Shanghai University
Article Information
- Xiangming XIONG, Jianjun TAO
- Lower bound for transient growth of inclined buoyancy layer
- Applied Mathematics and Mechanics (English Edition), 2017, 38(6): 779-796.
- http://dx.doi.org/10.1007/s10483-017-2202-8
Article History
- Received Apr. 15, 2016
- Revised Sep. 27, 2016
The buoyancy boundary layer above a heated inclined plate immersed in a stratified cold fluid was introduced by Prandtl[1] as a model of mountain and valley winds (see Fig. 1). When the temperature difference between the hot plate and the cold fluid at the same height is a constant, the basic flow of the inclined buoyancy layer is a parallel flow. Such a boundary layer also appears above the inclined bottom of an ice-covered lake[2].
![]() |
Fig. 1 Schematic of inclined buoyancy boundary layer, where heated plate is denoted by thick line |
|
When the inclination angle of the plate relative to the horizon is χ= 90°, the inclined buoyancy layer becomes a vertical buoyancy layer. Early linear stability analyses of the buoyancy layers were mainly concerned with two-dimensional normal modes[3-5], although Squire's theorem does not apply here[3]. Tao and Busse[6] studied the three-dimensional stability of the inclined buoyancy layer, and found that the most unstable modes in some region of the parameter space are not the two-dimensional Tollmien-Schlichting (TS) waves but the oblique rolls.
The linear stability analysis only provides a sufficient condition of hydrodynamic instability, while the simplest necessary condition of instability is given by the energy stability analysis (the energy method). The energy stability of the vertical buoyancy layer was studied by Dudis and Davis[7], but the energy stability of the inclined buoyancy layer is still unknown at present, and this is the first motivation of this paper.
When the Grashof number is less than the critical Grashof number of energy stability GrE, the energy of any disturbance decays monotonically, which precludes any transient disturbance energy growth, and the basic flow is therefore globally stable[8]. In the Grashof number range between GrE and the critical Grashof number of linear stability GrL, the disturbance energy is possible to grow initially, and then the disturbance may decay eventually according to the linear stability or initiate a subcritical transition with the help of nonlinear effects. The transient disturbance energy growth in parallel shear flows can be as large as O(1 000) at subcritical Reynolds numbers[9-11]. However, the maximum transient energy growth has never been evaluated in the inclined buoyancy layers, and this is the second motivation of the present study.
Both the critical Grashof number of the energy stability and the transient energy growth depend on the definition of the disturbance energy. The conventional energy method uses the kinetic energy of disturbance as the definition, while the generalized energy method defines a positive-definite Lyapunov functional as a generalized disturbance energy[12-13], and may shrink the difference between GrE and GrL. In the case of the Rayleigh-Bénard convection, the energy stability and the linear stability theories share the same critical Rayleigh number[14]. In this paper, we will use the generalized energy method to prove that the basic flow of the vertical buoyancy layer is globally stable to streamwise-independent disturbances.
In the following, we will introduce the mathematical model of the inclined buoyancy layer in Section 2, and extend the conventional energy stability analysis of the vertical buoyancy layer[7] to the case of the inclined buoyancy layer in Section 3. The maximum transient growth at subcritical Grashof numbers in the inclined buoyancy layer will be calculated in Section 4. In Section 5, we will study the linear and nonlinear stabilities of the basic flow to streamwise vortices in the vertical buoyancy layer. The conclusions will be summarized in Section 6.
2 Mathematical modelA schematic of the inclined buoyancy boundary layer is shown in Fig. 1. The coordinates in the streamwise, wall-normal, and spanwise directions are x*, y*, and z*, respectively. The asterisks denote dimensional variables. The coordinate in the −g direction is s*, where g is the gravitational acceleration. The fluid in the far field is stably stratified with a linear temperature profile T ∞ * (s*) = T ∞ * (0) + ζ*s*, where T ∞ * (0) is a reference temperature, and ζ* > 0 is the magnitude of the temperature gradient in the stably stratified fluid. The temperature at the surface of the plate is assumed to be Tw *(s*) = T ∞ * (s*) + ∆T*, where ∆T* is the constant temperature difference between the hot plate and the cold fluid at the same height s*.
The dimensional governing equations with the Oberbeck-Boussinesq approximation are[6]
![]() |
(1) |
![]() |
(2) |
![]() |
(3) |
![]() |
(4) |
![]() |
(5) |
where U* = (U*, V*, W*) is the velocity of the fluid. T* and P* are the temperature and the pressure of the fluid, respectively. Parameters ρ0, γ, υ, and κ are the reference density, the coefficient of thermal expansion, the kinematic viscosity, and the thermal diffusivity of the fluid, respectively.
Define the non-dimensional coordinates, time, velocity, temperature, and pressure as[6]
![]() |
where P ∞ *(s*) = P ∞ * (0) + ρ0g(0.5γζ*s*2 − s*), and
![]() |
The Grashof number and the Prandtl number are defined as
![]() |
(6) |
where χ is the inclination angle of the plate relative to the horizon.
Then, we have the non-dimensional governing equations and boundary conditions,
![]() |
(7) |
![]() |
(8) |
![]() |
(9) |
![]() |
(10) |
![]() |
(11) |
![]() |
(12) |
![]() |
(13) |
The basic flow is a parallel flow[1],
![]() |
(14) |
The non-dimensional governing equations and boundary conditions of disturbance (u, v, w, T') = (U, V, W, T) − (U0, 0, 0, T0) are[6]
![]() |
(15) |
![]() |
(16) |
![]() |
(17) |
![]() |
(18) |
![]() |
(19) |
![]() |
(20) |
![]() |
(21) |
where u, v, and w are the velocity components of disturbance corresponding to x, y, and z, respectively. T' and p are the temperature perturbation and the pressure perturbation, respectively.
3 Energy methodThe disturbance energy is defined as
![]() |
(22) |
where λ is a positive parameter to be determined. As a result,
![]() |
(23) |
where
![]() |
(u1, u2, u3) ≡ (u, v, w), and (x1, x2, x3) ≡ (x, y, z). The Einstein summation convention is used here. Only when λ=0.25PrGr2 will the energy maximum problem be well posed, because the integral
![]() |
where A is a positive parameter, then the disturbance is second-order continuously differentiable and divergence-free, and satisfies the boundary conditions. When A → + ∞ , we have
![]() |
Therefore, the integral
Using λ=0.25PrGr2 and introducing θ=0.5GrT', we obtain
![]() |
(24) |
and
![]() |
(25) |
As the coefficient of the third integral term in (23) vanishes, the inclination angle χ influences
Define the critical Grashof number of energy stability as
![]() |
(26) |
The maximum is searched among all divergence-free disturbances satisfying the boundary conditions, and is evidently positive. The critical Grashof number of the energy stability is always finite, because
![]() |
(27) |
and
![]() |
(28) |
When Gr < GrE, we have
When calculating the maximum in (26), we obtain the same Euler-Lagrange equations as in the energy method of the vertical buoyancy layer[7],
![]() |
(29) |
![]() |
(30) |
![]() |
(31) |
![]() |
(32) |
![]() |
(33) |
where μ and ψ(x, y, z) are the Lagrange multipliers. The minimum positive eigenvalue μ is just the critical Grashof number of the energy stability GrE.
In the following, we will introduce a discretization method different from the one used in Ref. [7] to solve the eigenvalue problem (29) -(33).
Expanding the disturbances in Fourier integrals
![]() |
and introducing the wall-normal vorticity
![]() |
(34) |
![]() |
(35) |
![]() |
(36) |
![]() |
(37) |
![]() |
(38) |
![]() |
(39) |
where
Assume
![]() |
(40) |
where the summation is over all integers m and n such that 0 ≤ n ≤ N, −n ≤ m ≤ n, and m+n is even. This discretization is an extension of the original method of descending exponentials, which was used in the linear stability analysis of the saline boundary layer near the surface of a porous medium[17]. The original method of descending exponentials takes the form of
Substituting (40) into (34) -(39) and neglecting terms that are O(exp(−(k +N + 1)y)) when y → + ∞ , the eigenvalue problem of ordinary differential equations (34) -(39) is reduced to a matrix eigenvalue problem, which is solved with the QZ function in MATLAB.
Given α and β, denoting the minimum positive eigenvalue of (34) -(39) by µ1(α, β), we have
![]() |
(41) |
For Pr=0.72, the minimum in (41) is achieved at (α, β)=(0, 0.485 69) (see Fig. 2). Actually, the minimum in (41) is always achieved at α=0 within numerical accuracy for 0≤ Pr≤ 2. Therefore, the least stable modes of the energy stability are streamwise vortices. This is also the case for the energy stability of plane Poiseuille flow[19-20] and plane Couette flow[21]. When Pr > 20, μ1 is almost independent of (β/α)[7].
![]() |
Fig. 2 Contours of minimum positive eigenvalue μ1(α, β) for Pr=0.72 |
|
Define the spanwise wavenumber of the least stable mode as β3D. Therefore, GrE = µ1(0, β3D). We also define the critical Grashof number of the energy stability for two-dimensional disturbances as[7]
![]() |
(42) |
where α2D is the streamwise wavenumber of the least stable two-dimensional disturbance in the energy method. In this section, we use N=15 in the discretization, which is accurate enough for the calculation of eigenvalues (see Table 1).
![]() |
We calculate the critical wavenumbers of the least stable two-dimensional and three-dimensional modes in the energy method for Prandtl numbers between 0 and 100 (see Fig. 3(a)). The critical wavenumbers do not coincide with the result of Dudis and Davis[7], because the critical wavenumbers listed in their paper were only accurate to 0.05, which was implied in their later work[22]. The critical Grashof numbers they calculated nevertheless had only 5% relative errors compared with our results (see Fig. 3(b)). Not demanding better accuracy, they numerically integrated the eigenvalue equations only from y=0 to y = Ly = 8. To check accuracy of the present method, we use their method with Ly = 20 and identify the eigenvalues in Table 1 to be between 39.111 626 150 6 and 39.111 626 150 7, and between 54.383 008 825 and 54.383 008 826. Therefore, our algorithm is reliable, and our results are accurate enough. Note that the Grashof number in this paper is equivalent to 2R in Ref. [7].
![]() |
Fig. 3 Critical wavenumbers and critical Grashof numbers of energy stability and linear stability as functions of Pr |
|
We also calculate the critical Grashof numbers of the linear stability (GrL) for angles of inclination 30° and 90° with the Chebyshev-Tau method[23] (see Fig. 3(b)). The concern of Dudis and Davis[7] is confirmed, i.e., the critical Grashof numbers of the energy stability and the linear stability do not coincide (GrE < GrL). Actually, Dudis and Davis were not sure about it, because the only available linear stability result at that time concerned merely two-dimensional disturbances (β=0) [3]. On the contrary, the GrL curves plotted here concern all two-dimensional and three-dimensional disturbances (α2+β2 > 0).
When Gr < GrE, the basic flow is stable to all disturbances with finite amplitudes; when Gr > GrL, there exists at least one infinitesimal disturbance increasing exponentially. When GrE < Gr < GrL, the disturbance energy is able to increase at the beginning for some initial disturbances, and then the linear stability theory implies that disturbances will eventually decay if nonlinear effects are negligible.
We use SIMSON[24] to simulate the nonlinear evolution of disturbances in the inclined buoyancy layer for χ=30° and Pr=0.72. The discretization in the streamwise and spanwise directions uses a two-dimensional Fourier series expansion. Therefore, the solution is periodic in these directions. A Chebyshev series expansion is used in the wall-normal direction. The simulation is performed in a computational domain of size (Lx, Ly, Lz) = (4π, 30, 2π/β) using (64, 129, 64) spectral modes. The initial condition is the least stable normal mode in the energy method with (α, β)=(0, 0.485 69). Therefore, the disturbance energy will grow initially if Gr > GrE = 39.11. In this simulation, Lx can be arbitrary because the normal mode is independent of x. The evolution of the average disturbance energy E' = E/(LxLz) is shown in Fig. 4(a). When t=0, the growth rates of the average disturbance energy in the direct numerical simulation (DNS) are in accord with
![]() |
Fig. 4 Average disturbance energy and its growth rate in DNS (χ=30° and Pr=0.72) |
|
The contours of the wall-normal velocity v of the least stable normal mode in the energy method when Pr=0.72 are plotted in the zy-plane (see Fig. 5(a)). The profile is scaled so that the average disturbance energy E'=1. The contour levels are ± 0.1, ± 0.2, …, and the contours of negative values are plotted with dashed lines. The contours of v=0 are omitted for clarity.
![]() |
Fig. 5 Contours of v of normal modes corresponding to GrE, GrE, 2D, and GrL when χ=30° and Pr=0.72 |
|
We also plot the contours of v of the least stable two-dimensional normal mode in the energy method when Pr=0.72 in the xy-plane as shown in Fig. 5(b), and plot the contours of v of the least stable normal mode in the linear stability theory when χ=30° and Pr=0.72 in Fig. 5(c). Both profiles are scaled so that E'=1, and have maximal v at y ≈ 2.5. The profile of the two-dimensional least stable normal mode in the energy method, however, extends farther into the far field.
4 Transient energy growthWhen GrE < Gr < GrL, the disturbance energy is able to grow initially, and then the disturbance may be sustained by the nonlinear effects or decay according to the linear stability theory. The transient energy growth exists even when the nonlinear effects are negligible, because the growth rate of the disturbance energy
Formally, the solution to the linearized Navier-Stokes equation
![]() |
(43) |
where E is the disturbance energy defined in (24). The maximum growth for all time is defined as[10]
![]() |
(44) |
Define the maximum growth for all time and wavenumbers as[11]
![]() |
(45) |
where (αmax, βmax) are the wavenumbers of the optimal disturbance.
To calculate the energy growth function, we first discretize the linearized Navier-Stokes equation with a Chebyshev polynomial expansion in the interval [0, 30] in the wall-normal direction and with a two-dimensional Fourier series expansion in the streamwise and spanwise directions. After that the linear operator L is approximated by a finite-dimensional matrix L', the eigenvalues and the eigenvectors of which are solved with the QZ function in MATLAB. In the end, the energy growth function is calculated with the singular value decomposition of the matrix eL't. The detail of the procedure was described by Reddy and Henningson[11].
We plot the energy growth function G(t) for (α, β)=(0.4, 0.4), χ=90°, Pr=0.72, and Gr=200 in Fig. 6(a). The Grashof number is between GrE = 39.1 and GrL = 202.35 for χ=90° and Pr=0.72. The maximum growth for all time is Gmax = G(0.730) =11.19, i.e., the disturbance energy is able to grow to 11.19 times as large as the initial disturbance energy. Note that the points in the curve G (t)correspond to different initial disturbances. The contours of the wall-normal velocity v of the initial disturbance corresponding to Gmax are plotted in Fig. 6(b). The profile is scaled so that the average disturbance energy E'=1. The contour levels are ± 0.1, ± 0.2, …, and the contours of negative values are plotted with dashed lines. The contours of v=0 are omitted for clarity.
![]() |
Fig. 6 Energy growth function and initial disturbance corresponding to Gmax for (α, β)=(0.4, 0.4), χ=90°, Pr=0.72, and Gr=200 |
|
The calculation of the maximum transient growth is based on the eigenvalue decomposition of the linearized Navier-Stokes equation. We examine the validity of this algorithm by performing DNS with SIMSON[24] and comparing the evolution of the disturbance energy in the DNS with that predicted by E(t) = ||v(t)||2/2 = ||eL'tv(0)||2/2. The simulation is performed in a computational domain of size (Lx, Ly, Lz) = (2π/α, 30, 2π/β) using (64, 513, 64) spectral modes, where (α, β)=(0.4, 0.4). The initial disturbances in the DNS are given by scaling the disturbance in Fig. 6(b) to E'=0.1 and E'=10.
When the initial average disturbance energy is E'=0.1, the evolution of the disturbance energy agrees well with that predicted by the eigenvalue decomposition of the linearized Navier-Stokes equation (see Fig. 7(a)). The xz-plane averaged disturbance velocity and temperature
![]() |
Fig. 7 DNS verification for (α, β)=(0.4, 0.4), χ=90°, Pr=0.72, and Gr=200 |
|
When the initial average disturbance energy is E'=10, although the three-dimensional flow structure is similar to that of E'(0) =0.1 (see Fig. 7(e)), the amplitudes of the xz-plane averaged disturbance velocity and temperature are comparable to those of the basic flow (see Fig. 7(c)). For example, the relative modification of the basic flow uave/U0 is about 6% at y=π/4, where U0 achieves its maximum, and is about 72% at y=2.6, where uave achieves its maximum. The xz-plane averaged flow differs from the basic flow, which is the base of the linear transient growth analysis. Therefore, the evolution of E' in the case E'(0) =10 is diverted from the linear theoretical prediction (see Fig. 7(a)).
We summarize the maximum growth for all time Gmax and the corresponding tmax in the αβ-plane for χ=90°, Pr=0.72, and Gr=200 in Figs. 8(a) and 8(b). The maximum growth for all time and wavenumbers is S=11.25, which is achieved at (αmax, βmax)=(0.379, 0.449) and tmax = 0.708 2.
![]() |
Fig. 8 Maximum growth for all time for χ=90°, Pr=0.72, and Gr=200 |
|
However, the optimal disturbances that have the maximum growth in the whole wavenumber space are not always three-dimensional oblique rolls. We calculate the maximum growth for all time and wavenumbers S(χ, Gr, Pr) in the χGr-plane when Pr=0.72 (see Fig. 9(a)), and find that the optimal disturbances are streamwise vortices when χ≤ 50° (see Figs. 9(b) and 9(c)). In Fig. 9(a), the upper and lower thick lines represent GrL and GrE, respectively. According to the definitions (43) and (44), Gmax = + ∞ when Gr > GrL, and Gmax = 1 when Gr ≤ GrE. Therefore, we only plot the contours of S when GrE ≤ Gr ≤ GrL. When Pr=0.72, the maximum growth in the disturbance energy over all angles of inclination and GrE ≤ Gr ≤ GrL is S=11.62, which is achieved at χ=90° and Gr = GrL = 202.35.
![]() |
Fig. 9 Maximum growth for all time and wavenumbers S and optimal wavenumbers for Pr=0.72 |
|
In the vertical buoyancy layer, the least stable modes of the energy stability are streamwise vortices, which was already known by Dudis and Davis[7]. However, in the following, we will prove that the basic flow of the vertical buoyancy layer is linearly stable to all streamwise vortices no matter how large the Grashof number is.
Theorem 1 The basic flow of the vertical buoyancy layer is linearly stable to all streamwise vortices.
Substituting the temporal normal mode of linear stability
![]() |
(46) |
![]() |
(47) |
![]() |
(48) |
![]() |
(49) |
where
![]() |
Then, we have either λ < 0 or
![]() |
Eliminating
![]() |
(50) |
![]() |
(51) |
Multiplying (50) by the complex conjugate of
![]() |
Remark 1 This proof is inspired by the proof of "the solutions to the Squire equation are always damped"[8].
Although streamwise vortices are always damped in the linear stability theory, the critical Grashof numbers of the energy stability (GrE) of streamwise vortices are finite for all Prandtl numbers in the vertical buoyancy layer (see Fig. 3(b)). When Gr > GrE, although streamwise vortices may grow transiently, they will decay eventually if the nonlinear effects are negligible. Because GrE depends on the definition of the disturbance energy (24), we can increase GrE by defining other positive-definite Lyapunov functionals as the generalized disturbance energy. With this generalized energy method, we will prove in the following that the basic flow of the vertical buoyancy layer is stable to finite-amplitude streamwise-independent disturbances (including the streamwise vortices in the linear stability analysis).
Theorem 2 The basic flow of the vertical buoyancy layer is stable to all streamwise-independent disturbances.
For a streamwise-independent disturbance (u, θ), we have
![]() |
(52) |
![]() |
(53) |
where
The preceding discussion inspires us to define the weighted disturbance energy
![]() |
(54) |
where σ is a positive parameter that will be determined later. As a result,
![]() |
(55) |
Notice that
![]() |
(56) |
where the maxima are searched among all streamwise-independent divergence-free disturbances satisfying the boundary conditions. According to the definition (26), the right hand side of (56) is no larger than GrE−1. Therefore, we have
![]() |
(57) |
Furthermore, from (55) and with the same variable substitution as in (56), we have
![]() |
(58) |
For any 0 < σ < (GrE/Gr)2, we have
Remark 2 Joseph and Hung[25] have proven that the parallel flows with prescribed boundary values (such as the plane Poiseuille flow and the plane Couette flow) are stable to all streamwise-independent disturbances. Their proof relied on the Poincaré inequality in a finite interval,
To compare the different energy stabilities defined by (24) and (54), we simulate the evolution of the least stable mode in the energy method (α, β)=(0, 0.485 69) in the vertical buoyancy layer with SIMSON[24], and plot the average disturbance energy E' and
![]() |
(59) |
![]() |
Fig. 10 Average disturbance energy E' (solid lines) and ![]() |
|
and
![]() |
(60) |
where Ω = [0, Lx] × [0, 30]× [0, Lz] with a resolution of (64, 513, 64) spectral modes. We choose Lz = 2π/0.485 69, whereas Lx can be arbitrary because the disturbance is independent of x.
Because we have known GrE = 39.11 when Pr=0.72, we choose σ=0.03 < (GrE/Gr)2 in the definition (54), which leads to the above definition of
In this paper, we prove that the inclined buoyancy layer and the vertical buoyancy layer have the same critical Grashof number of energy stability. However, if the Grashof number is defined as Gr' = gγd*3∆T*/ν2 = Gr/ sin χ[4] instead of (6), the new critical Grashof number of the energy stability satisfies GrE'(χ) = GrE' (90°)/ sin χ. Furthermore, we take advantage of the exponential form of the basic flow, using the method of descending exponentials to discretize the eigenvalue equations in the energy stability analysis. The critical Grashof numbers and the critical wavenumbers are calculated with better accuracy than those listed in Ref. [7].
In the band between the energy stability curve and the linear stability curve in Fig. 3(b), the disturbance energy is able to increase transiently[7]. We calculate the transient growth in the disturbance energy for Pr=0.72 based on the linearized Navier-Stokes equation. When χ ≤ 50°, the optimal disturbances that have the maximum energy growth are streamwise vortices. This conclusion also holds for 55° ≤ χ ≤ 90° and Gr≤ 110. However, when 55° ≤ χ ≤ 90° and Gr ≥ 120, oblique rolls with nonzero streamwise and spanwise wavenumbers are the optimal disturbances. The maximum growth in the disturbance energy over all angles of inclination at subcritical Grashof numbers is S=11.62 for χ=90°.
In the vertical buoyancy boundary layer, we prove that the basic flow is always linearly stable to streamwise vortices. We also define the weighted disturbance energy (54) instead of the original disturbance energy (24) to prove that the basic flow of the vertical buoyancy layer is stable to any finite-amplitude streamwise-independent disturbance. However, this definition of weighted energy cannot be extended either to streamwise-dependent disturbances or to the inclined buoyancy layer.
The energy stability of the inclined buoyancy layer is reminiscent of that of parallel shear flows, where the critical Reynolds numbers of energy stability are less than those of linear stability. Although continuous turbulence exists even when Re < ReL in plane Poiseuille flows[26], whether nonlinear effects are able to maintain a finite-amplitude disturbance in the inclined buoyancy boundary layer when GrE < Gr < GrL is worthy of further study.
Acknowledgements The work was carried out at National Supercomputer Center in Tianjin, and the calculations were performed on TianHe-1(A).[1] | Prandtl, L. Essentials of Fluid Dynamics, Blackie and Son, London, 422-425 (1952) |
[2] | Malm, J. Bottom buoyancy layer in an ice-covered lake. Water Resources Research, 34, 2981-2993 (1998) doi:10.1029/98WR01904 |
[3] | Gill, A. E. and Davey, A. Instabilities of a buoyancy-driven system. Journal of Fluid Mechanics, 35, 775-798 (1969) doi:10.1017/S0022112069001431 |
[4] | Iyer, P. A Instabilities in buoyancy-driven boundary-layer flows in a stably stratified medium. Boundary-Layer Meteorology, 5, 53-66 (1973) doi:10.1007/BF02188311 |
[5] | McBain, G. D., Armfield, S. W., and Desrayaud, G. Instability of the buoyancy layer on an evenly heated vertical wall. Journal of Fluid Mechanics, 587, 453-469 (2007) |
[6] | Tao, J. and Busse, F. H. Oblique roll instability in inclined buoyancy layers. European Journal of Mechanics B/Fluids, 28, 532-540 (2009) doi:10.1016/j.euromechflu.2009.01.002 |
[7] | Dudis, J. J. and Davis, S. H. Energy stability of the buoyancy boundary layer. Journal of Fluid Mechanics, 47, 381-403 (1971) doi:10.1017/S0022112071001113 |
[8] | Schmid, P. J. and Henningson, D. S. Stability and Transition in Shear Flows, Springer-Verlag, New York (2001) |
[9] | Butler, K. M. and Farrell, B. F. Three-dimensional optimal perturbations in viscous shear flow. Physics of Fluids A: Fluid Dynamics, 4, 1637-1650 (1637) |
[10] | Reddy, S. C., Schmid, P. J., and Henningson, D. S. Pseudospectra of the Orr-Sommerfeld operator. SIAM Journal on Applied Mathematics, 53, 15-47 (1993) doi:10.1137/0153002 |
[11] | Reddy, S. C. and Henningson, D. S. Energy growth in viscous channel flows. Journal of Fluid Mechanics, 252, 209-238 (1993) doi:10.1017/S0022112093003738 |
[12] | Galdi, G. P. and Padula, M. A new approach to energy theory in the stability of fluid motion. Archive for Rational Mechanics and Analysis, 110, 187-286 (1990) doi:10.1007/BF00375129 |
[13] | Straughan, B. The Energy Method, Stability, Nonlinear Convection, 2nd ed. , Springer-Verlag, New York (2004) |
[14] | Joseph, D. D. On the stability of the Boussinesq equations. Archive for Rational Mechanics and Analysis, 20, 59-71 (1965) |
[15] | Joseph, D. D. Stability of Fluid Motions Ⅱ, Springer-Verlag, Berlin, 29-33 (1976) |
[16] | Dongarra, J. J., Straughan, B., and Walker, D. W. Chebyshev tau-QZ algorithm methods for calculating spectra of hydrodynamic stability problems. Applied Numerical Mathematics, 22, 399-434 (1996) doi:10.1016/S0168-9274(96)00049-9 |
[17] | Van Duijn, C. J. , Wooding, R. A. , and van der Ploeg, A. Stability Criteria for the Boundary Layer Formed by Throughflow at a Horizontal Surface of a Porous Medium: Extensive Version, Eindhoven University of Technology, Eindhoven (2001) |
[18] | Hardy, G. H. and Riesz, M. The General Theory of Dirichlet 's Series, Cambridge University Press, Cambridge (1915) |
[19] | Joseph, D. D. and Carmi, S. Stability of Poiseuille flow in pipes, annuli, channels. Quarterly of Applied Mathematics, 26, 575-599 (1969) doi:10.1090/qam/1969-26-04 |
[20] | Busse, F. H. Bounds on the transport of mass and momentum by turbulent flow between parallel plates. Zeitschrift für angewandte Mathematik und Physik, 20, 1-14 (1969) doi:10.1007/BF01591113 |
[21] | Joseph, D. D. Nonlinear stability of the Boussinesq equations by the method of energy. Archive for Rational Mechanics and Analysis, 22, 163-184 (1966) |
[22] | Dudis, J. J. and Davis, S. H. Energy stability of the Ekman boundary layer. Journal of Fluid Mechanics, 47, 405-413 (1971) doi:10.1017/S0022112071001125 |
[23] | Orszag, S. A. Accurate solution of the Orr-Sommerfeld stability equation. Journal of Fluid Mechanics, 50, 689-703 (1971) doi:10.1017/S0022112071002842 |
[24] | Chevalier, M. , Schlatter, P. , Lundbladh, A. , and Henningson, D. S. SIMSON: A Pseudo-Spectral Solver for Incompressible Boundary Layer Flows, KTH Mechanics, Stockholm (2007) |
[25] | Joseph, D. D. and Hung, W. Contributions to the nonlinear theory of stability of viscous flow in pipes and between rotating cylinders. Archive for Rational Mechanics and Analysis, 44, 1-22 (1971) |
[26] | Patel, V. C. and Head, M. R. Some observations on skin friction and velocity profiles in fully developed pipe and channel flows. Journal of Fluid Mechanics, 38, 181-201 (1969) doi:10.1017/S0022112069000115 |