and Sc -> ~ by Hurle et al.3 for the problem with no magnetic field. In the order e5/2(72) yield solution48 which depend on B and f. Integrating Eq. (72c) in the order e5'2 in the vertical direction from z = 0 to z = « and using (81) and solution in the order e5/2 together with the boundary conditions for C5/2, yield
^-o= °-
((83) 83)
dA
Hence, the amplitude function A does not vary with respect to y. In the order e3 Eqs. (72a,b) yield solution^ which depend on A and f . Integrating (72c) in the order e3 with respect to z from z = 0 to z = ~ aad ussng ((81 )nd solution in the order e3 together with the boundary conditions for C3, yield the following nonlinear evolution equation for the amplitude A: „„K A
dA dA
dd22AA
,d ,d44AA
11 dd
[|(^)] + , A f = 0, W*'h*4-°-
(84) (84)
where the lengthy expression for the coefficient / is given in Riahi48 and lx is a constant. Using (83), (84) together with appropriate boundary conditions for f atjc = ±~ and }> = ±oo° yield ^ = 0. dy dy
(85)
The results (83) and (85) imply that disturbance are two-dimensional in (x,z) plane. Thus, presence of a horizontal magnetic field (along y-axis) eliminates three-dimensional disturbances. Next, we reduce the number of coefficients in (84) first by choosing y = 1 which defines e for a given R. Then set 1/2
^ l==S/l,t 5//,r = = f/,x = Jc(/) . K tl,x=x(lf-
(86) (86)
A + A^ Aiiii++ 0.5(A 0.5(A2)H 0, A;!+SA + SA + + AH + )H == 0,
(87)
Using (85-(86), (84) becomes
where a subscript iorx means /<# or d/dx. Equatton (87) )i effecttvely the same amplitude equation as that derived by Riley and Davis45 for non-magnetic field case where S equals a constant times Kx. Here it is found that S increases with Q but is independent of B22. Thus presence of a vertical magnetic field leads to an effective segregation coefficient 5 given by
144
S = Kl/e\
(88)
larger than the physicochemical value K by an amount which is determined by R and Q. Young and Davis44 considered Eq. (87) in the non-magnetic field case subject to the periodic boundary conditions in the internal 0 < x < Jc0. They solved equation (87) using an explicit finite difference scheme for various values of x0 chosen as multiples of the wavelength 2->/2;r which gives maximum amplification in the linear theory of Eq. (87). This linear theory give S < 0.25 for instability. Figure 7 in Young and Davis44 presents the calculated A for S = 0.2, 0.24 and 0.4. When S > 0.25, initial disturbances decay to zero and the zero basic state value (A = 0) is regained. For S < 0.25, the system is unstable and a cellular structure forms. For S « 0.25, the results reported in Young and Davis 44 indicate an apparent secondary instability where the tips of the cells show a tendency to split. It is clear from these results the presence of a strong vertical magnetic field can stabilize the system, as is the case in the presence of a strong vertical rotation49, and can inhibit the onset of the cellular structure by effectively magnifying the segregation coefficient. The case of presence of combined rotation and magnetic field is presently under investigation by Riahi81. It is of interest to note that the following evolution equation for the function B is obtained from the order e1/2 system by applying the same procedure as that applied to the order e3 system which determined the evolution equation for A: Bt+SB + Bn + B^+lAB)-^.
(89)
This equation is linear in B. For S < 0.25, A * 0 and the solution for B is not needed since A is the leading order amplitude. For S > 0.25, A = 0 and Eq. (89) yields solution which decays to zero. 8. Some Concluding Remarks We reviewed nonlinear stability analysis for a number of convective flow problems dealing with the preferred patterns in horizontal layers and in spherical shells, stability region of realizable solutions with different wave-lengths, steady and oscillatory solutions, simple and mixed type solutions, effects of external constraints and solidification, and for cases described by discrete modal and wave packet representation of the flow solutions. We found that certain mathematical methods, such as perturbation, multiple scales and scaling procedure are quite effective to carry out the analysis for the nonlinear system. While the results based on the analyses presented in this review article are general in the sense that they require only the horizontal symmetry of the basic state in the horizontal layer cases and the spherical symmetry of the basic state in the spherical shell cases, they are limited in most cases by the condition that only small deviations from the basic state are permitted. But since most of the results are based on the analysis that is mainly concerned with symmetry properties, those results are hardly affected by the condition of small e. The symmetry of the patterns based on the full nonlinear stability system cannot be changed by the contributions of higher order in e unless further bifurcation takes place. Because of
145
the exceptional symmetry of such preferred patterns it is expected that those patterns retain their form and distinction for an extensive range of R. Although there have been very few studies so far on the problems of wave packet convection patterns as well as problems of convective flows influenced by the boundary modulation effects, the results of the nonlinear stability analyses presented here indicate that the importance of such problems should not be under estimated. Wave packet convective flows, which are generally dispersive65, provide a larger manifold of realizable patterns due to solutions of initial value problems, while a discrete modal approach often fails to provide such manifold. Convective flow models with boundary corrugation or modulation effects can provide a large manifold of realizable solutions and can provide an operating procedure to control instabilities and the flow structure. The results of nonlinear stability system for convective flows in spherical shells reviewed here has been based on the analyses of much more general subject of solutions generated by bifurcations from spherical symmetric basic states which has received much attention in the recent literature. The problem of degeneracy and its removal by nonlinear effects dealt in the analyses is common to all eigenvalue problems with spherical symmetry. Related problems which can be investigated using the analysis of the type used here include buckling of spherical shells, cracking patterns of symmetrically cooling solid spheres and oscillations of non-rotating stars.
146 References 1. P.G. Drazin and W.H. Reid, Hydrodynamic Stability (Cambridge Univ. Press, Cambridge, U.K., 1980), p. 527. 2. W.V.R. Malkus and G. Veronis, / . Fluid Mech., 4, (1958) 225. 3. A. Schluter, D. Lortz and F.H. Busse, / . Fluid Mech., 23, (1965) 129. 4. F.H. Busse and N. Riahi, J. Fluid Mech., 96, (1980) 243. 5. F.H. Busse, / . Fluid Mech., 30, (1967) 625. 6. N. Riahi, / . Fluid Mech., 152, (1985) 113. 7. N. Riahi, ZAMP, 31, (1980) 261. 8. F.H. Busse, In Instability of Continuous Systems, ed. H. Leipholz (1971) 41. 9. N. Riahi, Development in TAM ,11, (1982a) 357. 10. N. Riahi, / . Phys. Soc. Japan, 51, (1982b) 4104. 11. N. Riahi, Proc. Symp. Fluid Dyn., (Urbana, EL) (1984a) 315. 12. D.N. Riahi, Proc. Conf. Math Appl. Fluid Mech. and Stability (Troy, NY) (1986a) 257. 13. D.N. Riahi, Proc. AIAA/ASME Heat Trans, and Thermophys. Conf. (Boston, MA), 56 (1986b) 41. 14. D.N. Riahi, Acta Mech, 60, (1986c) 143. 15. N. Riahi, / . Phys. Soc. Japan, 53, (1984b) 4169. 16. D.N. Riahi, Int. J. Non-linear Mech., 21, (1986d) 97. 17. D.N. Riahi, / . Math. Phys. Sci., 22, (1988a) 161. 18. D.N. Riahi, Acta Mech., 64, (1986e) 155. 19. D.N. Riahi, Acta Mech.,71, (1988b) 249. 20. D.N. Riahi, Proc. SECT AM XIII (USQSC), (1986f) 91. 21. D.N. Riahi, J. Phys. Soc. Japan, 56, (1986g) 3515. 22. D.N. Riahi, CAM Conf on Wave Phenomena (Edmonton, Canada) (1992a) 145. 23. D.N. Riahi, / . Fluid Mech., 246 (1993a) 529. 24. D.N. Riahi, TAM-Report #757 (TAM Dept., UIUC, Urbana, IL) (1994a). 25. D.N. Riahi, Proc. Royal Soc. London series A, (1995a) in press. 26. D.N. Riahi, Bifurcation Phenomena and Chaos in Thermal Convection, ASME, HTD-Vol. 214/A MD-Vol. 138 (1992b) 69. 27. S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, (Clarendon Press, Oxford, U.K., 1961), p. 654. 28. F.H. Busse, J. Fluid Mech., 72, (1975) 67. 29. F.H. Busse and N. Riahi, / . Fluid Mech., 123, (1982) 283. 30. F.H. Busse and N. Riahi, / . Tera Cognita, 4, (1984) 240. 31. N. Riahi, / . Phys. Soc. Japan, 53, (1984c) 2506. 32. N. Riahi, G. Geiger and F.H. Busse, Geophys. Astrophys. Fluid Dyn., 20, (1982) 307. 33. F.H. Busse and D.N. Riahi, / . Nonlinearity, 1, (1988) 379. 34. D.N. Riahi and F.H. Busse, ZAMP, 39, (1988) 699. 35. D.N. Riahi, Proc. SECT AM XVII, (Hot Springs, AK., USA) (1994c) 469. 36. S.H. Davis, / . Fluid Mech., 212, (1990) 241. 37. M.R.E. Proctor, / . Fluid Mech., 113, (1981) 469.
147 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65.
66. 67. 68. 69. 70. 71. 72. 73. 74. 75.
C J . Chapman and M.R.E. Proctor, / . Fluid Mech., 101, (1980) 759. D.TJ. Hurle, EJakeman and A.A. Wheeler, Phys. Fluids, 26, (1983) 624. D.N. Riahi, Phys. Fluids, 31, (1988c) 27. S.H. Davis, U. Muller and C. Dietsche, / . Fluid Mech., 144, (1984) 133. D.N. Riahi, Phys. Fluids A, 3, (1991) 2816. D.N. Riahi, Int. J. Engr. Sci., 27, (1989a) 689. G.W. Young and S.H. Davis, Physical Rev. B, 34, (1986) 3388. D.S. Riley and S.H. Davis, Physica D., 39, (1989) 231. D.N. Riahi, Phys. Rev. B, 44, (1992c) 4170. D.N. Riahi, / . Math. Phys. Sci., 26, (1992d) 429. D.N. Riahi, Proc. ASME Symp. on Bifurcation and Chaos (Anaheim, CA), Vol. 214/A MD-Vol. 138, (1992e) 93. D.N. Riahi, Acta Mech., 99, (1993b) 95. D.N. Riahi, Int. J. Engr. Sci., 30, (1993c) 551. R.E. Kelly and D. Pal, Proc. Heat Transf. and Fluid Mech. Inst. (Stanford University Press, Stanford, CA., 1976), 1. R.E. Kelly and D. Pal, / . Fluid Mech., 86, (1978) 433. D.N. Riahi, / . Fluid Mech., 129, (1983) 153. D.N. Riahi, Phys. Fluids A, 2, (1990) 353. D. Pal and R.E. Kelly, Proc. 6th Intl. Heat Transf. Conf. (Toronto, Canada, 1978) 235. D.N. Riahi, / . Abstracts (AMS) (1995b) 398. D.N. Riahi, submitted for publication (1995c). N. Riahi, / . Aust. Math. Soc. (Series B), 25, (1984d) 406. D.N. Riahi, Bull. Am. Phys. Soc, 34, (1989b) 2284. D.N. Riahi, Proc. Royal Soc. London Series A, 436, (1992f) 33. D.N. Riahi, Proc. of Pan Am. Cong. ofAppl. Mech. (Brazil), (1993d) 327. D.N. Riahi, Proc. of First World Cong. Nonlinear Analysis (Tampa, Fl) (1995d) in press. A.C. Newell and J.A. Whitehead, / . Fluid Mech., 39, (1969) 279. G.B. Whitham, Linear and Nonlinear Waves (Wiley, NY, 1974), p. 636. D.N. Riahi, Nonlinear Dispersive Waves: The Math. Theory with Applications to Fluids and Plasmas ed. L. Debnath , (World Sci. Publishing Co., 1992) (1992g) 417. G. Kuppers and D. Lortz, / . Fluid Mech., 35, (1969) 609. G. Kuppers, Phys. Lett. A, 32, (1970) 7. R.M. Clever and F.H. Busse, / . Fluid Mech., 94, (1979) 609. N. Riahi, ZAMP, 33, (1982c) 81. J. Tavantzis, E.L. Reiss and B J . Matkowsky, SIAM J Appl. Math., 34, (1978) 322. F.H. Busse, Rep. on Prog, in Phys., 41, (1978) 1929. D.D. Joseph and S. Carmi, / . Fluid Mech., 26, (1966) 769. G. Geiger, Ph.D. Dissertation (University of Munich, 1977). P. Chossat, SIAM J Appl. Math., 37, (1979) 624. M. Golubitsky and D.G. Schaeffer, Commun. Pure Appl. Math., 4, (1982) 500.
148
76. D. Thompson, On Growth and Form (Cambridge University Press, Cambridge, U.K., 1942), p. 250. 77. W.T. Koiter, Proc. Koninklijke Nederland Akad. Van Wetenschapen, B72, (1969) 40. 78. D.N. Riahi and F.H. Busse, Proc. ASME Winter Meeting (Boston, MA) (1987) 109. 79. L. Hadji and D.N. Riahi, Proc. SECT AM XVII (Hot Springs, AK) (1994) 444. 80. S.R. Coriell, M.R. Cordes, W.J. Boettinger and R.F. Sekerka, / . Crystal Growth, 49, (1981) 13. 81. D.N. Riahi, In preparation (1995e).
149 MODELING AND SIMULATION FOR PRIMARY INSTABILITIES IN SHEAR FLOWS
D.N. RIAHI Department of Theoretical and Applied Mechanics 216 Talbot Laboratory, University of Illinois 104 S. Wright St., Urbana, IL 61801
ABSTRACT
This chapter reviews primary hydrodynamic instabilities and their mathematical modeling and numerical simulation for several important shear flow problems of practical importance. The specific problems considered here deal with instabilities which may take place in a number of shear flow problems such as bounded and unbounded mixing layers, rotating disk flows, flows over wings and boundary layers. Modeling is accomplished by making use of certain mathematical methods such as perturbation techniques, multiple scales and method of stationary phase. Computation is done using finite differences, spectral, collocation and Galerkin methods. In addition, the latest research results, for each particular shear flow problem, accomplished by the author, are also presented.
1. Introduction This chapter deals with a review of the fundamental understanding of the primary instabilities of a number of shear flows which are of importance in relation to practical applications of such flows. A number of primary instabilities of the shear flows lead to completely different form of such flows with different behavior and in certain cases can facilitate transition to turbulence. We shall present in this chapter a review of some of the mathematical modelings and numerical simulations that have been done for the primary instabilities which may take place in bounded and unbounded mixing layers, plane wake flows, rotating disk flows, boundary layers, flows over wings and in channel flows. Lie and Riahi1 investigated linear spatial primary instabilities of viscous flows in both bounded and unbounded mixing layers for arbitrary values of the ratio A between the difference and the sum of the velocities of the two flow streams. They applied spectral and Galerkin methods with expansions in Chebyshev polynomials and some mathematical analysis and modeling to determine the critical values of the Reynolds number at which the primary instability arises first in the mixing layers. They detected new modes of viscous instability, and they studied the effects of the presence of solid boundaries on the primary instability processes. Earlier, there was some studies on the linear temporal viscous primary instability of unbounded mixing layers for X = l 2 . However, it is known 3-5 that the spatial instability is generally preferred over the temporal instability in mixing layers, particulary for the primary instability and the early development of the mixing layers which are mostly influenced by the spatial events. The parameter A also plays a significant role in
150 mixing layers. For example, it can have significant effects on the critical Reynolds number for the onset of secondary motion1, and the spreading rate of the mixing layers decrease with A 6 ' 7 . Numerical simulations in mixing layers are particularly efficient if they are based on spectral methods 1 with expansions of disturbance variables in Chebyshev polynomials which require less computer time and storage to achieve accurate results when compared with those required by finite difference methods and is known to lead to accurate results in other problems as well8. More recent modeling search for mixing layers9 led to the conclusion that a mathematically rigorous approach for a time evolving base flow in a mixing layer seems to be the singular value decomposition of the linear operator transforming some initial perturbations into the solution at a given later time. Another important type of free shear flow problems are those of plane wake flows10. In a series of papers, Kuah and Riahi 11-13 reported results for the computational investigations of primary instabilities in plane wakes behind a flat plate parallel to the upstream uniform flow. Such shear flow system is an important flow in practice and, in particular, is of interest in aerodynamics and Naval Sciences. Just as in the mixing flow cases discussed in the previous paragraph, an improved fundamental understanding of the primary instabilities and early transition processes would open significant possibilities to the design engineers to delay transition and control or promote turbulence in such shear flow systems. The results reported in the above references 1113 are based on viscous modes of instabilities which are in contrast to most of the investigations of plane wakes due to in viscid modes of instabilities14-18. Boundary layer flow over a rotating disk is an important prototype shear flow system to study the fundamental aspects of the cross flow instability mechanism which is operative in many three-dimensional boundary layer flows with aerodynamic applications. There have been notable recent contributions on the rotating-disk primary instability aspects both experimentally19 and theoretically 20-23 . All these theoretical studies were based on linear stability theory. Very few nonlinear investigations have been done to date. Malik24 used a spectral method in a Navier-Stokes simulation and found some realistic mode interactions similar to those found experimentally25 in a swept wing flow. Itoh26 predicted the supercritical stability of rotating disk flow using a weakly nonlinear model based on the Orr-Sommerfeld equation. MacKerrel 27 used an asymptotic analysis as the Reynold's number R approaches infinity to find that nonlinear effects can be destabilizing for some small wavelength modes near the lower branch of the neutral stability curve, although these modes may not be preferred at such large R. More recently Vonderwell and Riahi 28 carried out a rigorous weakly nonlinear study based on the full governing equations of the preferred disturbances for the rotating disk primary instability problem. They found, in particular, that subcritical bifurcation exists in rotating disk flow, at least for the preferred primary instability modes near the minimum of the upper branch of the neutral stability curve 2 2 . It should be noted that the above study was based on the assumption that disturbance consists of a monochromatic wave with constant amplitude for all the horizontal variables. More realistic is the consideration of localized disturbances whose behavior can be studied by using the wave packet approach 29-31 . Riahi and Vonderwell32 investigated wave packet development of primary instability modes of rotating disk flow using the method of multiple scales. The system of equations for small amplitude disturbances was solved by using the weakly nonlinear theory to determine the wave packet solution which represented the time and space variations of the wave pattern. The
151 amplitude of disturbance satisfied a nonlinear partial differential equation, a generalization of the time dependent equation and of the linear model of earlier works30"32. The solution of the amplitude equation was found to become more and more concentrated at the center of the wave packet disturbance becoming infinite there in a finite time. This bursting type behavior suggested possible fast transition to turbulence in rotating disk flow. A general weakly nonlinear and non-parallel boundary layer theory for primary instabilities, which include those in two-dimensional Blasius flows, three-dimensional flows over infinite and finite span wings, was developed recently 33-34 . This theory is an extension of an earlier spectral theory 35 and provide dominant contributions to the nonlinear terms by subharmonic components of primary instability modes at various frequencies. The results are consistent with experimental observations5*36. Recently some progress has been made on fundamental understanding of the roles and mechanism played by the primary instability modes and their interactions in threedimensional compressible boundary layer flows over swept wings 37-38 . Asymptotic and triple-deck investigation of the primary modes of instability for such flows37 indicated that there are basically two kinds of modes. Inviscid modes having wavelengths scaled on the boundary layer thickness which are relevant upper branch modes. Wall modes having a triple deck structure which are relevant lower branch modes. Stationary and long time scale non-stationary modes are preferred. Compressibility effects lead to several types of primary wall modes which depend on particular disturbance density and temperature scaling. Value of the wave number for the wall modes was found to depend critically on the wall values of the mean flow velocity gradients, density and temperature as well as on the wall value of the viscosity and the viscosity gradient, while value of the wave number for the inviscid modes depends on the mean flow quantities throughout the boundary layer. Primary instability modes that satisfy the triad resonance conditions were investigated using collocation and multiple scale methods 38 . The most critical resonant triads were determined, and the roles played by the primary modes and their interactions in evaluating the compressibility effects were detected. Primary instability of two and three dimensional rotating or non-rotating channel flow modes can be of algebraic and exponential linear or nonlinear types for non-rotating and rotating flow cases, respectively 39 . The primary instability modes found for the rotating system is stronger, but less probable, than the corresponding one which occurs for the non-rotating system. 2. Mixing Layers We shall first review and discuss the primary instabilities in the confined mixing layers 1 . The well known tangent-hyperbolic profile 2'40 u(y) is often used as the basic flow profile for a mixing layer generated by two parallel streams with different velocities ux and u2 and with an assumed arbitrary velocity ratio A = (ux - u2)/(ul +u2) for ul>u2. Here v is the cross stream (transverse) variable. Guided by previous inviscid instability results 40 , y = ±5 can be chosen to represent the upper and lower limits of v of the flow field (base flow plus the primary instability modes). As a computational modeling approach to the problem, spectral method, used earlier by Orszag for the channel flow8, can be applied along with the expansions in Chebyshev Polynomials 41 of the eigenfunctions for the Orr-Sommerfeld equation to be discussed below. This suggests that
152 u and its second derivative u" with respect to v which are needed in the governing equation for the primary instability modes to be discussed below can be expressed in terms of Chebyshev Polynomials Tm(y/5) as «(y) = l + i 5 „ r . ( y / 5 ) ,
(la)
m=0
«"(>) = i t t f r / s ) , M<|
(ib)
m=0
where am and a^ are constants for a given A. Here a prime (') represents derivative with respect to y. The stream function of the primary spatial instability mode is assumed of the form 0(y)exp[/(ccc - wtj\ in terms of a complex wave number a , phase velocity c = w/a, real frequency w, eigenfunction 0(y), streamwise variable x, pure imaginary number / = V-T and time variable t. Since Squire's theorem 42 holds only two-dimensional modes of instability are of interest here which are the most critical ones. Also only spatial instability modes are considered here due their dominant effects in the early development of free shear flows 1 ' 42 . Using the above description for the stream function of a primary instability mode in the Navier-Stokes equations and eliminating the pressure terms with the aid of the continuity equation for the incompressible fluid flow cases leads us to the so-called OrrSommerfeld equation which is given below 0(/v)(y) - 2 a 2 0 " (y) + a > ( y ) = iaR {[u(y) - c\t"(y) ~ a2(/>(y)] - u"(yMy)}
§
(2)
where R = AuO/n *s t n e Reynolds number, Au = ul-u2, 0 = A M / ( 4 M ^ ) (momentum thickness40) and JLL is the kinematic viscosity. The Eq. (2) together with (1) indicate that
+ (a + K2)(
(iga)
and
(18 b)
Using (14 c) and the approximate expressions for fn and gln in (18) yield relation between f0 and gl0 for v/5 = ±1. Next, this relation is used in (14 c) and then take derivatives of (14 c) with respect to v. After some manipulation the results for the approximate forms for 0 and
(19)
Here only the last four entries in the array of B are nonzero which are due to the conditions for 0 and <j>' and y/5 = ±1. The numerical procedure for modeling cases described so far is presented in details elsewhere1. Here we briefly review only the main aspects of the numerical procedure. The coefficients am and a{2) introduced in (1) which affect the solution to (2) are computed by applying the particular Gaussian quadrature formula1 to (1). By reducing R gradually and applying to condition (7), one can locate the critical Reynolds number, Re, which is achieved for a particular ar = real (a), for given A, by checking that at = imaginary part of a equals zero. Searching process for Re can be achieved by using a Newton-Raphson scheme1. The results of the numerical computations1 indicate that the primary instabibty mode at the onset of the secondary motion holds the same character for both bounded and unbounded cases, although Re for unbounded case is much smaller than the corresponding one for the bounded case. The effect of boundaries is essentially to restrain the growth of the primary instability, so that Re becomes very small for the unbounded cases. In both bounded and unbounded cases, the wave number and frequency of the critical instability mode are very close and they held essentially the same character. It was found that the
156 primary viscous instability modes are more critical and unstable than the corresponding primary inviscid instability modes, and the most dangereous primary instability modes are intrinsically viscous in nature and are due to intermediate values of the frequency w. 3. Rotating Disk Flows We consider the problem of incompressible fluid over an infinite horizontal plane rotating about a vertical axis with angular velocity Q. It is convenient to use cylindrical coordinates r*,0,z\ with z = 0 being the plane of the disk. Let the over-bar variables P and U = («,v,vv) denote the primary steady base flow quantities for pressure and velocity in the radial (r*), azimuthal (0) and axial (z*) directions, respectively, in the rotating coordinate system. The well known Von-Karman's exact solution of the NavierStokes equations43 is obtained by setting u = r*QF(z% v = r*ClG(z), w = (MG)* #(*)» p = p/*lP(z),
(20a,d)
where z = z*(£l/jj) • Here ji is the kinematic viscosity and p is a constant reference density. The functions F , G , / / a n d P satisfy a well known system 43 whose solution is easily determined. To study the stability of such solution with respect to super imposed small disturbances, we first non-dimensionalize the governing Navier-Stokes equations by using (fi/tif* as the reference length, r*Q as the reference velocity, and P 2 .Q 2 as the reference pressure, where r* denotes the radial location near which the analysis is made28. The primary instability mode variables w,v,n\ and/? superimposed on the base flow variables (20) are used to satisfy the Navier-Stokes equations in cylindrical coordinates. The resulting equations for the primary instability mode variables, as functions of nondimensional space and time variables (r,0,z) and t can be simplified by applying the usual quasi-parallel flow approximation22. Here we are interested in the stability near a radial location such that the Reynolds number, R = r*(Q.//i) , is the minimum Reynolds number, Re, below which no stationary primary mode is amplified. The quasi-parallel approximation, then, amounts to substituting Re for r, since r* corresponds to r = R. Further simplification is obtained by neglecting terms of order R~5 and smaller. The final equations for the primary instability modes are lengthy28 and will not be given here. Next, the method of multiple scales is applied. It is assumed that R does not deviate much from Re and a slowly varying time variable ts is defined ^(fl-jg/*,,
(21a)
ts = e\
(21b)
where e is a small perturbation parameter and /^ is a constant. The sign and magnitude of #2 is to be selected so as to insure that e2 remains small and positive. It is then assumed that the solution to the governing system for the primary mode varies on two different time scales such that
dt
dts
(22)
157 Applying the weakly nonlinear theory 44-45 , we pose expansions in powers of e of the form F = e{EAFn) + e2(E2A2F22+\A\2F20) + £ 3 (£ 3 A 3 F 33 + |A| 2 EAF 3 1 ) + C.C,
(23)
where F is any primary mode variable, E = exp[/(ar + fiRcO - wf)], A = A(ts), \A\2 = AA where the over-bar indicates complex conjugate, and c.c. indicates the complex conjugate of the entire preceding expression. Next, use (23) in the governing system and collecting coefficients of powers of e leads to the systems of orders e\ e2 and £3 whose complete forms and solutions are discussed in details in Vonderwell and Riahi28. An equation for A is obtained by applying an existence condition28 for the solution of the order e3E system. The equation for A and its complex conjugate are then simplified to ^ - = M + c|A|2A,
(24 a)
-^- = bA + c\A\2A. dts
(24 b)
where b and c are constants whose complete expressions are given in Vonderwell and Riahi 28 . Multiplying Eq. (24 a) by A, Eq. (24 b) by A, and summing yields a Landautype equation42
J-|A|2=2/>,|A|2+2c,|A|\
(25)
where twice the real parts of b and c, 2b r , and 2cr, are recognized as the linear growth rate of \A\ , and the negative of the Landau constant42, respectively. A Chebyshev collocation method was used to solve the systems for the base flow (20) and for the primary mode in the orders £ and e2 as well as the adjoint system to the linear system for the primary mode whose solution were needed to determine the various results including the values for b and c. In each case the physical domain, 0 < z < z ^ , was mapped onto a computation domain, |7]| < 1, by 2 8 £ = 1.8(1 + ^/(1.18-77).
(26)
Adjoint and linear systems for the primary instability modes represent eigenvalue problems with identical eigenvalues. Both were solved by using (26) and about 181 Gauss-Lobatto points28. The minimum Reynolds number for neutral stability of stationary (w = 0) modes and the associated wave numbers were found to be 28 Rc= 275.36, a = 0.38315, 0 = 0.078232.
(27 a,b,c)
These results are in agreement with those obtained by use of the finite difference method22. The same collocation simulation and mapping (27) with 181 Gauss-Lobatto points in the
158
computational domain were used to determine the solutions in the order e2 for the primary instability mode systems. The constants b and c are determined by evaluating their expressions 28 using Simpson's rule on 180 intervals to approximate the integrations28. The constants are b = (2.8993 x 10"5 - 2.5127 x lO-5*)/^,
(28 a)
c = 0.09759 + 5.1792/.
(28 b)
Since the Landau constant, -2c r , is negative, it is known 42 that |A| increases superexponentially for /^ > 0 (R > Rc). For /^ > 0 the solution breaks down after a finite time 28 ' 42 and (29) \A\^>ooasts^>:±-ln\l+_ 2br \ crK, where Ao is some initial value of \A\. It is anticipated that in this case there is a fast transition to turbulence28'42. For ft, < 0(ft < Rc), subcritial instability occurs since / (30) \A\ -> oo as ts -» lnl ^ — I A2-4? 2b, for AQ > Ac, where Ac = (-br/cr) . It is suggested46 that in this case there should be a lower critical value /^ of ft below which the bifurcated primary instability solution does not exist so that values of \A\ begin to increase with R. It is expected that the base flow is globally asymptotically stable with respect to all primary instability modes for R< Rc. So far we have assumed a primary instability mode consisting of a monochromatic wave with constant amplitude for all rand 0. More realistic is the consideration of localized primary instability modes whose properties can be studied by using the wave packet modeling approach29-32. Riahi30 and Vonderwell and Riahi31 used the method of multiple scales29'47 to examine the linear primary rotating disk flow instability, in order to make a better evaluation of the details of evolution of wave packet disturbances and their behavior which develops in both space and time. They assumed/? > Rc but R does not deviate much from Rc and defined slowly varying quantities 6S = £0, ts = et, E2R2 = R-Rc,
(31 a,b,c)
where £ is a small perturbation quantity and ft, is a positive constant. It is assumed that the primary instability modes vary on two different length and time scales d
d +£
M-*TO M-/
d
d
d
*->*
+fi
d
,„
*
(32a b)
^
'
The primary instability mode variables are expanded in powers of e and are used in the linear instability system. To lowest order in e it is assumed that any primary mode variable fx is of the form f^A{es,ts)f,{z)E,
(33)
159 and the goal has been to determine the form of the amplitude function A. To order e2 the existence condition yield dA
_ dA
+ G
^
_
= 0
^
„.. (34)
-
where Rf-Gg is the group velocity in the azimuthal direction. This equation implies A = A{ds - Gets) and suggests that instead of (31a,b) one should use the slowly varying quantities T = £2f
£ = e(e-Get),
(35a,b)
and proceed as before. The solution varies on two different time and length scales d
d
d
+e
(36a)
Te^to %> d
d
2
d
^
d
n^****-***
(36b)
To orders e and e2 one obtains the same linear eigenvalue problem as before, and the solutions are found. To order e3 the existence condition yield the equation for A dA
d2A
JT*W
,-_
'
( }
where ax and b are constant30. If the initial disturbance is centered at 0 = 0, r = /^, the solution for small T will be concentrated near £ = 0. Using the method for separation of variables, the solution for (37) is found to be A = A0 exp(fcT)T_>* exp(-^ 2 /4a 1 r),
(38)
where \ is a constant and the condition that A —> 0 as % —> °° is used to derive (38). This condition as well as the numerical computation of a{30 indicated that the real part of Oj is indeed positive. It is found that disturbance spreads out as T increases and at the center it grows exponentially, after an initial decrease produced by r~^ factor, rate of growth being proportional to (R - Rc). Riahi and Vonderwell32 extended the above analysis to nonlinear and more general propagating motion of the primary wave packet modes. It is found appropriate to define the slowly varying quantities (35) and r) = e{r-Grt\
(39)
where £ is defined in (31 c) and G = (GrGe) is the group velocity vector in the non-axial direction. Apply the nonlinear theory29, we pose expansion of the form (23) where now A = A(£,77,T). Substituting such expressions into the primary instability mode system and collecting coefficients of powers e leads to systems to orders eJe2E1£2E2ye2E° and e3E which are needed to determine an equation for A. The first four systems together with the linear adjoint system are solved by using a Chebyshev collocation method28. The solutions to these systems are then used in the solvability condition for the system to order e3E to
160 determine the following partial differential equation for A: dA
a
d2A
d2A
„
d2A
'W+2a2^+aiW)
= bA + c\A?A,
(40)
where the coefficients a,. (i = 1,2,3), b and c are constants. Define
£,=(S + M)/2, H=(« + J,l?)/2,
(41a)
where ^ and Aj are the roots of the quadratic equation a3A2+2^ +^ = 0
(41b)
for X. Using (41), (40) is converted to the form d2A
dA a-
= M + c|A|2A, a^a}-al/a3.
(42 a,b)
Linear solution of (42) is of the form A = (A0/r)exp(bT)exp[-ri£J(aT)l
(43)
where AQ is a constant. The maximum of the real amplitude of the wave packet derived from (43) decreases initially and then amplifies. In the time that elapses before the wave packet begins to grow it will propagate in the radial direction a distance leading to a value of Rc found to be closer to the experimental value48 than that due to the linear stability theory based on the monochromatic wave disturbances28. Using (43), the conditions for (42) can be written as A = \r~l Qxp[-T]^J(ar)] as T->0
(44 a)
|A|->0 as |77,£|->~.
(44 b)
System (42) and (44) is then well posed. For the purpose of the present discussion, we consider the case where d/drj = 0. Then (40) or (42) becomes ^ - a
l
^ = bA + c\A?A,
(45)
where the coefficients take the values a{ = R^\ b = (2.899 - 2.513/)/^(l0- 5 ), c = 0.098 + 5.179/,
(46a,b,c)
and Rc = 275.36 28 . The equation (45) is of the general mathematical form studied by Hocking and Stewartson 49 for various values of the coefficients. They found that the solution of the equation of the type (45) becomes more and more concentrated at the center (^ = 0) of the wave packet, becoming infinite there in a finite T , provided R^ is not too small. Otherwise, no bursting type of behavior is exhibited by A. Studies based on a less realistic case of a monochromatic wave disturbance28 do not exhibit this later result. The
161 results of the bursting for order one values of /^ are suggestive of possible fast transition to turbulence due to primary instability modes. It should be noted that a number of features used in the present model are determined before experimentally48 such as the results that the stationary spiral vortices are the primary instability mechanism which propagate and grow as wave packets, and the wave patterns from each point source spreads out circumferentially downstream of the source. 4. Spectral Theory In this section we review a spectral theory developed first by Plaschko and Hussain 35 for two dimensional shear layers and more recently extended 33 ' 34 to threedimensional shear layers such as slowly divergent free shear layers and boundary layers. Although there have been a number of theoretical modelings for secondary instabilities in shear layers 50 ' 51 , the spectral theory 33-35 , to be reviewed here, is apparently the only credible theory which provides some significant insights into the nature of the primary instability modes in non-parallel shear layers, and the results are consistent with experimental observations5'36. Consider a slowly diverging base flow field (w0, v0, w0) = u0 (JC, v, z). Here (JC, y, z) are the space variables along the streamwise, transverse and spanwise directions, respectively. Designate the flow divergence parameter by e and define xs = ex and zs - ez as the slowly varying streamwise and spanwise coordinates, under the assumption that nonparallel effects are equally important along the x and z directions. Based on the quasi-parallel assumption22'42, u0 = (u.evQ, w0) where («o,v0,w0) *S a n OI"der one quantity in each component. Next, superimpose three-dimensional primary instability modes on the base flow. Subtracting the non-dimensional forms of the governing Navier-Stokes and continuity equations for the base flow from the corresponding ones for the total flow (base flow + primary instability modes) lead to the following stability system for the primary modes f — + W 0 - V - - V 2 ] W + W - V W 0 + V P = -M-VW,
(47 a)
V w = 0,
(47b)
w = 0 at y = yx,y2,
(47 c)
where u = (w,v,w) and P are the velocity vector and pressure for the primary instability modes, R is the appropriately defined Reynolds number of the flow, and yl and v2 are the values of v at the lower and upper boundaries of the flow, respectively. Now expand the primary mode dependent variables in powers of e with the assumption that nonlinear and nonparallel effects are equally important. L p ) = / M p ^ l + e 2 f« 2 ,P 2 j + ...
(48)
Using (48) and (47) and collecting terms of equal powers in £, yield systems to various orders in e. The order e system 3334 admits solution of the form
162
UiiXs.y.ZstW.P)]
r
Pi(xs>y>zsw>P) \ +pz - wt]}dwdp,
(49)
where w is the frequency, P is the span wise wave number and g(xs,zs>w,P) is a function to be determined later. The fast variable in (49) is defined to be g/e. The solution of the form (49) is given in general form to cover the following three special cases: (i) Twodimensional primary instability modes acting on two-dimensional base flow 35 . Here zs = 0, P = 0 and (49) reduces to a single integral over frequencies. An example for this case is two-dimensional primary instability of Blasius flow, (ii) Three-dimensional primary instability modes acting on three-dimensional z -independent base flow. Here zs = 0 in (49). An example for this case is three-dimensional boundary layer primary instabilities of flow over an infinite span wing, (MI) Three-dimensional primary instability modes acting on three-dimensional z-dependent base flow. Here P = 0 and (49) reduces to a single integral over frequencies. An example for this case is three-dimensional boundary layer primary instabilities of flow over a finite span wing. Using (49) in the e order system yield the system for \ul,Pl
. This system represents the three-dimensional
extension of the well known Orr-Sommerfeld equation governing a linear primary instability mode in a parallel flow if one defines —— and /?+— as the streamwise and dxs
dzs
spanwise wave numbers, respectively 33-35 . In order to determine the weakly varying amplitude function Al(xs,zs,w9P), given in (49), one need to consider the e2 order system and determines its solution 33>35. Also, one need to consider the adjoint eigenvalue problem of the linear system which is a system for the adjoint variables \ux, Px adjoint linear solution is of the form (49), provided
AjWpAjPi
33 35
" . The
is replaced by
M 1 (X 5 ,V,Z 5 ,W,J8)P 1 (^ 5 ,V,Z S ,W,^)J. Applying the solvability condition to the system for the y -dependence functions of the e2 order system33-35, yield the following integro partial differential equation for Al: k(xSJzSyWiP)—L + l(xs1zSyWip)-^ + m(xSizs1w,P)Al dxs dzs
=
[exp{-ig/£)]'\ZLH{xs>Zs^P>wi>Pi)A{xs^s^ ■txp[i[g(xs,zs,wiypl)
+ g(xs,zs,w-wl,p-p])]/e}dvv]dpi,
(50)
163 where the coefficients kj,m, and H depend on the solutions of the adjoint and e order systems, but they are independent of A, 33 " 35 . Equation (50) is a kind of generalized Landau equation where the nonlinear term in the right hand side represents nonlocal interactions between different wave modes over the frequency and spanwise wave numbers domain of primary instability waves. Equation (50) is too-complicated and no analytic solution appears to be possible, except for large x and z. However, this equation can be solved by numerical iterative process starting with the solution of the linear part of (50). Let us now consider the phase function g given in (50) and define § = a(*,z s ,w,j8), § = 7(*5,z,w\jB).
(51a,b)
as the local wave numbers along x and z axes, respectively. It then follows from (51) that, to the leading orders, a is a function of x only and y is a function of z only. Therefore -g{xs,zs,w,p)
= jXoa(x\zs,w,p)dx'
+
jj(xs,z\w,p)dz''
(52)
Now the oscillating term, which appear in the integral in the right hand side of (50), is of the form e x p | i j J a r ( ^ z 5 , w P A ) + a r (jc / ,z 5 ,w-w I ,i8-A)] dx^i^r^z^w^P^y^z^w-w^P-P^dz^
(53)
where ar and yr are the real parts of a and y, respectively. We use the convention from the parallel flow stability theory 2 ' 42 that ar and yr are positive for positive frequencies. Thus, for locations far downstream (JC,Z » 1), the integral in (53) takes large oscillations and we can use the method of stationary phase 52 to determine the dominant contribution of the nonlinear term in (50). The dominant contribution comes from frequencies w i = w'^ and wave numbers at which the phase in (53) is stationary 33-35 . Using this result in (50) yield k(x9z9w,P)^
+ l(x,z,w,P)^
+ m(x9z,w,p)Ai =
r(x,z,w,j3)A 2 (^z,f4),
(54)
where T is independent of y^ 33-35 . This equation can be solved easily by using the method of characteristics34-35• It is seen from (54) that, for given w and /?, the dominant nonlinear contribution comes from the subharmonic mode at the frequency w/2 and at the spanwise wave number P/2. This result is consistent with experimental data 5 ' 36 and agrees with the theory due to Plasckho and Hussain35. Previous nonlinear theories 35 ' 53 failed to predict this role of the subharmonic. Let us now pay attention to the fact that the experimental results36 also indicate dominance of other subharmonic modes at frequencies
164 w/n(n>2) for slightly larger values of R. The theory developed by Plaschko and Hussain35 is unable to predict the dominance of any subharmonic mode other than just one at w/2 and, indeed, solvability conditions at orders en(n > 2) of that theory may contradict each other. In particular, the solvability condition in the order e3 imposes further condition on Ay which may not be obeyed by (50). However, the reformulated and generalized theory due to Riahi 33-34 is capable of predicting the dominance of the subharmonics at frequencies w/n(n>2) and at spanwise wave numbers p/n(n>2) and there are no contradictions between the solvability conditions at different orders in e. As explained in Riahi 33 ~34, the main reason for the deficiency of the original theory35 has been the fact that only a particular solution for the e2 order system was considered and a large subset of frequencies contributing to the most general solution for the e2 order system was discarded in that theory Riahi33"34 considered the most general solution to the e2 order system with inclusion of all the frequencies and the spanwise wave numbers which contribute to such solution. It turns out that, due to contribution of all the frequencies and spanwise wave numbers, the solvability condition at each order en(n>2) leads to an equation for an amplitude function An_x which is related to Am(m
+ rr dz*}-
(55)
Using the method of stationary phase to determine the dominant contribution of the nonlinear term in the equation for ^4n_, and apply similar procedure to that used for ( 5 0 ) 3 3 " 3 4 , we find the phase is stationary at wl=w2=... = wn_l=w/n and Px = P2 = ••• = Pn-\ = P/n. Hence for given w and /?, the dominant nonlinear contribution in the equation for An_l comes from the subharmonic mode at the frequencies w/n and at the spanwise wave number P/n. 5.
Flows Over Swept Wings
Another important example of shear flow with practical applications is that of flow over swept wings. Such flows in aerodynamic applications are essentially within threedimensional boundary layers which are subjected to primary crossflow instability producing cross flow modes as well as primary viscous instability producing the TollmeinSchlichting (TS) modes. TS modes are the primary instability modes in two-dimensional boundary layers. For three-dimensional boundary layer flows over swept wings, crossflow modes (CF) dominate at the leading edge of the wing, while TS modes show larger growth rates in the mid-chord region. In between, there is a region where both CF
165 and TS modes are present. We shall first briefly review the asymptotic analysis of the linear primary instability modes of the three dimensional compressible boundary layer over a swept wing which has recently been done 37 and led to some understanding of the behavior and characteristics of such primary modes. The governing system for the linear primary instability modes of the present problem are given by several authors 38 ' 54-55 and will not be given here. Consider a coordinate system with x-axis along the normal chord, the z-axis along the span and the y-axis normal to the surface of the wing. The boundary layer on the wing is assumed slightly non-parallel, and the non-parallel effects do not enter the governing linear primary stability system to the leading terms. The non-dimensional form of the governing primary stability system contains mean flow variables as coefficients, four non-dimensional paramenters R, Pr, M and / , and the expressions for the viscosity ju and the conductivity K and for their first and second derivatives with respect to temperature which are given by Sutherland type laws 56 . Here R is the Reynolds number, Pr is the Prandtl number (defined based on the free stream quantities), M is the Mach number (defined based on the free stream quantities) and y is the ratio of specific heats. It is assumed that R is sufficiently large. Basically there are two sets of primary modes: inviscid modes which are useful to describe the upper branch disturbances and wall modes which are useful to describe the lower branch disturbances57. For the inviscid modes the appropriate small parameter £ is given by £ = R~y\
(56)
and any dependent variable F of the primary modes is of the form F = f{y)exV{(i/e3)[ja(x,e)dx
+ zP(e) - e'M^l
(57)
where T is time, a is the streamwise wave number, p is the spanwise wave number, w is the frequency and f(y) is the y-dependence part of F. We restrict our attention to neutral disturbances. The boundary layer for these modes is divided into an inviscid zone of thickness 0(e 3 ) and a thin viscous zone of thickness 0(e 4 ). In the inviscid zone, the governing equation are simplified. These equations have coefficients which depend on the mean flow quantitites. The equations are solved using spectral collocation method. The mean flow data used in the computation is based on the boundary layer flow on an infinite span 23° swept wing with suction, M = 0.82 and chord Reynolds number of 2 x 107. The boundary layer edge velocities are determined by integration of the surface Euler equations knowing the pressure coefficient. The mean flow solution is found using the compressible boundary layer finite difference code of Iyer 58 ' 37-38 . Since the long time scale modes are found to be preferred, the results are independent of w to the leading terms. Computational results for the eigenvalue problem of the inviscid zone led to the values for the wave number and the wave angle arctan (p/oc) at several nodes37. It turns out that it is possible to have more than one solution at a node. Each of such solution at a particular node may correspond to a particular w which can be confirmed only by some higher order analysis and calculation57. In the viscous zone, the equations are simplified and solved using matching condition with the inviscid zone solution at the upper boundary. Next, the wall modes are considered, where the appropriate small parameter £ is given by
166 £ = R~yi\
(58)
and any dependent variable F of these modes is of the form F = f{y)exp{(i/e4)[ja(x,e)dx
+ zP(e) - e4Tw^)]}-
(59)
These modes have a triple-deck23 structure. In the upper deck of thickness 0(e 4 J, the variables are properly scaled and both viscous and conduction effects are negligible. The governing equations are then simplified and solutions for the dependent variables are found 37 . In the main deck of thickness 0(£ 8 ), the variables are properly scaled and both viscous and conduction terms are negligible. The governing equations are then simplified, and solutions for the dependent variables which match those for the upper deck at the upper bound are found. An interesting result for the main deck is that alternative scalings for the density and temperature variables of the primary modes are possible which can lead, to leading order terms, to zero vertical velocity (vertical vorticity mode) or zero temperature (isothermal mode) or zero vertical velocity and temperature (vertical vorticity and isothermal mode). It turns out that the solution to all the components of the velocity vector do not match the no slip condition at the wall which requires consideration of a lower deck zone. In the lower deck of thickness 0(e 9 ), the variables are properly scaled and the governing equations are simplified and then solved. The requirement of no slip condition at the wall and the matching condition with the main deck solution at the upper boundary lead to the results for the wave number and the wave angle37. The complete incompressible and compressible Navier-Stokes, continuity, state and energy equations for the primary instability modes have also been solved numerically 38 ' 59-60 to determine the solutions for such modes in the three-dimensional boundary layers over an infinite span swept wing. Since such systems have coefficients which depend on the steady state base flow solution of the governing Navier-Stokes, continuity, state (for compressible case) and energy equations, such governing equations, subjected to the boundary condition for an adiabatic wing with weak suction 59 , were solved first for flow over a 23° swept infinite span wing 38 ' 60 . A finite difference code 58 , fourth order accurate in the v -direction (wall normal direction) and second order accurate in the x -direction, was used to determine the base flow solution for a flow with free-stream Mach number of 0.82 38 ' 60 . It should be noted that both baseflow and stability systems are derived in a systematic way by the following procedure 38 ' 60 . First, introduce the slowly varying variables xs = ex, zs = £z, ts = £?,
(60 a,b,c)
where the small parameter e characterizes the magnitude of the disturbance quantity and we assume 59 that it simultaneously characterizes the nonparallelism of the boundary layer. Next, pose as a solution to the basic governing system for total quantities (base flow + disturbance) the form F = F0{xs,y) + eFl(x,y9zJ9xSJzsts)
+ e2F2(xty9zXxS9zS9ts)9
(61)
167
where F is any dependent variable, F0 is the variable for the base flow and FY and F2 are the variables for the disturbance. For the total viscosity we assume the form
A£ = ,io v0(r") +... 0) + A ^7; +
(62) (62)
with a similar expansion for the total conductivity. Here T denotes temperature. Substituting these forms into the system for the total quantities, the base flow system is obtained to the zeroth order in e. To the first order in £ the linear stability system is obtained which is appropriately simplified by assuming the following form for F{: ^^x ==^n(^,y,z 5 ,r 5 )exp(/0i**), n ), Fi.(xs>y.Zs'h)*M
(63) (63)
where the phase function (/>n satisfies = a
x
z
x
W X [\j£
(64) (64)
Here, an,pn, and wn are the streamwise wave number, spanwise wave number and frequency of the disturbance, respectively. The linear stability system for dependent variables like Fln is an eigenvalue problem. Once an and pn are selected, a non-trivial solution exists only for certain values of wn. The linear stability system is solved by the collocation method 38 ' 60 for both incompressible and compressible cases. Both the incompressible and compressible systems result in a linear stability eigenvalue problem which has a dispersion relation of the form (65) (65)
,Pn,w„,R) = 0. D(an,3„,w„,R) 0.
The real part of the frequency, wnr of the disturbance remains constant downstream, and for the special case of spatial instabilities with zero spanwise growth rate on an infinite span swept wing, pnr is fixed as well. The chordwise wave numbers, anr, will change as the boundary layer grows, however, as will the growth rate given by -ani. Once the values of wnr and pnr are specified, a search for the remaining eigenvalues must be undertaken at each node of interest. A Newton-Raphson search is used to find the correct value of In addition to the individual primary instability modes determined from the linear stability system described in the previous paragraph, instability modes that satisfy the triad resonance conditions can also be determined using the wave interaction modeling 38 ' 42 ' 60 . Using (61)-(62) in the system for the total quantities and retain only term of order e 2 lead to a system whose solvability condition lead to a system for the amplitudes an(xs,zsts), provided solution to the e order linear system is assumed to be in the form of
,=i^,,,ssW*)K(*,,«) ,=i^„, w*)K(*,,«)
(66)
in place of (63). To form a triad, it is required that (a^P3r,w3 (cc,r,P wlr) }r,wrir)) = (alr,plr lrWlr
+ (a2r,p2r,w2r)
+ e(<7 e(a a,<7 a,a^a p,a^, w),
(67) (67)
168 where the last term in the right hand side of (67) represents a small deviation from perfect resonance. The solvability condition for the e2 order system is a system of three coupled first order partial differential equations in xs,zs and ts for an(n = 1,2,3) with quadratic nonlinearities in an3S>60. In actual computation 38 ' 60 , the special case of an infinite span wing and spatial stability theory are considered, so that the solvability system reduces to a system of first order ordinary differential equations in xs for an(n = 1,2,3). To make the interaction coefficients of the nonlinear terms in the solvability system unique, the eigenfunctions were normalized by setting the eigenfunction magnitude of all three members (n = 1,2,3) equal to one, and then preserves the eigenfunctions magnitudes downstream. With initial values of an specified, the system for an is integrated using a standard fourth-order variable interval Runge-Kutta method 38 ' 60 . The results of the computations indicate that if the interaction coefficient for a mode is small, then the mode's amplitude can be boosted only if it is much smaller than the amplitude of the other two members of triad. The combinations of two TS modes and one CF mode always show a very strong crossflow interaction coefficient and very small TS interaction coefficients. Thus a small amplitude CF mode may be boosted after the TS modes have reached some finite amplitude. A comparison between the results for the incompressible and compressible stability system cases indicate that the popular and simpler incompressible modeling may not yield meaningful results and, thus, the compressibility effects must be taken into account in the stability system. 6.
Some Concluding Remarks
In this chapter we reviewed the structure and characteristics of the primary instability modes and their important effects on the evolution of the fluid flows in several basic shear flow systems. In particular, we emphasized the roles of the primary viscous and spatial instability modes in confined and unbounded mixing layers, the subcritical and bursting behavior of primary instability modes, in the form of either monochromatic waves of wave packets, in the rotating disk flows with subsequent possible fast transition to turbulence due to such modes, the dominance of primary instability subharmonic modes in non-parallel shear layers, triple-deck structure and long time scale behavior of primary instability modes in three-dimensional boundary layer flows over airfoils and large growth rates of primary instability crossflow modes due to triad interactions with primary TS modes in flows over swept wings. In the last several decades there have been significant progress in understanding the shear flow instabilities and their roles in important practical applications for flow control and for turbulence promotion. Most of the large amount of work in the shear flow instability area that have been reported by many authors in the literature so far, have been on the secondary instabilities61 in the boundary layers and free shear flows, and primary instabilities have been referred to as long as their relation to the secondary instability modeling and simulation were concerned. Despite this, the importance of primary instability modes by themselves cannot be under estimated. The present review chapter, which is devoted entirely to primary instabilities, demonstrates the significance and important roles of primary instabilities on the overall shear flow behavior and on the past, present and future evolution of shear flows. Primary instability modes are essential parts
169 of any shear flow instability mechanism and, indeed, secondary and higher instability modes cannot exist without the presence of the primary ones. Finally it should be noted that there are other important basic shear flow systems, such as those of wake flows and channel flows referred to only briefly in the section 1, or those of jets and pipe flows which were not even referred to in this chapter. However, such flow systems can be studied using the same types of mathematical and computational methods introduced in this chapter that were found to be so effective to determine the results for the primary instabilities. Also certain types of primary instabilities such as primary algebraic instabilities62 can be operative not only in channel flows but in any wall shear flows. In regard to wake flows, a separate article63 is devoted to the computational simulations of primary instabilities in planewake flows using techniques which can also be employed in other free shear flow systems.
170 References 1. K.H. Lie and D.N. Riahi, Int. J. Eng. Sci. 26 (1988) 163. 2. R. Betchov and W.O. Criminale, Stability of Parallel Flow (Academic Press, NY, 1966). 3. A. Michalke, 7. Fluid Mech. 23 (1965) 521. 4. P. Freymuth, J. Fluid Mech. 25 (1966) 683. 5. R. W. Miksad, J. Fluid Mech. 56 (1972) 695. 6. G. Brown and A. Roshko, /. Fluid Mech. 83 (1974) 641. 7. D. Oster and I. Wyganski, J. Fluid Mech. 123 (1983) 91. 8. S.A. Orszag, J. Fluid Mech. 50 (1971) 689. 9. D.N. Riahi, R.D. Moser and F. Waleffe, Proceed - Summer Research at Center for Turbulence Research (Stanford, CA) (1990) 205. 10. A.S. Burger, Laminar Wakes (American Elsevier Publishing Co., NY 1971). 11. T.H. Kuan and D.N. Riahi, TAM Report #703 (UIUC 1992) 1. 12. T.H. Kuan and D.N. Riahi, Develop, in TAM 17 (1994) 557. 13. T.H. Kuan and D.N. Riahi, Proceed. IMAC 14th World Cong, on Comput. and Appl. Math. (Atlanta, GA) I (1994) 275. 14. H. Sato and K. Kuriki, /. Fluid Mech. 11 (1961) 321. 15. G.E. Mattingly and W.O. Criminale, J. Fluid Mech. 51 (1972) 233. 16. E. Meiburg and J.C. Lasheras, /. Fluid Mech. 190 (1988) 1. 17. D.T. Papageorgiou and F.T. Smith, Proc. R. Soc. Lond. A 419 (1988) 1. 18. D.T. Papageorgiou and F.T. Smith, /. Fluid Mech. 208 (1989) 67. 19. S.P. Wilkinson and M.R. Malik, AIAA paper no. 83-1760 (1983). 20. M.R. Malik, S.P. Wilkinson and S.A. Orszag, AIAA J. 19 (1981) 1131. 21. L. Mack, AIAA paper no. 85-0490 (1985). 22. M.R. Malik, J. Fluid Mech. 164 (1986) 275. 23. P. Hall, Proc. R. Soc. Lond. A. 406 (1986) 93. 24. M.R. Malik, AIAA paper no. 86-1129 (1986). 25. W.S. Saric and L.G. Yeates, AIAA paper no. 85-0493 (1985). 26. N. Itoh, In Laminar-Turbulent Transition, ed. V.R. Kozlov (1984) 463. 27. S. MacKerrell, Proc. R. Soc. Lond. A. 413 (1987) 497. 28. M.P. Vonderwell and D.N. Riahi, Int. J. Eng. Sci. 31 (1993) 549. 29. K. Stewartson and J.T. Stuart, J. Fluid Mech. 48 (1971) 525. 30. D.N. Riahi, Instability and Transition, Vol. II, eds. M.Y. Hussaini and R.G. Voight (1990) 160. 31. M.P. Vonderwell and D.N. Riahi, Bull. APS. 35 (1990) 2322. 32. D.N. Riahi and M.P. Vonderwell, Proceed. Pan Am. Cong. Appl. Mech. (Brazil) (1993) 331. 33. D.N. Riahi, Proceed. IMAC 14th World Cong. Comput. and Appl. Math. (Atlanta, GA) I (1994) 421. 34. D.N. Riahi, Submitted for publication (1995). 35. P. Plaschko and A.K.M.F. Hussain, Phys., Fluids A. 27 (1984) 1603. 36. Z.D. Husain and A.K.M.F. Hussain, AIAA J. 21 (1982) 1512. 37. D.N. Riahi and M.P. Vonderwell, Proceed, of NASA Langley Workshop on Transition, Turbulence and Combustion (Hampton, VA) I (1994) 235.
171 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63.
M.P. Vonderwell and D.N. Riahi, submitted for publication (1995). D.N. Riahi, submitted for publication (1995). P.A. Monkewitz and P. Huerre, Phys. Fluid 25 (1982) 1137. L. Fox and LB. Parker, Chebyshev Polynomials in Numerical Analysis (Oxford Univ. Press, Oxford, U.K., 1968), 205. P.G. Drazin and W.H. Reid, Hydrodynamic Stability (Cambridge Univ. Press, Cambridge, U.K., 1980) 527. H. Schlichting, Boundary Layer Theory (McGraw-Hill Book Co., 1979) p. 817. J.T. Stuart, J. Fluid Mech. 4 (1958) 1. J.T. Stuart, J. Fluid Mech. 9 (1960) 353. L.D. Landau and E.M. Lifschitz, Fluid Mechanics, 2nd edition (Pergamon Press, NY, 1987), 539. A. Davey, L.M. Hocking and K. Stewartson, J. Fluid Mech. 63 (1974) 529. S.P. Wilkinson and M.R. Malik, AIAA J. 23 (1985) 588. L.M. Hocking and K. Stewartson, Proc. R. Soc. Lond. A. 326 (1972) 289. T. Herbert, AIAA paper no. 85-0489 (1984). T.M. Fischer and U. Dallman, AIAA paper no. 84-0009 (1984). G.F. Carrier, M. Krook and C.E. Pearson, Functions of a Complex Variable (McGraw-Hill Book Co., NY, 1966), 438. D.J. Benney and S.A. Maslowe, Stud. Appl. Math. 54 (1975) 181. N.M. El-Hady, AIAA paper no. 80-1374 (1980). L.M. Mack, AGARD report no. 709 (1984). F.M. White, Viscous Fluid Flow, 2nd Edition (McGraw-Hill Book Co., NY, 1991), 614. D.N. Riahi, In preparation (1995). V. Iyer, NASA contractor report 4269 (1990). N.M. El-Hady, Phys. Fluids A 1 (1989) 549. M.P. Vonderwell, Ph.D. Dissertation, UIUC, Illinois (1994). T. Herbert, Ann. Rev. Fluid Mech. 20 (1988) 487. M.T. Landahl, J. Fluid Mech. 98 (1980) 243. T.H. Kuan and D.N. Riahi, in preparation (1995).
173
GORTLER VORTICES WITH SYSTEM ROTATION Abdelfattah Zebib Rutgers University Department of Mechanical & Aerospace Engineering Piscataway, NJ 08855-0909, USA and Alessandro Bottaro Ecole Polytechnique Federate de Lausanne IMHEF-DME CH-1015 Lausanne, Switzerland and Barbro G.B. Klingmann Volvo Aerospace Corporation Space Propulsion Division S-461 81 Trollhattan, Sweden
ABSTRACT Spatial development of longitudinal vortices is determined by a linear, streamwise marching, stability approach as well as by a nonlinear, finite-volumes computational method. The basic flow is a Blasius boundary layer on a rotating concave surface, with the rotation vector parallel to the span. It is shown that linear marching results agree well with nonlinear spatially developing solutions up to fairly large disturbance amplitudes. It is also shown that while rotation influences the growth of the "mushroom" structures (rotation in the sense of the basic flow enhances the development of the vortices while counter-rotation dampens the vortices), the nonlinear saturation energy for a given spanwise wavelength is not dependent on rotation. Wavelength selection based on maximum linear amplification is investigated within a range of rotation numbers including zero, and is found to be rather weak, particularly in the case of counter-rotation. Under random disturbance inflow conditions, perturbation energies decrease initially in a linear filtering phase (which does not depend on rotation, but is a function of the inlet noise distribution) until coherent patches of vorticity near the wall emerge and can be amplified by the instability mechanism. The average wavenumber found is close to that predicted by the linear criterion.
174 1.
Introduction
Spatial development of the boundary layer flow over a concave surface subject to system rotation, with axis of rotation perpendicular to the plane of the base motion, is the subject of this paper. In addition to applications in rotating machinery, interest in vortices developing along curved boundary layers stems from their similarities with coherent structures in the near-wall region of turbulent shear flows1. Experiments 2 have provided a basis for the computational studies of, among others 3 ' 4 . While computations have faithfully reproduced laboratory results when experimentally established initial and boundary conditions are used, only theory 3 ' 5 has provided some criteria for the wavelength selection. It has been argued that a wavelength in the neighborhood of the curve of maximum linear amplification has a good chance of being selectively amplified. This has been confirmed in experiments where the upstream flow perturbations are randomly distributed along the span, but it is at variance with other experimental evidence that indicates that different wavelengths are selected in different experiments 2 ' 6 " 8 . In fact, the available literature for the nonrotating Gortler flow provides average values of the selected span wise wavenumbers that differ by up to a factor of 10. This fact indicates two things: the downstream vortex development is strongly influenced by the inlet flow and disturbance field (typical of a convective instability), and the wavelength selection mechanism is weak. In the present paper we report results from numerically solving the steady, nonlinear equations of motion to assess the influence of rotation on the spatial development of the flow along a concave wall, and the wavelength selection mechanism. These equations are parabolic, and are thus solved by straightforward marching 3 . In addition, the linearized version of these equations is solved by a finite difference-spectral marching procedure. Linear and nonlinear results, including those from local nonparallel linear stability 9 , for different rotation numbers complement and confirm each other convincingly; both codes are then used to address issues of wavelength selection.
2.
Mathematical Formulation
We consider the rotating boundary layer flow over the concave wall with constant radius of curvature R subjected to system rotation Q. with the axis of rotation coinciding with the axis of curvature (see Figure 1). Following9, we take / to be a typical length along the wall and define the curvature parameter e, the Reynolds number Re and the Gortler number G by: e = /fl-i,
Re = ^ - , v
G2 = £Re"2,
(1)
175 where Uoo is a free stream velocity and v is the kinematic viscosity. The theory is developed for 8 -» 0, Re -» «> such that G = 0(1). In a cylindrical coordinate system (r, 9, 0 with corresponding velocity components (vr, vG, v^) we introduce the dimensionless boundary layer coordinates x, y, and z: y=-^^Rel/2,
z = ^Rel/2.
(2)
and the dimensionless velocities u, v, w; and pressure p:
•■-£■
Figure 1.
v = - u ^Re'/2 >
w = y^Re I / 2 ,
P=
pU 2 >
Re.
(3)
Sketch of the problem.
where p is the density and p' is the dimensional pressure. When these variables are introduced into the Navier-Stokes equations, written in the rotating frame of reference of Figure 1, we find the leading system: du 5x
3v 5y
3w _ ~dz~ '
(4)
176
d
r
r [
U
^ ^
+ V
^ ^
d2u 3 2 u
d,
d
+ W
3, 3i]V
-,~ r> + G2u2 =
d d d, r [U^-+V^-+W^-JW L J dx
dy
dz
3D 3 2 V 3 2 V - _ ^ry - ^ + ^2 + ^2-2R0G2u'
3p 3 2 w 3 2 w = - V : + -r-=r + -=-7T. dz 3 y 2 9Z2
/CN
(6)
,_. >
v(7)
The boundary conditions associated with the parabolic system (4) - (7) are: u=v=w=0
aty = 0,
(8a)
uy=v =w=0
asy-»oo.
(8b)
where the infinite field was truncated at y e = 50. It also was checked that the adoption of conditions at y e such as 3 : u = 1, Vy = 0, w = 0, produced the same results as the boundary conditions (8b). Most computations are performed for one-half wavelength (7t/oc, where a is the wavenumber at x = 1), and along z the symmetry conditions u z = v z = w = 0,
for z = 0,7t/oc,
(8c)
are adopted. The numerical solutions of the nonlinear parabolic system is accomplished by a finite volume procedure 10 . Briefly, the y-z computational domain is divided into rectangular cells with the grid points located at the geometric centers of these small cells. Additional boundary points are included to incorporate the boundary conditions. The discretized equations are obtained by integrating the conservation equations over control volumes having these cells as a base and extending a distance Ax in the streamwise direction. Local linear y and z dependence in any of the primitive variables is assumed, resulting in a second order accurate scheme. Staggered location for the y-z velocity components is adopted to avoid unrealistic pressure fields and associated numerical instabilities. Streamwise marching is accomplished through a fully implicit first order forward Euler scheme. For the parameter values of this work, we have determined that a 71x25 y-z grid (stretched in y and uniform in z) and Ax = 0.04 are adequate for computations covering one-half wavelength. Numerous computations were also performed on a finer grid (121x50 and Ax = 0.01); each one of these latter calculations required of the order of 15 CPU hours on the Cray-2 supercomputer at EPFL. We also consider the linear spatial development of a perturbed Blasius profile:
177 u = (U(x,y) + ui(x,y,z), V(x,y) + vi(x,y,z), wi(x,y,z))
(9)
where (U, V) is the Blasius solution, which remain unchanged in the limit e -* 0. The linear stability equations are obtained by linearizing equations (4) - (7), and then reducing the system to a two-equation set in which the pressure and the spanwise disturbance velocity component are eliminated. The resulting equations for the perturbation velocities ui and vi are equivalent to those given by Hall 11 , with an additional rotation term. They are solved by a downstream marching procedure employing a second order accurate finite-difference scheme in x and Chebyshev collocation in y12> 13 > 14 . A local nonparallel analysis is also performed 9 ' 14 . It provides local growth rates, a, and eigenfunctions which can be used as initial conditions for both linear and nonlinear spatial developments.
3.
Linear and Nonlinear Development of the Vortices
It has been shown experimentally2 that the initial spatial development of Gortler vortices is steady. This development depends on a large number of variables, such as the shape and the streamwise position / of the initial disturbances. Other parameters at play fix the values of Re, G and Ro. Whereas the locally scaled wavenumber a increases as x 1 / 2 and the local G as x 3/4 , the dimensionless wavelength parameter A, defined as
A = ^ ' V ^ = G(-) 1 - 5 v
V"
(10)
a
remains constant at all streamwise locations (unless merging and/or splitting of vortices takes place and the physical wavelength A,' of the vortices is modified). Here we consider the influence rotation would have on previously reported experiments 2 and computations 3 . The detailed measurements 2 were made at Uoo = 500 cm/s, R = 320 cm, v = 0.15 cm 2 /s, for a vortex pair with a wavelength A/ of 1.8 cm, giving A = 450. The simulations are started at / = 40 cm, so that at x = 1, G = 6.756, a = 0.3824, and an initial disturbance (ui, vi, wi) composed of the local linear eigenfunctions for Ro = 0 with u i m a x = 0.042 is used 3 to correspond to the experimental data 2 . Figure 2 illustrates the "mushroom" growth corresponding to Ro between -0.2 and 0.2. Here contour plots of constant u are shown at 5 axial locations. The results obtained at Ro = 0 are in excellent agreement with previously published results 3 . The stabilizing (destabilizing) influence of negative (positive) rotation which is predicted by the local linear analysis in 9 is verified by the nonlinear calculations. The penetration of low momentum fluid into the free stream, indicated by the 0.9 u-level, is seen to increase (decrease) according to Ro > 0 (< 0). Further evidence of this influence can be seen from the "peak" and "valley" wall shear stress variation with x in Figure 3. At the mushroom
178
Figure 2.
Development of vortices in x corresponding to different values of Ro and for a disturbance wavenumber a = 0.3824. At x = 1, G = 6.756 and a perturbation proportional to the local linear eigenfunction at Ro = 0 normalized according to ui = 0.042 is assumed. Deeper penetration of low momentum fluid into the free stream at larger values of Ro is observed. The y-z grid used is 121x50, with Ax = 0.01; the range of y in the plot is [0,20] and the plots are scaled correctly. Spacing between neighboring isolines is 0.1.
179 stem (peak) positive rotation first decreases the wall gradient, with a subsequent increase which is more rapid the larger the rotation number. In the valley, the spatial development is also faster at large positive Ro, and the maximum value attained is higher. Far downstream, the shear stresses at the peak and valley locations reach almost constant values which depend weakly on Ro.
Figure 3.
Variation of wall shear stress with x for the same conditions of Figure 1. The destabilizing influence of rotation can be inferred from the decrease (increase) in shear stress at peak (valley) locations.
The growth of the perturbation energy with x as function of Ro (all other parameters fixed as stated above) is compared in Figure 4 to linear marching results. The nonlinear perturbation energy is defined as: a ^e E = 5 f fui 2 dydz,
(ID
180
Figure 4.
Growth of the perturbation kinetic energy E as defined in Eqns. (11) and (12) corresponding to nonlinear and linear marching predictions, respectively.
and the curves of linear growth shown in Figure 4 are computed according to X
E = E 0 e 2 fadx,
(12)
where Eo is the kinetic energy of the perturbation at x = 1. It can be seen that the linear marching results shown in Figure 4 faithfully reproduce the nonlinearly computed growth rates until quite large values of E (local linear results9 slightly overestimate the growth of the Gortler vortices14). Far downstream, nonlinear saturation leads to a departure from the linearly predicted energy growth. An interesting observation to be made from Figure 4 is that, despite the large influence of rotation on the initial development, the vortices have the same perturbation energy regardless of Ro, once the nonlinear saturation stage is reached.
181 4.
Wavelength Selection
According to the local linear theory 9 , the most amplified wavelength A lies in the neighborhood of 260 (160) for Ro = -0.3 (0.3). In the following parametric study, the linear marching code was used to obtain solutions for a large number of A's and Ro's.
Figure 5.
Amplification rates a for different A at different constant values of G.
Each run was initiated at G = 1 with the solution of the local stability problem. It was noted that varying the starting point of the downstream marching did not significantly affect the results. Figure 5 shows how the amplification rates vary with A and Ro, at fixed values of G. In all cases, the dependence of a on A (wavelength selection) is weak at low G where the linear amplification starts, giving similar amplification rates within a wide range of A's between 100 and 2000. As G increases, the region of intensively
182 amplified A's becomes slightly narrower, particularly for positive Ro. Some nonlinear calculations did indeed show that vortices with widely varying A had practically equal amplification rates, and more so for Ro < 0 than for Ro > 0. With increasing Ro the most amplified wavelengths shift towards lower values. This is seen clearly in Figure 6, where the most unstable A (i.e. the A where a is locally maximum at a given downstream position, G), is plotted as a function of Ro. When Ro < 0 the range of the most unstable A's varies quite broadly with G, whereas the variation is contained for Ro > 0; for example, when Ro = 0.5, the most amplified A is in a narrow range centered on A = 160, whereas for Ro = 0, it varies between A = 185 and 240 as G increases from 2 to 15. This fact could be interpreted as an indication that the wavelength selection mechanism is weaker at negative Ro.
Figure 6.
4.1
Variation of the most amplified wavelength with Ro, at different values of G. The crosses correspond to results of section 4.2, for natural wavelength selection, when the vortices begin to appear.
Nonlinear Development
It is by now clear that a system subject to controlled perturbations can amplify vortices of almost any desired wavelength within a large range. On the other side, a randomly distributed field of incoming disturbances can be expected to develop into vortices with the linearly most amplified wavelength. This can be verified by considering a long (in z) test section and applying at the entrance of the computational domain a white noise perturbation field. This is the strategy adopted in the following nonlinear calculations. These runs start from G = 6.756 with a computational box of dimensionless spanwise length equal to 89.7. A 72x152 mesh was found adequate for these simulations (a finer
183 mesh was also employed to verify the accuracy of the results). We performed two calculations both starting from the same global disturbance amplitude, but with different random noise distribution with Ro = 0.3. The perturbation energies in Figure 7 differ between the two cases, except clearly for x = 1, and also the saturation levels reached are slightly different. The latter is an effect of the different filtering selection actuated by the equations in the linear regime when E decreases, and of the different average wavelengths achieved in the two cases in the nonlinear stages. Note, however, that the streamwise extent of the initial energy-decreasing phase is the same (from x = 1 to about x = 1.5) for the two different random noise distributions. The streamwise velocity fields are shown in Figure 8 at five different x values; the spatial evolution produced are qualitatively similar. Most of the vortices are bent towards a side evidencing a disparity in streamwise vorticity levels on either side of the vortex pair.
Figure 7.
Perturbation energy versus x, for two different inlet noise distributions but same initial amplitude levels; Ro = 0.3.
Such asymmetries are due to interactions originating from an Eckhaus instability4. The main difference between the two cases lies in the average wavelength produced; the flow field on the right side of Figure 8 (corresponding to the dotted curve in Figure n) presents consistently one less upwash region than the case on the left. Simple hand counting of the visible upflows in the cross-sections indicates that there are 11, 10 and 9 pairs for x equal to, respectively, 2.4, 2.8 and 3.2 (flow field on the right). Even this count is approximate, since some vortices do not have "companion".
184
Figure 8.
Contours of the streamwise velocity fields for the two cases of Figure 7 (the left side corresponds to the dotted curve in Figure 7). The plots are scaled correctly.
Starting with the same white noise but with different amplitudes, the energy of the disturbance decreases (Figure 9a); this occurs over an extent which is independent of the initial amplitudes.
Figure 9.
a. b.
Perturbation energy versus x, Ro = 0.3. Different initial amplitude levels. Collapse of the curves when a "virtual origin" is introduced.
185 The perturbation amplitudes chosen on the different velocity components are 1%, 4.2%, 10% and 33.3%, corresponding to curves 1, 2, 3 and 4, respectively. Thus, the boundary layer equations act as a linear filter between x = 1 and xo = 1.52 (vertical line in the figure) for whatever initial perturbation amplitude. Quasi-exponential growth follows xo according to A_ = e2a( V x 0 ),
(13)
where A is the amplitude of the energy perturbation, a is the linear growth rate, and xj > xo, i = 1, 2, 3, 4, is the streamwise location where the level A is achieved for any initial level Aoi. Thus, we find, using the fact that Aoi = 0.014, A02 = 0.233, A03 = 1.329 and A04 = 16.130 at the common minimum point of the three curves, xo = 1.52, the relations: xi -X3= 1.61 (xi -X2) xi - X4 - 2.50 (xi - x 2 )
(14)
This scenario is confirmed by sliding curve 2 horizontally a distance L until it coincides, for x large enough, with curve 1. When curve 3 is likewise shifted a distance 1.61 L to the right and curve 4 a distance 2.5 L, we find the nice collapse indicated in Figure 9b. The overlap of the four curves for some x range till the nonlinear saturation is a powerful result which indicates conclusively that a linear receptivity operation is actuated by this flow when subject to steady, random perturbations.
4.2
Effect of Rotation
The effect of rotation on this selection process is investigated from numerical experiments starting from the same random noise distribution and level as case 1 above, three cases have been compared: Ro = - 0.3, 0, and 0.3. In Figure 10 we have plotted isolines of the streamwise velocity for these three cases at different downstream positions x. As already shown, the vortices start growing at random locations in z, and vortices with different individual wavelengths and amplitudes coexist at each x. The average wavelength at the x-position where the vortices first appear can be obtained by simply counting the number of outflow regions; this gives A = 125 (for Ro = 0.3), 181 (Ro = 0) and 250 (Ro = -0.3). Because of the irregularity of the vortices, the value of the average wavelength is approximate to = ±10%. It is also clear (Figure 7) that different inlet noises would produce slightly different results. The average "natural" wavelengths obtained in this way are in agreement, although slightly lower, with those predicted by linear stability (shown as crosses in Figure 6). This confirms that, on the average, the linearly most amplified wavelength is selected when the flow is subjected to random inlet perturbations.
186
A notable difference between the three cases computed is the fact that A changes significantly for the positive rotation case (Figure 10c), in the range of x considered, because of merging of vortices.
Figure 10.
a. Isolines of the streamwise velocity for Ro = -0.3 at, from the top, x = 1.6, 2, 2.4, 2.8, 3.2. The isolines are spaced 0.1 apart. Note that, for comparison purposes, a small horizontal arrow has been drawn in correspondence to the edge of the undisturbed Blasius boundary layer, i.e. i\ = 5, at each x station. b. (next page) Ro = 0, same x values as Figure 10 a. c. (next page) Ro = 0.3, same x values as Figure 10 a. A tendency towards further merging events, for x > 3.2, manifests itself.
For Ro = 0.3, there are 13 vortex pairs at x = 2 (average A is 125), 12 at x = 2.4 (A = 138), 11 at x = 2.8 (A = 157) and 10 at x = 3.2 (A = 181). To a lesser degree this also occurs for Ro = 0, where only one merging event occurs for 2.2 < x < 2.4, whereas the average wavelength remains constant for the negative value of Ro. This points out the fact that the case Ro = 0.3 is more susceptible to an Eckhaus instability than Ro = -0.3; hence, in the nonlinear regime for Ro > 0 an Eckhaus criterion should be used to interpret the selected wavelength.
187
188 The vortex merging at positive Ro can be clearly observed both in the contours of u in Figure 10c and in the streamwise vorticity field of Figure 11. A clear example is the merging of two vortex pair near the center of the cross-section at x = 2.4 into one pair at x = 2.8; a similar phenomenon is also produced between x = 2.8 and 3.2. Both events are marked by vertical arrows in Figure 11 for easy identification. Typically, the merging involves two vortex pairs, where one pair is much stronger than its neighbor. The stronger pair "jumps" above the weaker one with a spanwise shift of half a wavelength. In so doing, at first one cell of the weaker pair is annihilated and secondly two co-rotating cells (originally belonging to two different pairs), placed one above the other, merge into one. The end result is one strong pair (central pair at x = 3.2) with the upwash region somewhat inclined, and a tilting of the neighboring vortices.
Figure 11.
Streamwise vorticity field in the cross-section at the same x values as Figure 10, Ro = 0.3. Isolines spacing are 0.03 (for x = 1.6), 0.6 (x = 2), 1.2 (x = 2.4) and 2 (x = 2.8, 3.2); zero lines are not drawn. The horizontal arrows mark the edge of the undisturbed Blasius boundary layer, i.e. r| = 5.
189 5.
Concluding remarks
In this paper, results for nonlinear spatially developing Gortler vortices subject to system rotation have been presented. Comparisons with recent nonparallel linear theories, using both local and nonlocal solution procedures, have also been shown. It is shown how positive rotation enhances the development of the vortices while negative rotation retards their spatial development. We have performed numerical experiments to study the natural formation of steady Gortler vortices in a long (along the span) box, subject to random inlet perturbation fields. Initially, the flow fields go through a linear receptivity phase which selects the modes to be amplified independently of the initial perturbation amplitude level. However, different inflow random noise distributions yield different vortical structures downstream, but approximately same average wavelengths. This indicates that the initial filtering phase is similar for different initial conditions. The natural average wavelengths obtained are very close to those predicted by the linear analysis. As the vortices develop into the strongly nonlinear stage, the wavelength increases due to merging of the most narrowly spaced vortices, a process related to the Eckhaus instability. This is seen most clearly at positive Ro, for which the nonlinear development of the vortices is enhanced. Reference 14 gives more details and results about this problem.
6.
Acknowledgements
We thank Professor I. L. Ryhming for initiating this collaborative effort. A.Z. has been supported during his stay at EPFL by an ERCOFTAC Visitor Grant. A.B. acknowledges the Swiss National Fund, grant no. 21-36035.92, for travel support associated with this research. This work was also supported by the Swedish Board of Technical Development (NUTEK), the Swedish Technical-Scientific Council (TFR) and an ERCOFTAC Visitor Grant, through which the stay of B.G.B.K. at the EPFL was made possible. Cray-2 computing time for this research was generously provided by the Service Informatique CentraleofEPFL.
7.
References
1.
R.F. Blackwelder, Analogies between transitional and turbulent boundary layers. Phys. Fluids 26, 2807, (1983).
2.
J.D. Swearingen and R.F. Blackwelder, The growth and breakdown of streamwise vortices in the presence of a wall. /. Fluid Mech. 182, 255, (1987).
3.
K. Lee and J.T.C. Liu, On the growth of mushroomlike structures in nonlinear spatially developing Gortler vortex flow. Phys. Fluids A 4, 95, (1992).
190 4.
Y. Guo and W.H. Finlay, Wavenumber selection and irregularity of spatially developing nonlinear Dean and Gortler vortices. J. Fluid Mech. 264, 1, (1994).
5.
J.M. Floryan and W.S. Saric, Wavelength selection and the growth of Gortler vortices. AIAAJ. 22, 1529, (1984).
6.
I. Tani, Production of longitudinal vortices in the boundary layer along a curved wall. J. Geophysical Research 67, 3075, (1962).
7.
H. Bippes and H. Gortler, Dreidimensionale Storungen in der Grenzschicht an einer konkaven Wand. Acfa Mechanica 14, 251, (1972).
8.
H. Bippes, Experimental study of the laminar-turbulent transition of a concave wall in a parallel flow. NASA TM. 75243, (1978).
9.
A. Zebib and A. Bottaro, Gortler vortices with system rotation: Linear theory. Phys. Fluids A 5, 1206, (1993).
10.
S.V. Patankar and D.B. Spalding, A calculation procedure for heat, mass and momentum transfer in three-dimensional parabolic flows. Int. J. Heat Mass Transfer 15, 1787,(1972).
11.
P. Hall, The linear development of Gortler vortices in growing boundary layers. J. Fluid Mech. 130,41,(1983).
12.
F.P. Bertolotti, Linear and nonlinear stability of boundary layers with streamwise varying properties. Ph.D. Thesis, The Ohio State University, (1991).
13.
B.G.B. Klingmann, Linear parabolic stability equations applied to threedimensional perturbations in a boundary layer. Rep. T-93-12. IMHEF-DME, Swiss Federal Institute of Technology, Lausanne, (1993).
14.
A. Bottaro, B.G.B. Klingmann and A. Zebib, Gortler Vortices with system rotation. Submitted to Theor. & Comp. Fluid Dyn. (1995).